ArticlePDF Available

Global Climate Impacts of Fixing the Southern Ocean Shortwave Radiation Bias in the Community Earth System Model (CESM)

Authors:

Abstract and Figures

A large, long-standing, and pervasive climate model bias is excessive absorbed shortwave radiation (ASR) over the midlatitude oceans, especially the Southern Ocean. This study investigates both the underlying mechanisms for and climate impacts of this bias within the Community Earth System Model, version 1, with the Community Atmosphere Model, version 5 [CESM1(CAM5)]. Excessive Southern Ocean ASR in CESM1(CAM5) results in part because low-level clouds contain insufficient amounts of supercooled liquid. In a present-day atmosphere-only run, an observationally motivated modification to the shallow convection detrainment increases supercooled cloud liquid, brightens low-level clouds, and substantially reduces the Southern Ocean ASR bias. Tuning to maintain global energy balance enables reduction of a compensating tropical ASR bias. In the resulting preindustrial fully coupled run with a brighter Southern Ocean and dimmer tropics, the Southern Ocean cools and the tropics warm. As a result of the enhanced meridional temperature gradient, poleward heat transport increases in both hemispheres (especially the Southern Hemisphere), and the Southern Hemisphere atmospheric jet strengthens. Because northward cross-equatorial heat transport reductions occur primarily in the ocean (80%), not the atmosphere (20%), a proposed atmospheric teleconnection linking Southern Ocean ASR bias reduction and cooling with northward shifts in tropical precipitation has little impact. In summary, observationally motivated supercooled liquid water increases in shallow convective clouds enable large reductions in long-standing climate model shortwave radiation biases. Of relevance to both model bias reduction and climate dynamics, quantifying the influence of Southern Ocean cooling on tropical precipitation requires a model with dynamic ocean heat transport.
No caption available
… 
No caption available
… 
No caption available
… 
No caption available
… 
No caption available
… 
Content may be subject to copyright.
Journal of Climate
EARLY ONLINE RELEASE
This is a preliminary PDF of the author-produced
manuscript that has been peer-reviewed and
accepted for publication. Since it is being posted
so soon after acceptance, it has not yet been
copyedited, formatted, or processed by AMS
Publications. This preliminary version of the
manuscript may be downloaded, distributed, and
cited, but please be aware that there will be visual
differences and possibly some content differences
between this version and the final published version.
The DOI for this manuscript is doi: 10.1175/JCLI-D-15-0358.1
The final published version of this manuscript will replace the
preliminary version at the above DOI once it is available.
If you would like to cite this EOR in a separate work, please use the following full
citation:
Kay, J., C. Wall, V. Yettella, B. Medeiros, C. Hannay, P. Caldwell, and C. Bitz,
2016: Global climate impacts of fixing the Southern Ocean shortwave radiation
bias in the Community Earth System Model (CESM). J. Climate.
doi:10.1175/JCLI-D-15-0358.1, in press.
$PHULFDQ0HWHRURORJLFDO6RFLHW\
AMERICAN
METEOROLOGICAL
SOCIETY
1
1
Global climate impacts of fixing the Southern Ocean shortwave radiation bias in the
2
Community Earth System Model (CESM)
3
4
Jennifer E. Kay1, Casey Wall2, Vineel Yettella1, Brian Medeiros3, Cecile Hannay3, Peter
5
Caldwell4, and Cecilia Bitz2,
6
7
1Cooperative Institute for Research in Environmental Sciences (CIRES) and Department of
8
Atmospheric and Oceanic Sciences (ATOC), University of Colorado at Boulder
9
2 Department of Atmospheric Sciences, University of Washington - Seattle
10
3 Climate and Global Dynamics, National Center for Atmospheric Research
11
4 Livermore National Lab, Department of Energy
12
13
Corresponding Author:
14
Jennifer E. Kay, Jennifer.E.Kay@colorado.edu,
15
CIRES/ATOC, University of Colorado Boulder
16
216 UCB, Boulder, CO 80309
17
Phone: 303-492-6289
18
19
Journal of Climate
20
Revised March 17, 2016
21
Submitted May 18, 2015
22
Manuscript (non-LaTeX) Click here to download Manuscript (non-LaTeX)
Kay_JClim_SouthernOceanASR_revised_March17,2016_text,fig
2
ABSTRACT:
23
A large, long-standing, and pervasive climate model bias is excessive absorbed shortwave
24
radiation (ASR) over the mid-latitude oceans, especially the Southern Ocean. This study
25
investigates both the underlying mechanisms for and climate impacts of this bias within the
26
Community Earth System Model with the Community Atmosphere Model version 5 (CESM-
27
CAM5). Excessive Southern Ocean ASR in CESM-CAM5 results in part because low-level
28
clouds contain insufficient amounts of supercooled liquid. In a present-day atmosphere-
29
only run, an observationally motivated modification to the shallow convection detrainment
30
increases supercooled cloud liquid, brightens low-level clouds, and substantially reduces
31
the Southern Ocean ASR bias. Tuning to maintain global energy balance enables reduction
32
of a compensating tropical ASR bias. In the resulting pre-industrial fully coupled run with a
33
brighter Southern Ocean and dimmer Tropics, the Southern Ocean cools and the Tropics
34
warm. As a result of the enhanced meridional temperature gradient, poleward heat
35
transport increases in both hemispheres (especially the Southern Hemisphere) and the
36
Southern Hemisphere atmospheric jet strengthens. Because Northward cross-equatorial
37
heat transport reductions occur primarily in the ocean (80%) not the atmosphere (20%), a
38
proposed atmospheric teleconnection linking Southern Ocean ASR bias reduction and
39
cooling with northward shifts in tropical precipitation has little impact. In summary,
40
observationally motivated supercooled liquid water increases in shallow convective clouds
41
enable large reductions in long-standing climate model shortwave radiation biases. Of
42
relevance to both model bias reduction and climate dynamics, quantifying the influence of
43
Southern Ocean cooling on tropical precipitation requires a model with dynamic ocean heat
44
transport.
45
46
3
1. Motivation
47
Excessive absorbed shortwave radiation (ASR) over the mid-latitude oceans is a
48
ubiquitous, large, and long-standing bias in both climate models and reanalyses (Trenberth
49
and Fasullo, 2010; Hwang and Frierson 2013). ASR biases are largest over the Southern
50
Ocean, where local differences between satellite-observed and modeled ASR often reach
51
10s of Wm-2. In many models, insufficient cloud optical depth explains the excessive mid-
52
latitude ocean ASR. Simply put, the model clouds are not bright enough. Studies using
53
cyclone compositing have shown that climate model ASR biases are largest in the post cold-
54
front regions of mid-latitude cyclones where low-topped shallow convective clouds are the
55
dominant cloud type (Bodas Salcedo et al. 2014; Williams et al. 2013; Bodas Salcedo et al.
56
2012).
57
58
Recognizing that large model radiation errors may have a profound influence on the
59
simulated climate, this study has two goals. First, we aim to use observations to motivate
60
improvements in climate model physics and reduce mid-latitude ocean ASR bias. Second,
61
we aim to assess the impacts of reduced mid-latitude ocean ASR bias on global climate. To
62
achieve these goals, we employ a widely used global coupled climate model: the
63
Community Earth System Model with Community Atmosphere Model version 5 (CESM-
64
CAM5) (Hurrell et al. 2013). Like most climate models of its class, CESM-CAM5 has
65
excessive mid-latitude ocean ASR due to dim clouds that do not scatter enough incoming
66
sunlight back to space (Kay et al. 2012a).
67
68
4
Accomplishing the first goal of this study requires identifying and removing a
69
deficiency in the model equations used to predict mid-latitude ocean cloud properties.
70
Satellite and in-situ observations show ubiquitous supercooled cloud liquid in low-level
71
clouds over the mid-latitude oceans, especially the Southern Ocean (e.g., Hu et al. 2010;
72
Morrison et al. 2011; Huang et al. 2012; Cesana and Chepfer 2013; Chubb et al. 2013;
73
Bodas-Salcedo et al. 2015). Even close to the Antarctic continent during winter, total
74
glaciation in low clouds over the Southern Ocean is rare (e.g., Morrison et al. 2011; Huang
75
et al. 2012). Supercooled cloud liquid exerts a strong control on cloud radiative effects in
76
both models (e.g., Kay et al. 2014; Forbes and Ahlgrimm 2014) and observations (e.g.,
77
Shupe et al. 2004; Bodas-Salcedo et al. 2015).
78
79
Numerical weather and climate models struggle to reproduce observations of
80
ubiquitous supercooled liquid-containing clouds in the extratropical atmosphere (e.g.,
81
Cesana and Chepfer, 2013; Forbes and Ahlgrimm 2014; Gettelman et al. 2015). The
82
processes that produce and remove supercooled cloud liquid are not explicitly resolved in
83
these models because they occur at a scale that is much smaller than the model resolution.
84
But beyond this scale challenge, it is important to realize that the ubiquitous presence of
85
supercooled liquid is in itself surprising, and not fully understood (Morrison et al. 2012). It
86
is well known that when ice and supercooled liquid co-exist, the ice grows at the expense of
87
the liquid by the Wegener-Bergeron-Findeisen mechanism (Wegener 1911; Bergeron
88
1935; Findeisen 1939). Indeed, the turbulent and microphysical processes that enable
89
supercooled liquid to persist in the atmosphere despite the Wegener-Bergeron-Findeisen
90
mechanism are complex, coupled, and enigmatic.
91
5
Motivated by observations of ubiquitous supercooled liquid in mid-latitude oceanic
92
clouds and by knowing that cloud liquid water content exerts a strong control on cloud
93
radiative properties, we propose a hypothesis: mid-latitude ocean clouds are not bright
94
enough in climate models because the modeled clouds contain insufficient amounts of
95
supercooled liquid water. To test this hypothesis, we change the model physics in CESM-
96
CAM5 to allow more supercooled liquid water in shallow convective clouds. In support of
97
our hypothesis, increasing supercooled liquid in shallow convective clouds makes those
98
clouds brighter and substantially reduces the excessive mid-latitude oceanic ASR bias,
99
especially over the Southern Ocean.
100
101
Having introduced our first goal, we next introduce and motivate our second goal:
102
documenting the influence of mid-latitude ocean ASR bias reduction on the global climate
103
system. Given the large magnitude of the ASR biases we aim to reduce, we expect bias
104
reduction will produce changes in global energy budgets and in atmospheric and oceanic
105
circulation. In particular, we investigate the possibility that brightening mid-latitude clouds
106
will lead to a tropical precipitation response via an atmospheric teleconnection. As
107
described in recent review papers by Chiang and Friedman (2012) and Schneider et al.
108
(2014), the central idea behind this atmospheric teleconnection is simple: a hemispheric
109
asymmetry in heating or cooling shifts tropical rainfall and associated atmospheric
110
circulation towards the relatively warmed hemisphere. Especially pertinent to this study,
111
Hwang and Frierson (2013) invoke this atmospheric teleconnection when hypothesizing
112
that fixing the Southern Ocean ASR climate model bias will preferentially cool the Southern
113
Hemisphere, and shift tropical precipitation northward to ameliorate another long-
114
6
standing climate model bias: the double Intertropical Convergence Zone (ITCZ) bias (e.g.,
115
Lin 2007). We test this proposed extratropical-tropical teleconnection in a hierarchy of
116
model configurations with varying degrees of atmosphere-ocean coupling. As we will show,
117
dynamic ocean heat transport mutes the atmospheric teleconnection linking Southern
118
Ocean cooling with northward ITCZ shifts.
119
120
2. Model and Experiments
121
122
a. Mid-latitude oceanic clouds in CESM-CAM5
123
We use a single climate model for our numerical experiments: CESM-CAM5. All of
124
our numerical experiments were done with the CESM Large Ensemble project code base
125
(Kay et al. 2015), which uses CAM version 5.2. CESM-CAM5 is a state-of-the-art global
126
coupled climate model that participated in Coupled Model Intercomparison Project version
127
5 (CMIP5, Taylor et al. 2012). A full description of CESM-CAM5 and its capabilities can be
128
found in a special collection of the J. Climate. Of particular relevance to this study are the
129
representation of double-moment cloud microphysics (Morrison and Gettelman 2008) and
130
shallow convection (Park and Bretherton 2009). A description of the integration of cloud
131
processes in CAM5 can be found in Park et al. 2014.
132
133
To introduce mid-latitude oceanic regions in CESM-CAM5, Figure 1a-c shows the
134
atmospheric circulation and cloud distribution from 30 to 70 °S for the annual mean,
135
Southern hemisphere winter (JJA), and Southern hemisphere summer (DJF). In the annual
136
mean, ascent from 50-70 °S associated with the mid-latitude stormtrack leads to relatively
137
7
large cloud fractions from the surface into the free troposphere, with the largest cloud
138
fractions below 700 mb. From 30 to 50 °S, subsidence associated with the descending
139
branch of the Hadley circulation leads to relatively small cloud fractions throughout the
140
troposphere and relatively shallow low cloud tops. Seasonal changes in insolation shift the
141
entire atmospheric circulation pattern poleward in DJF and equatorward in JJA, but the
142
general nature of the circulation and corresponding cloud distributions are seasonally
143
invariant.
144
145
In addition to cloud fraction, cloud properties (e.g., phase, temperature, particle
146
size) exert a strong control on cloud radiative effects. For example, Southern Ocean total
147
cloud fractions within CAM5 are within 0.10 of satellite observations, yet CAM5 still has
148
large Southern Ocean ASR biases indicative of cloud properties biases (Kay et al. 2012a).
149
Motivated by the need to understand cloud properties and specifically cloud water content,
150
Figure 1d-i shows the vertical distribution of Southern Ocean cloud water content in
151
CESM-CAM5 in the annual mean, DJF, and JJA. At all times of the year, the vast majority of
152
the CESM-CAM5 low-level (pressure > 700 mb) cloud water is liquid, not ice. Supercooled
153
liquid dominates between 268 K and 273 K. Below 268 K, the cloud water content is a mix
154
of water and ice. Observations show supercooled liquid dominance down to 253 K (Hu et
155
al. 2010; Morrison et al. 2010; Cesana and Chepfer 2013; Chubb et al. 2013). Indeed,
156
comparisons with satellite observations (Cesana and Chepfer 2013; O’Dell et al. 2008)
157
show that CAM5 clouds have too much ice and insufficient supercooled cloud liquid over
158
the Southern Ocean (Kay et al. 2016). In summary, CESM-CAM5 contains insufficient
159
supercooled cloud liquid at temperatures below 268 K.
160
8
161
Having identified that CESM-CAM5 clouds have insufficient supercooled cloud
162
liquid, how can this cloud phase bias be fixed? Cloud properties in climate models are
163
controlled by interactions between the grid-scale processes such as condensation and
164
evaporation, and the parameterized physics most notably the cloud microphysics and
165
convection. The dominant processes controlling modeled clouds can be identified using
166
tendencies, namely the change in cloud mixing ratio per unit time resulting from the
167
physical processes represented by the model.
168
169
Figure 2 shows the annual mean cloud liquid tendencies in both the subsidence
170
(30-50 °S) and ascent (50-70 °S) regions for CESM-CAM5 moist physical processes. Over
171
both Southern Ocean regions, detrainment from the shallow convection parameterization
172
(SHDLFLIQ) is the primary source of low-level cloud condensate. Once liquid cloud forms, it
173
is depleted by precipitation in the parameterized microphysics (MPDW2P), grid-scale
174
evaporation also known as the macrophysics” (MACPDLIQ), and to a lesser extent by
175
convective transport/mixing (entrainment of dry air) in the parameterized shallow
176
convection (CMFDLIQ). The conversion of cloud liquid to precipitation (MPDW2P)
177
primarily produces rain but also some snow in the ascent regime (not shown). Conversion
178
of cloud liquid to ice (MPDW2I), which occurs primarily via the Wegener-Bergeron-
179
Findeisen process (not shown), is a secondary sink for cloud liquid in the ascent regime
180
(Figure 2b). In the ascent regime, vertical diffusion (turbulent mixing) is also a sink for
181
cloud liquid produced by moist physical processes (not shown). Deep convection has a
182
9
negligible influence on Southern Ocean cloud liquid, and thus deep convective tendencies
183
are small.
184
185
Since cloud liquid and ice water content are governed by complex prognostic
186
equations involving many processes, coming up with a simple explanation for insufficient
187
supercooled liquid in CAM5 at first seems hopeless. Knowing that shallow convective
188
detrainment is the primary source of Southern Ocean cloud liquid (Figure 2) is extremely
189
helpful in this regard. Indeed, knowing that shallow convective detrainment is the primary
190
source of Southern Ocean cloud liquid (Figure 2) helps explain why most of the
191
supercooled cloud liquid occurs between temperatures of 268 K and 273 K (Figure 1).
192
While the cloud microphysics predicts cloud phase, the shallow convection prescribes
193
cloud phase for the detrained condensate that forms cloud using a piecewise linear
194
function of temperature:
195
196
f=0 ; for T > Tice
197
f= (Tice T)/ 30 ; for 238.15 K < T < Tice (Eq. 1)
198
f=1 ; for T < 238.15 K
199
200
where f is the glaciated fraction (unitless), T is temperature (K), and Tice is a constant that
201
specifies the temperature below which the shallow convection detrains ice (K). In the
202
default configuration, Tice = 268 K. In other words, detrained condensate below 268 K
203
contains ice, but above 268 K all detrained condensate is liquid.
204
205
10
Because the shallow convection is the primary source of low-level cloud condensate
206
over the mid-latitude oceans, changing Tice in Eq. 1 provides a simple lever to change the
207
amount of supercooled liquid in shallow convective clouds. Motivated by observations
208
showing supercooled liquid dominates at temperatures well below 268 K, we conduct
209
experiments in which we increase the supercooled liquid detrained from shallow
210
convection by changing Tice. Consistent with Figures 1 and 2, changing Tice from 268 K to
211
253 K implies that a much larger fraction of the cloud condensate detrained by the shallow
212
convection will be liquid over the Southern Ocean. We note that changing Tice has no effect
213
on low-level tropical clouds whose cloud top temperatures exceed 273 K. We also note that
214
Tice =253 K was the original value proposed by Park and Bretherton (2009), but it was not
215
adopted as the default value during CESM-CAM5 development.
216
217
b. CESM-CAM5 runs used for this study
218
We next introduce the model runs used in this study (Table 1). To accomplish goal
219
#1 testing the hypothesis that increasing supercooled liquid in detrained condensate
220
from the shallow convection reduces the excessive mid-latitude ocean ASR bias in CESM-
221
CAM5 we use atmosphere-only runs with fixed year 2000 sea ice and ocean conditions.
222
Atmosphere-only runs have prognostic atmosphere and land, but prescribed surface ocean
223
conditions and no ocean heat transport. Atmosphere-only runs are useful for quantifying
224
model biases resulting from the model representation of atmospheric processes. For
225
example, CESM-CAM5 has excessive Antarctic sea ice, a bias that complicates interpretation
226
of ASR bias at high Southern Ocean latitudes. We completed both a control” atmosphere-
227
only CAM5 run in which the physics is identical to the released version of CESM-CAM5 and
228
11
an experiment” atmosphere-only CAM5 run in which we decreased Tice from its control
229
value of 268 K to its experiment value of 253 K (see Eq. 1). To test our hypothesis, we
230
evaluate top-of-atmosphere ASR biases by comparing model runs to satellite-observed
231
radiative fluxes from version 2.8 of the Clouds and the Earth's Radiant Energy System-
232
Energy Balanced and Filled (CERES-EBAF) dataset (Loeb et al. 2009).
233
234
To accomplish goal #2 evaluating the impact of reduced mid-latitude ocean ASR
235
bias on global climate we use fully coupled CESM-CAM5 simulations with constant 1850
236
climate conditions starting with initial conditions from January 1, year 402 from the 1850
237
fully coupled control run of the CESM-CAM5 Large Ensemble Project (Kay et al. 2015). Fully
238
coupled models have prognostic atmosphere, ocean, land, and sea ice component models,
239
and as such can predict changes in atmosphere and ocean heat transport in response to
240
model physics changes like those proposed here. Like the atmosphere-only runs, we
241
completed fully coupled runs with differing values of Tice. To complement the fully coupled
242
runs, we completed slab ocean model (SOM) runs. SOM runs have prognostic mixed layer
243
ocean and sea ice, but constant prescribed ocean heat transport. As a result, SOM runs
244
come into equilibrium on the timescales of the mixed layer ocean (decades, proportional to
245
the mixed layer heat capacity) as opposed to the timescales of the deep ocean (centuries+).
246
We use the SOM runs to identify the importance of ocean heat transport for the model
247
climate response to ASR bias reduction.
248
249
250
251
12
3. Results
252
253
a. Shortwave radiation bias reduction in atmosphere-only runs
254
We begin by presenting ASR bias in present-day atmosphere-only runs. We compare
255
the control with Tice=268 K (ATM2000_cnt) to the experiment with Tice=253 K
256
(ATM2000_exp, Table 1). Decreasing Tice increases the cloud liquid water path (LWP) and
257
decreases ASR over the mid-latitude oceans in both hemispheres where shallow convective
258
clouds dominate (Figure 3a). At Southern Hemisphere mid-latitudes, oceans are the
259
dominant surface type and there is an active stormtrack year round. As a result, we find
260
decreasing Tice leads to the largest ASR bias reductions over the Southern Ocean. Not
261
surprisingly, global maps show the largest Southern Ocean ASR reductions occur in
262
Southern Hemisphere summer (DJF) (Figure 3b). In the North Atlantic and North Pacific,
263
ASR reductions occur in transition seasons (MAM, SON), but not during Northern
264
Hemisphere summer (JJA) when many shallow convective mid-latitude clouds have
265
temperatures above 273 K and mid-latitude cyclone activity is reduced (not shown).
266
267
Focusing on the region with the largest ASR bias and bias reduction, we next
268
compare zonal mean ASR model bias in the control and the experiment over the Southern
269
Ocean (Figure 4). Increasing the supercooled liquid condensate detrained from the shallow
270
convection reduces the DJF (annual) Southern Ocean 30-70 °S ASR bias by 11.1 (6.5) Wm-2
271
from 16.4 Wm-2 (7.3 Wm-2) to 5.3 Wm-2 (0.8 Wm-2). The largest ASR bias reductions occur
272
from 50 to 70 °S where detrained ice in the control (Figure 1g-i) became detrained
273
supercooled liquid in the experiment. In summary, the atmosphere-only model runs
274
13
demonstrate that increasing the supercooled liquid in shallow convective clouds reduces
275
mid-latitude ocean ASR biases, especially over the Southern Ocean.
276
277
b. Shortwave radiation bias reduction in fully coupled model runs
278
Having demonstrated that we can substantially reduce mid-latitude ocean ASR
279
biases by decreasing Tice in atmosphere-only runs, we next evaluate the impacts of the
280
same Tice changes in fully coupled model runs. We begin by comparing global mean energy
281
imbalance in a fully coupled experiment (FC1850_exp, Tice=253 K) with a fully coupled
282
control (FC1850_cnt, Tice=268 K). While the control has a small positive TOA energy
283
imbalance of 0.3 Wm-2, the experiment has a larger negative TOA energy imbalance of -0.9
284
Wm-2. This imbalance must be reduced to produce a stable coupled simulation. The
285
negative global energy imbalance in the experiment results from strong shortwave cloud
286
cooling (Table 2). Not surprisingly, the negative TOA energy imbalance in the experiment
287
leads to sustained global cooling. After 29 years, the global mean temperature in the
288
experiment decreases by 0.5 K and shows no sign of stopping. Worryingly, a similar
289
experiment completed within a slab ocean framework experienced runaway global cooling
290
of 40 K after 40 years with sea ice reaching the equator (not shown). Global cooling in this
291
snowball earth run was enhanced by low-latitude shortwave cloud and sea ice feedbacks
292
that were especially prominent after year 25 (not shown).
293
294
Decreasing Tice has a large effect on ASR, which brings the model out of global
295
energy balance. To obtain a stable fully coupled model run with the ASR bias reductions
296
resulting from decreasing Tice, additional model parameter changes are required. We found
297
14
appropriate additional model parameter changes by performing an assessment of the
298
regional contributions to global mean energy balance in CESM-CAM5. Despite having large
299
Southern Ocean ASR biases, the fully coupled control has a stable climate with a small
300
global energy imbalance. The juxtaposition of a large Southern Ocean ASR bias and small
301
global energy imbalance implies that the default version of CESM-CAM5 achieves a globally
302
balanced radiation budget with large compensating biases. Indeed, zonal mean plots of ASR
303
bias for the control (Figure 5a, FC1850_cnt, solid lines) show insufficient tropical ASR
304
compensates for excessive mid-latitude ASR in CESM-CAM5, a common occurrence in
305
climate models (Trenberth and Fasullo, 2010). An exciting opportunity for both
306
extratropical and tropical bias reduction emerges from globally compensating Southern
307
Ocean and tropical ASR errors. If we can combine model parameter changes to increase
308
tropical ASR with Tice decreases to decrease Southern Ocean ASR, we can obtain a stable
309
climate simulation with improved ASR in both the Tropics and the Mid-latitudes.
310
311
In order to increase tropical ASR, we made two modifications to the threshold
312
relative humidity for low cloud formation (rhminl). First, we increased rhminl from the
313
default value 0.8925 to 0.9175. Second, unlike in the control where rhminl over the land
314
was 0.1 lower than rhminl over the ocean, we specified that the land have the same rhminl
315
as the ocean. Both rhminl changes should be considered tuning” (Mauritsen et al. 2012) to
316
obtain global energy balance as they were motivated by energy balance considerations and
317
the desire for a stable coupled climate.
318
319
15
Having described the motivation to increase detrained supercooled liquid (decrease
320
Tice) and to tune the model (rhminl changes), we combine the Tice and rhminl changes to
321
obtain a fully coupled “tuned experiment” (FC1850_texp, Table 1). The fully coupled tuned
322
experiment has a near-zero global energy imbalance, and a stable global mean surface
323
temperature that is similar to the fully coupled control (Table 2). Notably, the Tice and
324
rhminl changes in the tuned experiment produce a stable coupled climate with reduced
325
ASR bias over both the Southern Ocean and the Tropics (Figure 5a, Table 2). Because the
326
rhminl changes primarily affect low clouds, their impact on shortwave cloud cooling was
327
larger than on longwave cloud warming (Table 2). The rhminl changes increased tropical
328
ASR by reducing shortwave cloud cooling in the tuned experiment when compared to the
329
control, but had only a small impact on the extra-tropics (Figure 5b). Confirming the zonal
330
mean picture, global maps reveal that the largest LWP increases and ASR decreases in the
331
fully coupled tuned experiment are over the Southern Ocean (Figure 6). The fully coupled
332
tuned experiment ASR reductions and LWP increases (Figure 6) are similar to the
333
atmosphere-only experiment (Figure 3). In other words, we retained the desirable
334
Southern Ocean ASR bias reduction in a fully coupled framework.
335
336
The climate of the fully coupled tuned experiment (FC1850_texp) quickly adjusts to
337
the large radiation changes and reaches a new equilibrium climate state. Over the first 30
338
years, the global ocean heat imbalance (a measure of ocean thermal disequilibrium) in
339
FC1850_texp was small (0.06 Wm-2). During years 150-200 of FC1850_texp, the global
340
ocean heat imbalance is an order of magnitude smaller (0.007 Wm-2) than during the first
341
30 years of FC1850_texp. In addition, the global top-of-atmosphere imbalance remains
342
16
small (<0.1 Wm-2) over the entire 200 years of simulation of FC1850_texp. We did not find
343
a large climate-relevant transient response over 200 years of simulation. These results
344
suggest that the 200-year-long FC1850_texp simulation can be used to understand the fully
345
coupled climate system response to brightening the Southern Ocean and dimming the
346
tropics.
347
348
c. Climate influence of shortwave radiation bias reduction in fully coupled model runs
349
To evaluate the equilibrium climate system response to the ASR bias reductions
350
shown in Figure 5a, we next compare years 150-200 of the fully coupled tuned experiment
351
(FC1850_texp) with the fully coupled control (FC1850_cnt). Figure 7 compares zonal mean
352
air temperatures and zonal winds. Due to an enhanced meridional temperature gradient,
353
the Southern Hemisphere jet strengthens in the tuned experiment as compared to the
354
control. A stronger Southern Hemisphere jet is seen in all seasons, but especially in DJF
355
(Figure 7d). Small poleward jet shifts are also evident in the Southern Hemisphere. The
356
maximum wind speed at 850 mb, often used to identify the eddy-induced jet, shifts
357
poleward by 0.1° (from -51.9 to -52.1 °S) in the annual mean and shifts poleward by 0.6°
358
(from 50.9 to 51.5 °S) in DJF.
359
360
The results shown thus far demonstrate that ASR bias reduction can alter global
361
energy budgets, temperature gradients, and atmospheric circulation. These findings
362
motivate an assessment of the influence of ASR bias reduction on global heat transport.
363
Figure 8 shows northward heat transport in the control and tuned experiment, including a
364
partitioning into the atmosphere and the ocean appropriate for equilibrium conditions (see
365
17
Appendix of Kay et al. 2012b for methods). Consistent with enhanced meridional
366
temperature gradients, poleward heat transport increases in the tuned experiment as
367
compared to the control. The Southern Hemisphere poleward heat transport increase is
368
more than double the Northern Hemisphere poleward heat transport increase. In addition
369
to inter-hemispheric magnitude differences, there are inter-hemispheric differences in the
370
atmospheric and oceanic contributions to total heat transport change. In the Northern
371
Hemisphere, poleward heat transport increases are due entirely to the atmosphere while in
372
the Southern Hemisphere, both the ocean (dominant in the Tropics) and atmosphere
373
(dominant in the mid-latitudes) contribute to increased poleward heat transport.
374
375
We next examine the mechanisms underlying the increased poleward oceanic heat
376
transport in the Southern Hemisphere in FC1850_texp as compared to FC1850_cnt. The
377
increased Southern Hemisphere poleward ocean heat transport response emerges in the
378
first 30 years of FC1850_texp. This quick response implicates the shallow wind-driven
379
ocean circulation, not the deep thermohaline ocean circulation. Figure 9 shows the control
380
surface zonal wind stress climatology and wind stress response (FC1850_texp -
381
FC1850_cnt). Over the Southern Hemisphere sub-tropics, easterly wind stress increases,
382
especially in the Pacific basin. Figure 10 shows the meridional overturning circulation
383
climatology and response to the wind stress changes shown in Figure 9. In the top 1000
384
meters of the Southern Hemisphere subtropical ocean, there is a deepening and
385
intensification of the shallow overturning ocean circulation. The enhanced easterly winds
386
shown in Figure 9b/c increase poleward Ekman transport leading to increased poleward
387
oceanic heat transport. There is also a poleward shift in the location of upwelling in the
388
18
Southern Ocean. By comparing Figure 9b/10b based on years 1-30 of FC1850_texp and
389
Figure 9c/10c based on years 150-200 of FC1850_texp, it is clear that this shallow wind-
390
driven ocean circulation response emerged almost entirely in first 30 years and remains
391
stable over 200 years of coupled simulation. After 200 years of simulation, changes in the
392
deep ocean thermohaline circulation are small, second-order, and have a minor impact on
393
the ocean heat transport response in Figure 8 (not shown).
394
395
Table 3 compares cross-equatorial heat transport and tropical precipitation in the
396
fully coupled simulations. While both the atmosphere and the ocean contribute, the ocean
397
is responsible for 80% of the reductions in northward cross-equatorial heat transport in
398
FC1850_texp as compared to FC1850_cnt. Consistent with a small change in atmospheric
399
cross-equatorial heat transport, the tropical atmospheric circulation and precipitation
400
responses are modest. Never-the-less, the sense of the responses are consistent with
401
Hwang and Frierson (2013): ameliorating the Southern Ocean ASR bias reduces northward
402
cross-equatorial atmospheric heat transport and shifts tropical precipitation northward
403
reducing the tropical precipitation asymmetry index bias. Unlike FC1850_cnt, atmospheric
404
cross-equatorial heat transport in FC1850_texp is within the uncertainty bounds of modern
405
observationally based estimates (Loeb et al. 2015). Yet, the total and oceanic cross-
406
equatorial heat transports are underestimated in FC1850_texp when compared with
407
modern observationally based estimates. If comparison with modern observationally based
408
estimates is appropriate, FC1850_texp has less realistic total and oceanic cross-equatorial
409
heat transport than FC1850_cnt.
410
411
19
d. Influence of ocean heat transport on climate response to shortwave radiation bias
412
reduction in slab ocean model runs
413
The results shown in Figures 8-10 suggest that including dynamic ocean heat
414
transport limits a proposed atmospheric teleconnection linking Southern Ocean ASR bias
415
reduction with tropical precipitation shifts. Because slab ocean models are run with
416
prescribed ocean heat transport, they provide a useful framework to quantify the influence
417
of ocean heat transport on a modeled climate response. Leveraging the capability to
418
prescribe ocean heat transport within SOMs, we next identify the influence of ocean heat
419
transport on the climate response to ASR bias reduction. Specifically, we compare a SOM
420
experiment with prescribed ocean heat transport changes resulting from ASR bias
421
reduction (SOM1850_texp_ocnht=texp) to a SOM experiment with fixed ocean heat
422
transport (SOM1850_texp_ocnht=cnt). Both SOM experiments have ASR bias reduction
423
similar to the fully coupled experiment (Figure 5).
424
425
We begin by comparing atmospheric heat transport in the SOMs and fully coupled
426
runs (Figure 11). Like the fully coupled experiment with dynamic ocean heat transport, the
427
SOM experiment with changed ocean heat transport has small changes in cross-equatorial
428
atmospheric heat transport. In contrast, the SOM with fixed ocean heat transport has large
429
increases (0.3 Petawatts) in atmospheric cross-equatorial heat transport. When ocean heat
430
transport is fixed, ASR bias reduction produces increases in both northward cross-
431
equatorial moisture transport and southward cross-equatorial energy transport (Figure
432
11b).
433
434
20
If cross-equatorial atmospheric heat and moisture transport change (Figure 11b),
435
we expect a large tropical atmospheric circulation, moisture convergence, and precipitation
436
response. Indeed, Figures 12 and 13 confirm the importance of cross-equatorial
437
atmospheric heat transport for shifting the ITCZ. When ocean heat transport is fixed, large
438
cross-equatorial atmospheric heat transport changes occur. As a result, the tropical
439
atmospheric response to reduced ASR includes reduced subsidence and increased
440
precipitation north of the equator, and enhanced subsidence and decreased precipitation
441
south of the equator (Figure 12a, Figure 13). The tropical precipitation asymmetry index
442
in SOM1850_tex_ocnht=cnt is 0.33, which is much larger than in FC1850_texp (0.15) or in
443
modern-day observations (0.20) (Table 3). On the other hand, when ocean heat transport
444
changes are predicted or prescribed, a muted tropical atmospheric circulation and
445
precipitation response occurs (Figure 12b/c, Figure 13). These results confirm that with
446
dynamic ocean heat transport, tropical precipitation is less affected by Southern Ocean
447
cooling. In the climate model experiments with dynamic or prescribed ocean circulation
448
changes, Southern Ocean ASR bias reduction leads to small ITCZ northward shifts because
449
the ocean, not the atmosphere, dominates the cross-equatorial heat transport response. To
450
summarize these results, Figure 14 contains a schematic contrasting the climate impacts of
451
ASR bias reduction with and without dynamic ocean heat transport.
452
453
454
455
456
21
4. Discussion
457
Our results demonstrate that shallow convective cloud phase exerts a strong control
458
on global energy balance. Specifically, increasing supercooled liquid in shallow convective
459
clouds to better match observations enables large reductions in long-standing shortwave
460
radiation biases in a state-of-the-art climate model (CESM-CAM5). The climate impacts of
461
these shortwave radiation bias reductions are profound. A cooler brighter Southern Ocean
462
and warmer dimmer Tropics leads to increased poleward heat transport, especially in the
463
Southern Hemisphere. In response to stronger meridional temperature gradients, the
464
Southern Hemisphere atmospheric jet increases in strength, and shifts slightly poleward as
465
proposed by Ceppi et al. (2012). This jet strengthening and small poleward jet shift is
466
consistent with a more poleward and realistic jet location in CESM-CAM5 (Kay et al. 2014)
467
when compared to other models of its class (Barnes and Polvani 2013). Unfortunately, the
468
westerly winds over the Southern Ocean are already too strong in CESM-CAM5, due in part
469
to a cold upper troposphere bias present in many models without a resolved stratospheric
470
circulation (e.g., Charlton-Perez et al. 2013). Though “fixing” the Southern Ocean shortwave
471
radiation bias is an important step forward and enables new science, it is not a panacea.
472
473
Yet, perhaps the most fundamental outcome of this study is a null result with
474
relevance to climate dynamics. In our fully coupled tuned experiment, ITCZ shifts are small
475
in response to a preferentially cooled Southern Hemisphere. Cooling the Southern Ocean
476
has little impact on tropical precipitation because cross-equatorial heat transport changes
477
occur primarily in the ocean (80% of the response), not the atmosphere (20% of the
478
response). For this reason, quantifying the influence of Southern Ocean cooling on global
479
22
climate requires the use of a model with dynamic ocean heat transport. Similarly, Deser et
480
al. (2015) emphasize the importance of dynamic ocean heat transport and atmosphere-
481
ocean coupling in producing the equatorially symmetric pattern of response to Arctic sea
482
ice loss. More broadly, our results raise a critical question for the climate dynamics of
483
extratropical-tropical teleconnections: Given inter-hemispheric temperature differences,
484
when does the ocean, not the atmosphere, accomplish the required cross-equatorial heat
485
transport? A dynamic ocean has been neglected in many (but not all) previous studies that
486
analyzed atmospheric teleconnections and ITCZ shifts (e.g., Seo et al. 2014; Hwang and
487
Frierson 2013 supplementary materials; Frierson and Hwang 2012, some but not all
488
references in Chiang and Friedman 2012). An important role for the ocean in cross-
489
equatorial heat transport is certainly possible. For example, recent work has emphasized
490
the importance of oceanic cross-equatorial heat transport in setting the mean position of
491
the ITCZ (Frierson et al. 2013; Fuckar et al. 2013; Marshall et al. 2014). Slab ocean models
492
are useful for isolating the role of dynamic ocean heat transport, but their use in
493
quantifying global teleconnections should be justified via comparison with fully coupled
494
model simulations.
495
496
While we have reduced a long-standing climate model bias, documented the climate
497
response, and highlighted the role of dynamic ocean circulation in global teleconnections,
498
important additional work remains. First of all, are our results robust in other models?
499
Recent work using an independent fully coupled climate model (HadGEM2-ES)
500
corroborates our finding that tropical rainfall is insensitive to southern hemispheric
501
cooling (Matt Hawcroft, personal communication). Second, this study focuses on mean
502
23
state 1850 climate, but more realistic shallow convective cloud phase in climate models
503
also has implications for transient simulations and cloud-climate feedbacks. While the
504
global mean cloud-climate feedback in response to increased greenhouse gases is likely
505
positive (Boucher et al. 2013), negative Southern Ocean cloud feedbacks are present in
506
most climate models (Zelinka et al. 2013) including CESM-CAM5 (Kay et al. 2014). These
507
negative cloud feedbacks are a consequence of optical depth increases (Zelinka et al. 2013),
508
with a large contribution from clouds changing in phase from ice to liquid (Tsushima et al.
509
2006, Kay et al. 2014) and a small contribution from poleward jet shifts (Kay et al. 2014,
510
Grise and Polvani 2014, Ceppi and Hartmann 2015). Yet, if cloud phase is diagnosed purely
511
as a function of temperature in climate models (e.g., as it is within the shallow convection
512
parameterization within CESM-CAM5 see Eq. 1), warming automatically converts ice to
513
liquid and increases cloud optical depth. Two obvious follow-on questions emerge. First,
514
could having more realistic liquid-dominated Southern Ocean clouds change the sign of the
515
Southern Ocean cloud-climate feedback from negative to positive in CESM-CAM5? We
516
expect the negative cloud feedback to be sensitive to both the amount of ice present in the
517
mean state and its susceptibility to warming (McCoy et al. 2015). Second, how realistic are
518
parameterizations that predict phase change purely a function of temperature? The
519
amount of liquid and ice in clouds in the real atmosphere depends on many factors beyond
520
temperature (Morrison et al. 2012). Prognostic microphysical processes that predict cloud
521
phase are increasingly common in models, but are still often neglected in parameterized
522
convection (e.g., Park et al. 2014, Forbes and Ahgrimm 2014).
523
524
525
24
526
527
5. Summary
528
We reduced large, long-standing, and pervasive shortwave radiation model biases
529
over the Southern Ocean and the Tropics in a state-of-the-art global coupled climate model
530
(CESM-CAM5). Bias reductions resulted from increasing the supercooled liquid water in
531
shallow convective clouds to better match observations and tuning the relative humidity
532
threshold for low clouds to obtain global energy balance. The climate impacts of the
533
reduced shortwave radiation biases included: a cooled Southern Ocean, a warmed Tropics,
534
increased poleward heat transport (especially in the Southern Hemisphere), and a stronger
535
Southern Hemisphere atmospheric jet. Cooling the Southern Ocean had a negligible
536
influence on tropical circulation and rainfall in our fully coupled model experiments. We
537
found a weak extratropical-tropical atmospheric teleconnection in our fully coupled runs
538
because Southern Ocean cooling increased cross-equatorial heat transport primarily in the
539
ocean, not the atmosphere. More broadly, this work demonstrates the importance of using
540
a global fully coupled model with dynamic ocean heat transport when quantifying the
541
global climate impacts of inter-hemispheric temperature gradients. Future work will
542
investigate the influence of more realistic cloud phase on transient climate simulations
543
including especially Southern Ocean shortwave cloud-climate feedbacks. Many important
544
climate questions remain for the Southern Ocean.
545
546
547
548
25
549
550
Acknowledgements: The authors wish to thank Clara Deser, Andrew Gettelman, Matthew
551
Woelfie, and Chris Bretherton for fruitful conversations related to this work, the
552
Yellowstone CESM CSL for computing resources, and the scientists and software engineers
553
that build CESM. This work was funded by start-up funds awarded to J. E. Kay by the
554
University of Colorado Cooperative Institute for Research in Environmental Sciences
555
(CIRES). P. Caldwell was supported by the Department of Energy's Office of Science
556
Biological and Environmental Research Division Global Climate Modeling Group at
557
Lawrence Livermore National Laboratory under Contract DE-AC52-07NA27344. The
558
National Center for Atmospheric Research is sponsored by the National Science
559
Foundation.
560
26
References
561
Adler, R.F. and Coauthors, 2003: The Version 2 Global Precipitation Climatology Project
562
(GPCP) Monthly Precipitation Analysis (1979 - Present). J. Hydrometeor., 4(6), 1147-
563
1167.
564
Barnes, E.A., and L.M. Polvani, 2013: Response of the midlatitude jets and of their
565
variability to increased greenhouse gases in the CMIP5 models, J. Clim., 26, 7117
566
7135, doi: http://dx.doi.org/10.1175/JCLI-D-12-00536.1.
567
Bergeron, T., 1935: On the physics of clouds and precipitation. Proces Verbaux de
568
l’Association de Météorologie, International Union of Geodesy and Geophysics, 156–
569
178.
570
Bodas-Salcedo, A. and Coauthors, 2012: The Surface Downwelling Solar Radiation
571
Surplus over the Southern Ocean in the Met Office Model: The Role of Midlatitude
572
Cyclone Clouds. J. Climate, 25:21, 7467-7486
573
Bodas-Salcedo, A. and Coauthors, 2014: Origins of the Solar Radiation Biases over
574
the Southern Ocean in CFMIP2 Models. J. Climate, 27, 4156. doi:
575
http://dx.doi.org/10.1175/JCLI-D-13-00169.1
576
Bodas-Salcedo, A. and Coauthors, 2015: Large contribution of supercooled liquid clouds to
577
the solar radiation budget of the Southern Ocean, submitted to J. Climate
578
Boucher, O., D. and Coauthors, 2013: Clouds and Aerosols. In: Climate Change 2013: The
579
Physical Science Basis. Contribution of Working Group I to the Fifth Assessment
580
Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K.
581
Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley
582
27
(eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY,
583
USA.
584
Ceppi, P., Y.-T. Hwang, D. M. W. Frierson, and D. L. Hartmann, 2012: Southern Hemisphere
585
jet latitude biases in CMIP5 models linked to shortwave cloud forcing, Geophys. Res.
586
Lett., 39, L19708, doi:10.1029/2012GL053115.
587
Ceppi, P., and D. L. Hartmann, 2015: Connections between clouds, radiation, and
588
midlatitude dynamics: a review, Current Climate Change Reports, doi:
589
10.1007/s40641-015-0010-x
590
Cesana, G., and H. Chepfer, 2013: Evaluation of the cloud thermodynamic phase in a
591
climate model using CALIPSO-GOCCP, J. Geophys. Res. Atmos., 118, 79227937,
592
doi:10.1002/jgrd.50376.
593
Charlton-Perez, A. J., and Coauthors 2013: On the lack of stratospheric dynamical variability
594
in low-top versions of the CMIP5 models, J. Geophys. Res. Atmos., 118, 24942505,
595
doi:10.1002/jgrd.50125.
596
Chiang, J.C.H., and A. R. Friedman, 2012: Extratropical Cooling, Interhemispheric
597
Thermal Gradients, and Tropical Climate Change. Annual Reviews of Earth and
598
Planetary Sciences, 40, 383-412.
599
Chubb, T. H., J. B. Jensen, S. T. Siems, and M. J. Manton, 2013: In situ observations of
600
supercooled liquid clouds over the Southern Ocean during the HIAPER Pole-to-Pole
601
Observation (HIPPO) campaigns, Geophys. Res. Lett., 40, 5280-
602
5285doi:10.1002/grl.50986.
603
Deser, C., Tomas, R. A., and L. Sun, 2015: The Role of OceanAtmosphere Coupling in
604
28
the Zonal-Mean Atmospheric Response to Arctic Sea Ice Loss. J. Climate, 28, 2168
605
2186. doi: http://dx.doi.org/10.1175/JCLI-D-14-00325.1
606
Findeisen, W., 1938: Kolloid-meteorologische Vorgänge bei Neiderschlags-bildung. Meteor.
607
Z., 55, 121133.
608
Frierson, D. M. W. and Y.-T. Hwang, 2012: Extratropical Influence on ITCZ Shifts in Slab
609
Ocean Simulations of Global Warming. J. Climate, 25, 720733.
610
doi: http://dx.doi.org/10.1175/JCLI-D-11-00116.1
611
Forbes, R. M. and M. Ahlgrimm, 2014: On the Representation of High-Latitude Boundary
612
Layer Mixed-Phase Cloud in the ECMWF Global Model. Mon. Wea. Rev., 142, 3425
613
3445. doi: http://dx.doi.org/10.1175/MWR-D-13-00325.1
614
Frölicher, T. L. and Coauthors, 2015: Dominance of the Southern Ocean in
615
Anthropogenic Carbon and Heat Uptake in CMIP5 Models. J. Climate, 28, 862
616
886. doi: http://dx.doi.org/10.1175/JCLI-D-14-00117.1
617
Fučkar, N. S., and Coauthors, 2013: Influence of the Extratropical Ocean Circulation on the
618
Intertropical Convergence Zone in an Idealized Coupled General Circulation Model. J.
619
Climate, 26, 46124629. doi: http://dx.doi.org/10.1175/JCLI-D-12-00294.1
620
Gettelman, A., Morrison, H., Santos, S., Bogenschutz, P. and P. M. Caldwell, (2015),
621
Advanced Two-Moment Bulk Microphysics for Global Models. Part II: Global Model
622
Solutions and AerosolCloud Interactions. J. Climate, 28, 12881307.
623
doi: http://dx.doi.org/10.1175/JCLI-D-14-00103.1
624
Grise, K. M. and L. M. Polvani, 2014: Southern Hemisphere CloudDynamics Biases in
625
CMIP5 Models and Their Implications for Climate Projections. J. Climate, 27, 6074
626
6092. doi: http://dx.doi.org/10.1175/JCLI-D-14-00113.1
627
29
Hu, Y., S. Rodier, K. Xu, W. Sun, J. Huang, B. Lin, P. Zhai, and D. Josset, 2010:
628
Occurrence, liquid water content, and fraction of supercooled water clouds from
629
combined CALIOP/IIR/MODIS measurements. J. Geophys. Res., 115, D00H34,
630
doi:10.1029/2009JD012384.
631
Huang, Y., S. T. Siems, M. J. Manton, A. Protat, and J. Delanoë, 2012: A study on the
632
low-altitude clouds over the Southern Ocean using the DARDAR-MASK. J. Geophys.
633
Res., 117, D18204, doi:10.1029/2012JD017800.
634
Hurrell, J., and Coauthors, 2013: The Community Earth System Model: A Framework
635
for Collaborative Research, Bull. Amer. Meteor. Soc., doi:
636
http://dx.doi.org/10.1175/BAMS-D-12-00121.1.
637
Hwang, Y.-T. and D. M. W. Frierson, 2013: A link between the double intertropical
638
convergence zone problem and cloud biases over the Southern Ocean, Proc. Nat.
639
Acad. Sci., doi: 10.1073/pnas.1213302110.
640
Kay, J. E., and Coauthors, 2015: The Community Earth System Model (CESM) Large
641
Ensemble Project: A Community Resource for Studying Climate Change in the
642
Presence of Internal Climate Variability, Bulletin of the American Meteorological
643
Society, 96, 13331349, doi:10.1175/BAMS-D-13-00255.1.
644
Kay, J. E., Bourdages, L., Chepfer, H., Miller, N., Morrison, A., Yettella, V., and B. Eaton , 2016:
645
Evaluating and improving cloud phase in the Community Atmosphere Model version
646
5 using spaceborne lidar observations, resubmitted March 14, 2016 with “minor
647
revisions” to Journal of Geophysical Research - Atmospheres, Available at:
648
http://www.cgd.ucar.edu/staff/jenkay/papers/2015JD024699-
649
T_Kay_JGR_CAM_COSP_cloudphase_revised_March14_2016.pdf
650
30
Kay, J. E., B. Medeiros, Y.-T. Hwang, A. Gettelman, J. Perket, and M. G. Flanner, 2014:
651
Processes controlling Southern Ocean shortwave climate feedbacks in CESM.
652
Geophys. Res. Lett., 41, doi:10.1002/2013GL058315
653
Kay, J. E., Hillman, B., Klein, S., Zhang, Y., Medeiros, B., Gettelman, G., Pincus, R.,
654
Eaton, B., Boyle, J., Marchand, R. and T. Ackerman, 2012a: Exposing global cloud
655
biases in the Community Atmosphere Model (CAM) using satellite observations and
656
their corresponding instrument simulators, J. Climate, 25, 5190-5207, doi:
657
http://dx.doi.org/10.1175/JCLI-D-11-00469.1
658
Kay, J. E., Holland, M. M., Bitz, C., Blanchard-Wrigglesworth, E.,Gettelman, A., Conley,
659
A., and D. Bailey, 2012b: The influence of local feedbacks and northward heat
660
transport on the equilibrium Arctic climate response to increased greenhouse gas
661
forcing in coupled climate models, J. Climate, 25, 5433-5450, doi:10.1175/JCLI-D-
662
11-00622.1
663
Lin, J.-L., 2007: The double-itcz problem in ipcc ar4 coupled gcms: Oceanatmosphere
664
feedback analysis. J. Climate, 20 (18), 44974525, doi:10.1175/JCLI4272.1,
665
Loeb, N. G. and Coauthors, 2009: Toward Optimal Closure of the Earth’s Top-of-
666
Atmosphere Radiation Budget, J. Climate, 22, 748766, doi:
667
10.1175/2008JCLI2637.1.
668
Loeb, N. G., H. Wang, A. Cheng, S. Kato, J. T. Fasullo, K. Xu, and R. P. Allan (2015),
669
Observational constraints on atmospheric and oceanic cross-equatorial heat
670
transports: Revisiting the precipitation asymmetry problem in climate models, Clim.
671
Dyn., doi:10.1007/s00382-015-2766-z.
672
Marshall J, Donohoe A, Ferreira D, McGee D, 2013: The ocean’s role in setting the mean
673
31
position of the Inter-Tropical Convergence Zone. Clim Dyn 113.
674
doi:10.1007/s00382-013-1767-z
675
Mauritsen, T. and Coauthors 2012: Tuning the climate of a global model, J. Adv. Model. Earth
676
Syst., 4, M00A01, doi:10.1029/2012MS000154.
677
McCoy, D. T., Hartmann, D. L., and D. P. Grosvenor, 2014: Observed Southern Ocean
678
Cloud Properties and Shortwave Reflection. Part II: Phase Changes and Low Cloud
679
Feedback. J. Climate, 27, 88588868.
680
doi: http://dx.doi.org/10.1175/JCLI-D-14-00288.1
681
Morrison, H., and A. Gettelman, 2008: A new two-moment bulk stratiform cloud
682
microphysics scheme in the NCAR Community Atmosphere Model (CAM3). Part I:
683
Description and numerical tests. J. Climate, 21, 36423659.
684
Morrison, H., G. de Boer, G. Feingold, J. Harrington, M. D. Shupe, and K. Sulia, 2012:
685
Resilience of persistent Arctic mixed-phase clouds. Nature Geoscience, 5, 11-17, doi:
686
10.1038/NGE01332.
687
Morrison, A. E., Siems, and S. T., and M. J. Manton, 2011: A Three-Year Climatology of
688
Cloud-Top Phase over the Southern Ocean and North Pacific. J. Climate, 24, 2405
689
2418. doi: http://dx.doi.org/10.1175/2010JCLI3842.1
690
O’Dell, C. W., F. J. Wentz, R. Bennartz, 2008: Cloud Liquid Water Path from Satellite-Based
691
Passive Microwave Observations: A New Climatology over the Global Oceans. J.
692
Climate, 21, 17211739.
693
Park, S., and C. S. Bretherton, 2009: The University of Washington shallow
694
convection and moist turbulence schemes and their impact on climate simulations
695
with the Community Atmosphere Model. J. Climate, 22, 34493469.
696
32
Park, S., Bretherton, C. S. and P. J. Rasch, 2014: Integrating Cloud Processes in the
697
Community Atmosphere Model, Version 5. J. Climate, 27, 68216856.
698
doi: http://dx.doi.org/10.1175/JCLI-D-14-00087.1
699
Schneider, T., T. Bischoff, and G. H. Haug, 2014: Migrations and dynamics of the
700
intertropical convergence zone. Nature, 513, 4553, doi:10.1038/nature13636.
701
Seo, J., Kang, S. M., and D. M. W. Frierson, 2014: Sensitivity of Intertropical Convergence
702
Zone Movement to the Latitudinal Position of Thermal Forcing. J. Climate, 27, 3035
703
3042. doi: http://dx.doi.org/10.1175/JCLI-D-13-00691.1
704
Shupe, M. D. and J. M. Intrieri, 2004: Cloud Radiative Forcing of the Arctic Surface:
705
The Influence of Cloud Properties, Surface Albedo, and Solar Zenith Angle. J. Climate,
706
17, 616628,
707
doi: 10.1175/1520-0442(2004)017<0616:CRFOTA>2.0.CO;2
708
Trenberth, K. E. and J. T. Fasullo, 2010: Simulation of Present-Day and Twenty-First-
709
Century Energy Budgets of the Southern Oceans. J. Climate, 23, 440454.,
710
doi:http://dx.doi.org/10.1175/2009JCLI3152.1
711
Tsushima, Y., et al. (2006), Importance of the mixed-phase cloud distribution in the control
712
climate for assessing the response of clouds to carbon dioxide increase: a multi-
713
model study, Clim. Dyn., (27): 113-126, DOI:10.1007/s00382-006-0127-7
714
Wegener, A., 1911: Thermodynamik der Atmosphäre. Leipzig, 331 pp.
715
Williams, K.D., and Coauthors, 2013: The Transpose-AMIP II Experiment and its
716
application to the understanding of southern ocean cloud biases in climate models. J.
717
Climate, 26, 3258-3274, DOI: 10.1175/JCLI-D-12-00429.1.
718
Zelinka, M. D., 2013: Contributions of Different Cloud Types to Feedbacks and Rapid
719
33
Adjustments in CMIP5*. J. Climate, 26, 50075027. doi:
720
http://dx.doi.org/10.1175/JCLI-D-12-00555.1
721
722
723
34
Table Captions:
724
Table 1. Description of global climate model runs. All runs use the Community Earth
725
System Model with the Community Atmosphere Model version 5 (CESM-CAM5, Hurrell et
726
al. 2013) at 1-degree horizontal resolution with the finite volume dynamical core.
727
728
Table 2. Global annual averages for fully coupled model runs. Observations are from
729
CERES-EBAF v2.8 (Loeb et al. 2009) and UWisc. (O’Dell et al. 2008). Root mean square
730
differences between model runs and observations are provided in parentheses. See Table 1
731
for a description of model runs.
732
733
Table 3. Cross-equatorial Heat Transport (CHT) and Tropical Precipitation
734
Asymmetry Index. The tropical precipitation asymmetry index (TPAI) is defined following
735
Hwang and Frierson (2013) as precipitation in the Northern Hemisphere tropics (0-20 °N
736
area-averaged) minus precipitation in the Southern Hemisphere tropics (0-20 °S area
737
averaged) normalized by the tropical mean precipitation (20 °S 20 °N area averaged). See
738
Table 1 for a description of model runs.
739
740
Figure Captions:
741
Figure 1. Atmospheric circulation, clouds, and air temperatures over the Southern
742
Ocean in the fully coupled control run (FC1850_cnt, Table 1). The left column shows
743
cloud fraction (colors) and subsidence (black lines) for the annual mean ANN (a), Southern
744
Hemisphere winter JJA (b) and Southern Hemisphere summer DJF (c). The middle column
745
shows grid-box cloud liquid mixing ratio (colors) and air temperature (black lines) for ANN
746
(d), JJA (e) DJF (f). The right column shows grid-box cloud ice mixing ratio (colors) and air
747
temperature (black lines) for ANN (g), JJA (h) and DJF (i).
748
749
Figure 2. Annual mean cloud liquid tendencies over the subsidence (30-50 °S, solid)
750
and ascent (50-70 °S, dash) regions of the Southern Ocean in the fully coupled
751
control run (FC1850_cnt, Table 1): a) vertical diffusion and all moist physics
752
tendencies, b) microphysics tendencies. Note: The total tendency due to moist physics
753
processes is the sum of MPDLIQ, MACPDLIQ, SHDLFLIQ CMFDLIQ, ZMDLIQ, and DPDLFLIQ.
754
The total tendency due to microphysical processes is the sum of MPDW2P, MPDW2I,
755
MPDW2V, QCSEDTEN, and QCSEVAP.
756
757
Figure 3. Global difference maps for atmosphere-only runs (ATM2000_exp
758
ATM2000_cnt): a) Absorbed shortwave radiation (ASR) difference for annual mean
759
(ANN) b) as in a) but for Southern Hemisphere summer (DJF), c) ANN grid-box mean
760
cloud liquid water path (LWP), d) as in c) but for DJF. See Table 1 for model run
761
descriptions.
762
763
Figure 4. Zonal mean Absorbed Shortwave Radiation bias over the Southern Ocean in
764
atmosphere-only runs. Values in parentheses indicate Southern Ocean (30-70 ° S) bias.
765
Observations are from CERES-EBAF (Loeb et al. 2009) v2.8 for the years 2000-2013. See
766
35
Table 1 for model run descriptions. Both annual means (ANN) and Southern Hemisphere
767
summer (DJF) means are plotted.
768
769
Figure 5. Zonal global mean absorbed shortwave radiation in fully coupled runs: a)
770
bias, b) difference from control (FC1850_cnt). Observations are from CERES-EBAF v2.8
771
(Loeb et al. 2009) for the years 2000-2013. An area-weighted axis is used to show the
772
balancing of mid-latitude and tropical ASR biases and differences. See Table 1 for model
773
run description. Both annual means (ANN) and Southern Hemisphere summer (DJF) means
774
are plotted.
775
776
777
36
Figure 6. Global difference maps for fully coupled climate model runs (FC1850_texp
778
FC1850_cnt): a) Annual mean (ANN) absorbed shortwave radiation (ASR) difference,
779
b) as in a) but for Southern Hemisphere summer (DJF), c) ANN grid-box mean cloud
780
liquid water path (LWP) difference, d) as in c) but for DJF. See Table 1 for model run
781
description. An average of years 150-200 is shown for FC1850_texp, while an average of
782
years 1-29 is shown for FC1850_cnt.
783
784
Figure 7. Global zonal vertical mean temperature and circulation in fully coupled
785
runs: a) Annual mean air temperature (black contours, FC1850_cnt) and
786
temperature response (colors, FC1850_texp - FC1850_cnt), b) as in a) but for
787
Southern Hemisphere summer (DJF), c) Annual zonal mean wind (black contours,
788
FC1850_cnt) and zonal mean wind response (colors, FC1850_texp - FC1850_cnt), d)
789
as in c) but for DJF. See Table 1 for model run descriptions. An average of years 150-200
790
is shown for FC1850_texp, while an average of years 1-29 is shown for FC1850_cnt.
791
792
Figure 8. Annual mean northward heat transport in fully coupled runs: a)
793
Climatology in fully coupled control run (FC1850_cnt, solid) and tuned experiment
794
(FC1850_texp, dash) b) Response to increased detrainment of supercooled liquid in
795
shallow convective clouds and tuning (FC1850_texp- FC1850_cnt). See Table 1 for
796
model run descriptions. An average of years 150-200 is shown for FC1850_texp, while an
797
average of years 1-29 is shown for FC1850_cnt. Heat transport calculated using methods
798
described in Kay et al. 2012b Appendix A. Total heat transport was calculated based on the
799
fact that, in steady state, horizontal heat flux convergence across a latitude band is
800
balanced by the net TOA flux poleward of that latitude. Ocean heat transport was calculated
801
inline. Atmospheric heat transport was calculated by integrating the residual between the
802
top-of-atmosphere and surface heat fluxes.
803
804
Figure 9. Surface zonal wind stress and wind stress response: a) climatology in fully
805
coupled control (FC1850_cnt, years 1-29), b) response to increased detrainment of
806
supercooled liquid in shallow convective clouds and tuning (FC1850_texp-
807
FC1850_cnt) using years 1-30 of FC1850_texp, c) as in b) but using years 150-200 of
808
FC1850_texp.
809
810
Figure 10. Meridional Overturning Circulation (MOC): a) climatology in fully coupled
811
control (FC1850_cnt, years 1-29), b) response to increased detrainment of
812
supercooled liquid in shallow convective clouds and tuning (FC1850_texp-
813
FC1850_cnt) using years 1-30 of FC1850_texp, c) as in b) but using years 150-200 of
814
FC1850_texp.
815
816
817
818
819
820
821
822
37
Figure 11. Atmospheric northward heat transport response to increased
823
detrainment of supercooled liquid in shallow convective clouds and tuning: a) total,
824
b) latent and dry static energy. Responses are shown for fully coupled model
825
(FC1850_texp - FC1850_cnt), slab ocean model with prescribed ocean heat transport
826
changes (SOM1850_texp_ocnht=texp - SOM1850_cnt), and slab ocean model run with fixed
827
ocean heat transport (SOM1850_texp_ocnht=cnt -SOM1850_cnt). Years 1-20 are shown for
828
all slab ocean experiments, years 1-29 are shown for FC1850_cnt and years 150-200 are
829
used for FC1850_texp. See Table 1 for model run descriptions.
830
831
Figure 12. Annual mean subsidence climatology (black contours) and subsidence
832
response (colors): a) slab ocean model run with fixed ocean heat transport
833
(SOM1850_cnt, SOM1850_cnt-SOM1850_texp_ocnht=cnt.), b) fully coupled model
834
(FC1850_cnt, FC1850_texp-FC1850_cnt), c) slab ocean model with prescribed ocean
835
heat transport changes (SOM1850_cnt, SOM1850_cnt-SOM1850_texp_ocnht=texp).
836
Years 1-20 are shown for all slab ocean experiments, years 1-29 are shown for FC1850_cnt
837
and years 150-200 are used for FC1850_texp. See Table 1 for model run descriptions.
838
839
Figure 13. Zonal annual mean precipitation a) climatology in fully coupled control
840
(FC1850_cnt, years 1-29) and Global Precipitation Climatology Project observations
841
(1979-2009) (Adler et al. 2003), b) ANN response, c) DJF response. Response shown
842
for fully coupled run (FC1850_texp - FC1850_cnt), slab ocean model run with changes in
843
ocean heat transport (SOM1850_texp_ocnht=texp - SOM1850_cnt), and slab ocean model
844
run without changes in ocean heat transport (SOM1850_texp_ocnht=cnt - SOM1850_cnt).
845
Years 1-20 are shown for all slab ocean experiments, years 1-29 are shown for FC1850_cnt
846
and years 150-200 are used for FC1850_texp. See Table 1 for model run descriptions.
847
848
Figure 14. Schematic showing climate response to absorbed shortwave radiation
849
bias reduction (less ASR in Southern Ocean, more ASR in the Tropics): a) with fixed
850
ocean heat transport, b) with dynamic ocean heat transport. All indicated changes
851
are anomalies relative to a mean state climate.
852
853
38
854
Name
(years)
Climate
Forcing
(year)
Description
ATM2000_cnt
10
2000
Atmosphere-only control with prescribed 2000
surface ocean
ATM2000_exp
10
2000
Atmosphere-only experiment (Tice =253 K) with
prescribed 2000 surface ocean
FC1850_cnt
29
1850
Fully coupled control, Initial condition January
1, year 402 of CESM-LE 1850 control run
b.e11.B1850C5CN.f09_g16.005 (Kay et al. 2015)
FC1850_exp
29
1850
Fully coupled experiment (Tice =253 K ), Initial
condition January 1, year 402 of CESM-LE 1850
control run b.e11.B1850C5CN.f09_g16.005 (Kay
et al. 2015)
FC1850_texp
200
1850
Fully coupled tuned experiment (Tice =253 K),
Initial condition January 1, year 402 of CESM-LE
1850 control run
b.e11.B1850C5CN.f09_g16.005 (Kay et al. 2015)
SOM1850_cnt
20
1850
Slab ocean model control with prescribed ocean
heat transport from FC1850_cnt, Initial
condition January 1, year 402 of CESM-LE 1850
control run b.e11.B1850C5CN.f09_g16.005 (Kay
et al. 2015)
SOM1850_texp_ocnht=texp
20
1850
Slab ocean model tuned experiment (Tice =253
K) with prescribed ocean heat transport from
FC1850_texp, Initial condition January 1, year
150 of FC1850_texp
SOM1850_texp_ocnht=cnt
20
1850
Slab ocean model tuned experiment (Tice =253
K) with prescribed ocean heat transport from
FC1850_cnt, Initial condition January 1, year
402 of CESM-LE 1850 control run
b.e11.B1850C5CN.f09_g16.005 (Kay et al. 2015)
Table 1. Description of global climate model runs. All runs use the Community Earth
855
System Model with the Community Atmosphere Model version 5 (CESM-CAM5, Hurrell et
856
al. 2013) at 1-degree horizontal resolution with the finite volume dynamical core.
857
858
39
859
Top-of-
model
Energy
Imbalance
(Wm-2)
Shortwave
Cloud
Radiative
Effect
(Wm-2)
Longwave
Cloud
Radiative
Effect
(Wm-2)
Total
Cloud
Fraction
(%)
Grid-box
Cloud
Liquid
Water Path
(g m-2)
Surface
Tempera
ture
(K)
CERES-EBAF
2000-2013
-47.2
26.0
n/a
UWisc.
1987-2000
87.2
FC1850_cnt
Yrs. 1-29
0.3
-47.7 (13.4)
22.6 (6.2)
63.1%
40.1 (60.9)
287.2
FC1850_exp
Yrs. 1-29
-0.9
-49.9 (13.0)
23.0 (5.8)
64.4%
45.4 (54.8)
286.6
FC1850_texp
Yrs. 1-29
0.0
-46.6 (10.9)
22.3 (6.1)
62.9%
44.0 (55.8)
287.5
FC1850_texp
Yrs. 150-200
0.0
-46.7 (10.9)
22.3 (6.1)
62.9%
44.0 (55.7)
287.5
Table 2. Global annual averages for fully coupled model runs. Observations are from
860
CERES-EBAF v2.8 (Loeb et al. 2009) and UWisc. (O’Dell et al. 2008). Root mean square
861
differences between model runs and observations are provided in parentheses. See Table 1
862
for a description of model runs.
863
864
865
866
Total
CHT
(PW)
Atmospheric
CHT
(PW)
Oceanic
CHT
(PW)
Tropical
Precipitation
Asymmetry Index
Observationally
Constrained
2001-2012
(Loeb et al. 2015)
0.20 ± 0.05
-0.24 ± 0.04
0.44 ± 0.07
0.20
FC1850_cnt
Yrs. 1-29
0.26
-0.18
0.44
0.11
FC1850_texp
Yrs. 1-29
0.13
-0.19
0.32
0.12
FC1850_texp
Yrs. 150-200
0.11
-0.21
0.32
0.15
Table 3. Cross-equatorial Heat Transport (CHT) and Tropical Precipitation
867
Asymmetry Index. The tropical precipitation asymmetry index (TPAI) is defined following
868
Hwang and Frierson (2013) as precipitation in the Northern Hemisphere tropics (0-20 °N
869
area-averaged) minus precipitation in the Southern Hemisphere tropics (0-20 °S area
870
averaged) normalized by the tropical mean precipitation (20 °S 20 °N area averaged). See
871
Table 1 for a description of model runs.
872
873
40
874
Figure 1. Atmospheric circulation, clouds, and air temperatures over the Southern
875
Ocean in the fully coupled control run (FC1850_cnt, Table 1). The left column shows
876
cloud fraction (colors) and subsidence (black lines) for the annual mean ANN (a), Southern
877
Hemisphere winter JJA (b) and Southern Hemisphere summer DJF (c). The middle column
878
shows grid-box cloud liquid mixing ratio (colors) and air temperature (black lines) for ANN
879
(d), JJA (e) DJF (f). The right column shows grid-box cloud ice mixing ratio (colors) and air
880
temperature (black lines) for ANN (g), JJA (h) and DJF (i).
881
41
882
Figure 2. Annual mean cloud liquid tendencies over the subsidence (30-50 °S, solid)
883
and ascent (50-70 °S, dash) regions of the Southern Ocean in the fully coupled
884
control run (FC1850_cnt, Table 1): a) vertical diffusion and all moist physics
885
tendencies, b) microphysics tendencies. Note: The total tendency due to moist physics
886
processes is the sum of MPDLIQ, MACPDLIQ, SHDLFLIQ CMFDLIQ, ZMDLIQ, and DPDLFLIQ.
887
The total tendency due to microphysical processes is the sum of MPDW2P, MPDW2I,
888
MPDW2V, QCSEDTEN, and QCSEVAP.
889
890
42
891
Figure 3. Global difference maps for atmosphere-only runs (ATM2000_exp
892
ATM2000_cnt): a) Absorbed shortwave radiation (ASR) difference for annual mean
893
(ANN) b) as in a) but for Southern Hemisphere summer (DJF), c) ANN grid-box mean
894
cloud liquid water path (LWP), d) as in c) but for DJF. See Table 1 for model run
895
descriptions.
896
897
43
898
899
Figure 4. Zonal mean Absorbed Shortwave Radiation bias over the Southern Ocean in
900
atmosphere-only runs. Values in parentheses indicate Southern Ocean (30-70 ° S) bias.
901
Observations are from CERES-EBAF (Loeb et al. 2009) v2.8 for the years 2000-2013. See
902
Table 1 for model run descriptions. Both annual means (ANN) and Southern Hemisphere
903
summer (DJF) means are plotted.
904
905
44
906
Figure 5. Zonal global mean absorbed shortwave radiation in fully coupled runs: a)
907
bias, b) difference from control (FC1850_cnt). Observations are from CERES-EBAF v2.8
908
(Loeb et al. 2009) for the years 2000-2013. An area-weighted axis is used to show the
909
balancing of mid-latitude and tropical ASR biases and differences. See Table 1 for model
910
run description. Both annual means (ANN) and Southern Hemisphere summer (DJF) means
911
are plotted.
912
913
914
45
915
Figure 6. Global difference maps for fully coupled climate model runs (FC1850_texp
916
FC1850_cnt): a) Annual mean (ANN) absorbed shortwave radiation (ASR) difference,
917
b) as in a) but for Southern Hemisphere summer (DJF), c) ANN grid-box mean cloud
918
liquid water path (LWP) difference, d) as in c) but for DJF. See Table 1 for model run
919
description. An average of years 150-200 is shown for FC1850_texp, while an average of
920
years 1-29 is shown for FC1850_cnt.
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
46
938
Figure 7. Global zonal vertical mean temperature and circulation in fully coupled
939
runs: a) Annual mean air temperature (black contours, FC1850_cnt) and
940
temperature response (colors, FC1850_texp - FC1850_cnt), b) as in a) but for
941
Southern Hemisphere summer (DJF), c) Annual zonal mean wind (black contours,
942
FC1850_cnt) and zonal mean wind response (colors, FC1850_texp - FC1850_cnt), d)
943
as in c) but for DJF. See Table 1 for model run descriptions. An average of years 150-200
944
is shown for FC1850_texp, while an average of years 1-29 is shown for FC1850_cnt.
945
946
947
948
949
47
950
Figure 8. Annual mean northward heat transport in fully coupled runs: a)
951
Climatology in fully coupled control run (FC1850_cnt, solid) and tuned experiment
952
(FC1850_texp, dash) b) Response to increased detrainment of supercooled liquid in
953
shallow convective clouds and tuning (FC1850_texp- FC1850_cnt). See Table 1 for
954
model run descriptions. An average of years 150-200 is shown for FC1850_texp, while an
955
average of years 1-29 is shown for FC1850_cnt. Heat transport calculated using methods
956
described in Kay et al. 2012b Appendix A. Total heat transport was calculated based on the
957
fact that, in steady state, horizontal heat flux convergence across a latitude band is
958
balanced by the net TOA flux poleward of that latitude. Ocean heat transport was calculated
959
inline. Atmospheric heat transport was calculated by integrating the residual between the
960
top-of-atmosphere and surface heat fluxes.
961
48
962
963
Figure 9. Surface zonal wind stress and wind stress response: a) climatology in fully
964
coupled control (FC1850_cnt, years 1-29), b) response to increased detrainment of
965
supercooled liquid in shallow convective clouds and tuning (FC1850_texp-
966
FC1850_cnt) using years 1-30 of FC1850_texp, c) as in b) but using years 150-200 of
967
FC1850_texp.
968
49
969
Figure 10. Meridional Overturning Circulation (MOC): a) climatology in fully coupled
970
control (FC1850_cnt, years 1-29), b) response to increased detrainment of
971
supercooled liquid in shallow convective clouds and tuning (FC1850_texp-
972
FC1850_cnt) using years 1-30 of FC1850_texp, c) as in b) but using years 150-200 of
973
FC1850_texp.
974
975
50
976
Figure 11. Atmospheric northward heat transport response to increased
977
detrainment of supercooled liquid in shallow convective clouds and tuning: a) total,
978
b) latent and dry static energy. Responses are shown for fully coupled model
979
(FC1850_texp - FC1850_cnt), slab ocean model with prescribed ocean heat transport
980
changes (SOM1850_texp_ocnht=texp - SOM1850_cnt), and slab ocean model run with fixed
981
ocean heat transport (SOM1850_texp_ocnht=cnt -SOM1850_cnt). Years 1-20 are shown for
982
all slab ocean experiments, years 1-29 are shown for FC1850_cnt and years 150-200 are
983
used for FC1850_texp. See Table 1 for model run descriptions.
984
51
985
Figure 12. Annual mean subsidence climatology (black contours) and subsidence
986
response (colors): a) slab ocean model run with fixed ocean heat transport
987
climatology (SOM1850_cnt) and response (SOM1850_cnt-SOM1850_texp_ocnht=cnt.),
988
b) fully coupled model climatology (FC1850_cnt) and response (FC1850_texp-
989
FC1850_cnt), c) slab ocean model with prescribed ocean heat transport changes
990
climatology (SOM1850_cnt) and response (SOM1850_cnt-
991
SOM1850_texp_ocnht=texp). Years 1-20 are shown for all slab ocean experiments, years
992
1-29 are shown for FC1850_cnt and years 150-200 are used for FC1850_texp. See Table 1
993
for model run descriptions.
994
52
995
Figure 13. Zonal annual mean precipitation a) climatology in fully coupled control
996
(FC1850_cnt, years 1-29) and Global Precipitation Climatology Project observations
997
(1979-2009) (Adler et al. 2003), b) ANN response, c) DJF response. Response shown
998
for fully coupled run (FC1850_texp - FC1850_cnt), slab ocean model run with changes in
999
ocean heat transport (SOM1850_texp_ocnht=texp - SOM1850_cnt), and slab ocean model
1000
run without changes in ocean heat transport (SOM1850_texp_ocnht=cnt - SOM1850_cnt).
1001
Years 1-20 are shown for all slab ocean experiments, years 1-29 are shown for FC1850_cnt
1002
and years 150-200 are used for FC1850_texp. See Table 1 for model run descriptions.
1003
53
1004
Figure 14. Schematic showing climate response to absorbed shortwave radiation
1005
bias reduction (less ASR in Southern Ocean, more ASR in the Tropics): a) with fixed
1006
ocean heat transport, b) with dynamic ocean heat transport. All indicated changes
1007
are anomalies relative to a mean state climate.
1008
1009
1010
1011
... The large uncertainty of ITCZ location change is mostly related to the uncertainty in the responses of clouds and sea ice. This ITCZ uncertainty persists even after ocean dynamics are included in models, where a dynamic ocean may mediate the extratropical influences on the ITCZ through changes in ocean heat transport 15,16 . For instance, the Coupled Model Intercomparison Project phase 5 (CMIP5) climate models show diverse ITCZ responses to a quadrupling of atmospheric CO 2 concentration 17,18 . ...
... We realize that the double ITCZ bias remains an issue in several generations of climate models [49][50][51][52][53] . This ITCZ bias has been suggested to be linked to Southern Ocean cloud bias 50,52 ; however, the teleconnection between the Southern Ocean and tropical precipitation biases is muted by adjustments in energy transports in the coupled climate system 15,16 . Furthermore, a direct relationship between the mean-state double ITCZ bias and ITCZ changes is not statistically significant 54 . ...
Article
Full-text available
Migrations of the intertropical convergence zone (ITCZ) have significant impacts on tropical climate and society. Here we examine the ITCZ migration caused by CO2 increase using climate model simulations. During the first one to two decades, we find a northward ITCZ displacement primarily related to an anomalous southward atmospheric cross-equatorial energy transport. Over the next hundreds or thousands of years, the ITCZ moves south. This long-term migration is linked to delayed surface warming and reduced ocean heat uptake in the Southern Ocean, which alters the interhemispheric asymmetry of ocean heat uptake and creates a northward atmospheric cross-equatorial energy transport anomaly. The southward ITCZ shift, however, is reduced by changes in the net energy input to the atmosphere at the equator by about two-fifths. Our findings highlight the importance of Southern Ocean heat uptake to long-term ITCZ evolution by showing that the (quasi-)equilibrium ITCZ response is opposite to the transient ITCZ response.
... BL clouds enhance downwelling shortwave reflection and absorb upwelling longwave radiation, which affects the surface and top-of-atmosphere energy budgets, especially in the Arctic and Southern Ocean (Bennartz et al., 2013;Bodas-Salcedo et al., 2016; D. L. Hartmann & Short, 1980;McCoy et al., 2017). To properly simulate the top-of-the-atmosphere and surface energy budgets in climate and numerical weather prediction models, it is crucial to accurately model the properties and morphological characteristics of BL clouds during CAOs (Abel et al., 2017;Bennartz et al., 2013;Bodas-Salcedo et al., 2012Forbes & Ahlgrimm, 2014;Kay et al., 2016;McCoy et al., 2017;Williams et al., 2013). ...
Article
Full-text available
The accurate representation of Cold Air Outbreaks (CAOs) and affiliated mixed‐phase boundary layer (BL) clouds in models is challenging. How BL cloud properties evolve during CAOs and their dependence on meteorological conditions is not well understood but is important for the simulation of Earth's energy budgets. Here the properties of polar BL clouds over the North Atlantic (NA) and Southern Ocean (SO) are compared using observations from the Measurements of Aerosol Radiation and CloUds over the SO (MARCUS) and CAOs in the Marine BL Experiment (COMBLE) conducted over the NA. MARCUS observations show a stronger BL inversion than COMBLE, with a higher mean EIS (estimated inversion strength)/LTS (lower tropospheric stability) of −0.03 K/13 K compared to COMBLE’s −3.2 K/9.3 K. 39% of CAOs observed during COMBLE were intense with M > 5 K, while MARCUS only had 1.3%. 78%/72% of clouds sampled in CAOs during COMBLE/MARCUS had cloud top heights <4 km. The mean BL cloud top height was over 400 m higher, and the BL was over 500 m deeper for M of 10 K compared to 0 K for both regions. MARCUS observed a 27% moister BL structure than COMBLE when M > 5 K due to stronger BL inversion trapping more moisture within the BL. Under the same LTS, EIS, and M conditions, MARCUS observed a 12% drier BL structure, and clouds were 46% more turbulent than COMBLE. During CAOs, 54% of single‐layer BL clouds sampled during MARCUS had liquid‐dominated bases compared to 39% during COMBLE.
... Existing biases in CESM2: Ensemble mean bias of 11 members of CESM2 historical simulations from CMIP6 data set in (a) JJAS precipitation, and 850 hPa wind with respect to GPCP (Adler et al., 2018) and ERA5 (Hersbach et al., 2020), respectively; (b) JJAS 200 hPa velocity potential relative to ERA5; and (c) annual mean Sea Surface Temperature relative to ERSST (Huang et al., 2015b); and (d) annual mean net downward top-of-atmosphere radiation (F TOA ) flux relative to CERES-EBAF (Loeb et al., 2018). ASR have little effect on tropical precipitation (Kay et al., 2016) but the same perturbation applied in lower latitudes shifts the ITCZ (Xiang et al., 2018). We hypothesize that SH subtropical ASR bias could alter the ITCZ locations during JJAS and affect the SASM precipitation. ...
Article
Full-text available
We study the sensitivity of South Asian Summer Monsoon (SASM) precipitation to Southern Hemisphere (SH) subtropical Absorbed Solar Radiation (ASR) changes using Community Earth System Model 2 simulations. Reducing positive ASR biases over the SH subtropics impacts SASM, and is sensitive to the ocean basin where changes are imposed. Radiation changes over the SH subtropical Indian Ocean (IO) shifts rainfall over the equatorial IO northward causing 1–2 mm/day drying south of equator, changes over the SH subtropical Pacific increases precipitation over northern continental regions by 1–2 mm/day, and changes over the SH subtropical Atlantic have little effect on SASM precipitation. Radiation changes over the subtropical Pacific impacts the SASM through zonal circulation changes, while changes over the IO modify meridional circulation to bring about changes in precipitation over northern IO. Our findings suggest that reducing SH subtropical radiation biases in climate models may also reduce SASM precipitation biases.
... This is strongly underestimated when employing CCN/INPclim-mean. Representing liquid clouds in the SO is a particularly common challenge for climate models in general (Bodas-Salcedo et al., 2016;Kay et al., 2016;Mc-Cluskey et al., 2023), but it is not obvious why the mean climatology would deviate so strongly from the default simulation. ...
Article
Full-text available
Aerosol particles influence cloud formation and properties. Hence climate models that aim for a physical representation of the climate system include aerosol modules. In order to represent more and more processes and aerosol species, their representation has grown increasingly detailed. However, depending on one's modelling purpose, the increased model complexity may not be beneficial, for example because it hinders understanding of model behaviour. Hence we develop a simplification in the form of a climatology of aerosol concentrations. In one approach, the climatology prescribes properties important for cloud droplet and ice crystal formation, the gateways for aerosols to enter the model cloud microphysics scheme. Another approach prescribes aerosol mass and number concentrations in general. Both climatologies are derived from full ECHAM-HAM simulations and can serve to replace the HAM aerosol module and thus drastically simplify the aerosol treatment. The first simplification reduces computational model time by roughly 65 %. However, the naive mean climatological treatment needs improvement to give results that are satisfyingly close to the full model. We find that mean cloud condensation nuclei (CCN) concentrations yield an underestimation of cloud droplet number concentration (CDNC) in the Southern Ocean, which we can reduce by allowing only CCN at cloud base (which have experienced hygroscopic growth in these conditions) to enter the climatology. This highlights the value of the simplification approach in pointing to unexpected model behaviour and providing a new perspective for its study and model development.
... When these elements are combined, the SO can serve as a proxy for a pre-industrial environment, providing a natural testbed for aerosol-cloudprecipitation interactions (ACPI) [5][6][7] . Yet the limited understanding of SO clouds results in radiation budget biases in both reanalysis products and climate simulations [8][9][10][11][12][13][14] . ...
Article
Full-text available
Southern Ocean (SO) air is amongst the most pristine on Earth, particularly during winter. Historically, there has been a focus on biogenic sources as an explanation for the seasonal cycle in cloud condensation nuclei concentrations ( N CCN ). N CCN is also sensitive to the strength of sink terms, although the magnitude of this term varies considerably. Wet deposition, a process encompassing coalescence scavenging (drizzle formation), is one such process that may be especially relevant over the SO. Using a boundary layer cloud climatology, N CCN and precipitation observations from Kennaook/Cape Grim Observatory (CGO), we find a statistically significant difference in N CCN between when the upwind meteorology is dominated by open mesoscale cellular convection (MCC) and closed MCC. When open MCC is dominant, a lower median N CCN (69 cm ⁻³ ) is found compared to when closed MCC (89 cm ⁻³ ) is dominant. Open MCC is found to precipitate more heavily (1.72 mm day ⁻¹ ) and more frequently (16.7% of the time) than closed MCC (0.29 mm day ⁻¹ , 4.5%). These relationships are observed to hold across the seasonal cycle with maximum N CCN and minimum precipitation observed during Austral summer (DJF). Furthermore, the observed MCC morphology strongly depends on meteorological conditions. The relationship between N CCN and precipitation can be further examined across a diurnal cycle during the summer season. Although there was again a negative relationship between precipitation and N CCN , the precipitation cycle was out of phase with the N CCN cycle, leading it by ~3 hours, suggesting other factors, specifically the meteorology play a primary role in influencing precipitation.
... Climate models continue to have large uncertainties in the cloud forcing over the SO, including their inability to reproduce the correct cloud phase, supercooled liquid cloud opacity, and cold cloud processes in the region (Cesana et al., 2022). The poor simulation of cloud forcing in SO, for instance, the abundance of supercooled liquid water in SO clouds, also lead to the significant biases of climate models in representing solar radiation budget over the region (Bodas-Salcedo et al., 2014Kay et al., 2016). These large uncertainties and biases have limited the ability of the models to represent important local climate features and their teleconnections such as surface warming, storm activity, and precipitation patterns (Ceppi et al., 2014;McFarquhar et al., 2021;Vergara-Temprado et al., 2018). ...
Article
Full-text available
Large satellite discrepancies and model biases in representing precipitation over the Southern Ocean (SO) are related directly to the region's limited surface observations of precipitation. To help address this knowledge gap, the study investigated the precipitation characteristics and rain rate retrievals over the remote SO using ship‐borne data of the Ocean Rainfall And Ice‐phase precipitation measurement Network disdrometer (OceanRAIN) and dual‐polarimetric C‐band radar (OceanPOL) aboard the Research Vessel (RV) Investigator in the Austral warm seasons of 2016–2018. Seven distinct synoptic types over the SO were analyzed based on their radar polarimetric signatures, surface precipitation phase, and rain microphysical properties. OceanRAIN observations revealed that the SO precipitation was dominated by drizzle and light rain, with small‐sized raindrops (diameter <1 mm) constituting up to 47% of total accumulation. Precipitation occurred most frequently over the warm sector of extratropical cyclones, while concentrations of large‐sized raindrops (diameter >3 mm) were prominent over synoptic types with colder and more convectively unstable environments. OceanPOL observations complement and extend the surface precipitation properties sampled by OceanRAIN, providing unique information to help characterize the variety of potential precipitation types and associated mechanisms under different synoptic conditions. Raindrop size distributions (DSD) measured with OceanRAIN over the SO were better characterized by analytical DSD forms with two‐shape parameters than single‐shape parameters currently implemented in satellite retrieval algorithms. This study also revised a rainfall retrieval algorithm for C‐band radars to reflect the large amount of small drops and provide improved radar rainfall estimates over the SO.
... Until recently, earth system models, including those participating in CMIP5 (Coupled Model Intercomparison Project Phase 5), and reanalysis products overestimated the occurrence of ice and had insufficient liquid cloud cover over the SO, leading to a large shortwave radiation bias in the region (Bodas-Salcedo et al., 2013;Gettelman et al., 2020;Kay et al., 2016aKay et al., , 2016bNaud et al., 2014). This imbalance was hypothesized to be due to over-prediction of cloud glaciation and ice precipitation processes (Frey & Kay, 2017;Kay et al., 2016a;Mace et al., 2020;McFarquhar et al., 2021;Tan et al., 2016;Vergara-Temprado et al., 2018). ...
Article
Full-text available
Supercooled liquid clouds are ubiquitous over the Southern Ocean (SO), even to temperatures below −20°C, and comprise a large fraction of the marine boundary layer (MBL) clouds. Earth system models and reanalysis products have struggled to reproduce the observed cloud phase distribution and occurrence of cloud ice in the region. Recent simulations found the microphysical representation of ice nucleation and growth has a large impact on these properties, however, measurements of SO ice nucleating particles (INPs) to validate simulations are sparse. This study presents measurements of INPs from simultaneous aircraft and ship campaigns conducted over the SO in austral summer 2018, which include the first in situ observations in and above cloud in the region. Our results confirm recent observations that INP concentrations are uniformly lower than measurements made in the late 1960s. While INP concentrations below and above cloud are similar, higher ice nucleation efficiency above cloud supports model simulations that the dominant INP composition varies with height. Model parameterizations based solely on aerosol properties capture the mean relationship between INP concentration and temperature but not the observed variability, which is likely related to the only modest correlations observed between INPs and environmental or aerosol metrics. Including wind speed in addition to activation temperature in a marine INP parameterization reduces bias but does not explain the large range of observed INP concentrations. Direct and indirect inference of marine INP size suggests MBL INPs, at least during Austral summer, are dominated by particles with diameters smaller than 500 nm.
Preprint
Full-text available
Ice-nucleating particles (INPs) have an important function in the freezing of clouds, but are rare in East Antarctica with concentrations between 6 × 10-6 L-1 and 5 × 10-3 L-1 observed at the Belgian Princess Elisabeth Station. These low concentrations offer a possible explanation for the occurrence of supercooled liquid water in clouds observed using the station's Micro Rain Radar and Ceilometer. We used COSMO-CLM² with an added aerosol-cycle module to test the cloud phase’s sensitivity in response to varying prescribed INP concentrations. We tested two cases, one in austral summer, one in austral winter, and analysed the differences resulting from INP concentration changes for an area around the station and over the Southern Ocean within the selected domain. Our results show a strong influence of the INP concentration on the liquid water path in both regions, with higher concentrations reducing the amount of liquid water. Over the ocean, this effect is stronger during winter: During summer, a significant portion of water remains in liquid state regardless of INP concentration. Over the continent, this effect is stronger during summer: Temperatures in winter frequently fall below -37 °C, allowing homogeneous freezing. The largest increase of the liquid water fraction of total cloud hydrometeor mass is simulated over the ocean in winter, from 9.8 % in the highest tested INP concentration to 50.3 % in the lowest. The radiative effects caused by the INP concentration changes are small with less than 3 W m-2 difference in the averages between different concentrations.
Article
Full-text available
Process-oriented observational constraints for the anthropogenic effective radiative forcing due to aerosol–cloud interactions (ERFaci) are highly desirable because the uncertainty associated with ERFaci poses a significant challenge to climate prediction. The contoured frequency by optical depth diagram (CFODD) analysis supports the evaluation of model representation of cloud liquid-to-rain conversion processes because the slope of a CFODD, generated from joint MODerate Resolution Imaging Spectroradiometer (MODIS)-CloudSat cloud retrievals, provides an estimate of cloud droplet collection efficiency in single-layer warm liquid clouds. Here, we present an updated CFODD analysis as an observational constraint on the ERFaci due to warm rain processes and apply it to the U.S. Department of Energy's Energy Exascale Earth System Model version 2 (E3SMv2). A series of sensitivity experiments shows that E3SMv2 droplet collection efficiencies and ERFaci are highly sensitive to autoconversion, i.e., the rate of mass transfer from cloud liquid to rain, yielding a strong correlation between the CFODD slope and the shortwave component of ERFaci (ERFaciSW; Pearson's R=-0.91). E3SMv2's CFODD slope (0.20 ± 0.04) is in agreement with observations (0.20 ± 0.03). The strong sensitivity of ERFaciSW to the CFODD slope provides a useful constraint on highly uncertain warm rain processes, whereby ERFaciSW, constrained by MODIS-CloudSat, is estimated by calculating the intercept of the linear association between the ERFaciSW and the CFODD slopes, using the MODIS-CloudSat CFODD slope as a reference.
Article
Full-text available
The evaluation and quantification of Southern Ocean cloud–radiation interactions simulated by climate models are essential in understanding the sources and magnitude of the radiative bias that persists in climate models for this region. To date, most evaluation methods focus on specific synoptic or cloud-type conditions that do not consider the entirety of the Southern Ocean's cloud regimes at once. Furthermore, it is difficult to directly quantify the complex and non-linear role that different cloud properties have on modulating cloud radiative effect. In this study, we present a new method of model evaluation, using machine learning that can at once identify complexities within a system and individual contributions. To do this, we use an XGBoost (eXtreme Gradient Boosting) model to predict the radiative bias within a nudged version of the Australian Community Climate and Earth System Simulator – Atmosphere-only model, using cloud property biases as predictive features. We find that the XGBoost model can explain up to 55 % of the radiative bias from these cloud properties alone. We then apply SHAP (SHapley Additive exPlanations) feature importance analysis to quantify the role each cloud property bias plays in predicting the radiative bias. We find that biases in the liquid water path are the largest contributor to the cloud radiative bias over the Southern Ocean, though important regional and cloud-type dependencies exist. We then test the usefulness of this method in evaluating model perturbations and find that it can clearly identify complex responses, including cloud property and cloud-type compensating errors.
Article
Full-text available
The Southern Ocean is a critical region for global climate, yet large cloud and solar radiation biases over the Southern Ocean are a long-standing problem in climate models and are poorly understood, leading to biases in simulated sea surface temperatures. This study shows that supercooled liquid clouds are central to understanding and simulating the Southern Ocean environment. A combination of satellite observational data and detailed radiative transfer calculations is used to quantify the impact of cloud phase and cloud vertical structure on the reflected solar radiation in the Southern Hemisphere summer. It is found that clouds with supercooled liquid tops dominate the population of liquid clouds. The observations show that clouds with supercooled liquid tops contribute between 27% and 38% to the total reflected solar radiation between 40° and 70°S, and climate models are found to poorly simulate these clouds. The results quantify the importance of supercooled liquid clouds in the Southern Ocean environment and highlight the need to improve understanding of the physical processes that control these clouds in order to improve their simulation in numerical models. This is not only important for improving the simulation of present-day climate and climate variability, but also relevant for increasing confidence in climate feedback processes and future climate projections.
Article
Full-text available
Climate models produce an increase in cloud optical depth in midlatitudes associated with climate warming, but the magnitude of this increase and its impact on reflected solar radiation vary from model to model. Transition from ice to liquid in midlatitude clouds is thought to be one mechanism for producing increased cloud optical depth. Here observations of cloud properties are used from a suite of remote sensing instruments to estimate the effect of conversion of ice to liquid associated with warming on reflected solar radiation in the latitude band from 40 degrees to 60 degrees S. The calculated increase in upwelling shortwave radiation (SW up arrow) is found to be important and of comparable magnitude to the increase in SW up arrow associated with warming-induced increases of optical depth in climate models. The region where the authors' estimate increases SW up arrow extends farther equatorward than the region where optical depth increases with warming in models. This difference is likely caused by other mechanisms at work in the models but is also sensitive to the amount of ice present in climate models and its susceptibility to warming.
Article
Full-text available
Satellite based top-of-atmosphere (TOA) and surface radiation budget observations are combined with mass corrected vertically integrated atmospheric energy divergence and tendency from reanalysis to infer the regional distribution of the TOA, atmospheric and surface energy budget terms over the globe. Hemispheric contrasts in the energy budget terms are used to determine the radiative and combined sensible and latent heat contributions to the cross-equatorial heat transports in the atmosphere (AHT_EQ) and ocean (OHT_EQ). The contrast in net atmospheric radiation implies an AHT_EQ from the northern hemisphere (NH) to the southern hemisphere (SH) (0.75 PW), while the hemispheric difference in sensible and latent heat implies an AHT_EQ in the opposite direction (0.51 PW), resulting in a net NH to SH AHT_EQ (0.24 PW). At the surface, the hemispheric contrast in the radiative component (0.95 PW) dominates, implying a 0.44 PW SH to NH OHT_EQ. Coupled model intercomparison project phase 5 (CMIP5) models with excessive net downward surface radiation and surface-to-atmosphere sensible and latent heat transport in the SH relative to the NH exhibit anomalous northward AHT_EQ and overestimate SH tropical precipitation. The hemispheric bias in net surface radiative flux is due to too much longwave surface radiative cooling in the NH tropics in both clear and all-sky conditions and excessive shortwave surface radiation in the SH subtropics and extratropics due to an underestimation in reflection by clouds.
Article
Full-text available
We review the effects of dynamical variability on clouds and radiation in observations and models and discuss their implications for cloud feedbacks. Jet shifts produce robust meridional dipoles in upper-level clouds and longwave cloud-radiative effect (CRE), but low-level clouds, which do not simply shift with the jet, dominate the shortwave CRE. Because the effect of jet variability on CRE is relatively small, future poleward jet shifts with global warming are only a second-order contribution to the total CRE changes around the midlatitudes, suggesting a dominant role for thermodynamic effects. This implies that constraining the dynamical response is unlikely to reduce the uncertainty in extratropical cloud feedback. However, we argue that uncertainty in the cloud-radiative response does affect the atmospheric circulation response to global warming, by modulating patterns of diabatic forcing. How cloud feedbacks can affect the dynamical response to global warming is an important topic of future research.
Article
Full-text available
The role of ocean-atmosphere coupling in the zonal-mean climate response to projected late twenty-firstcentury Arctic sea ice loss is investigated using Community Climate System Model version 4 (CCSM4) at 18 spatial resolution. Parallel experiments with different ocean model configurations (full-depth, slab, and no interactive ocean) allow the roles of dynamical and thermodynamic ocean feedbacks to be isolated. In the absence of ocean coupling, the atmospheric response to Arctic sea ice loss is confined to north of 308N, consisting of a weakening and equatorward shift of the westerlies accompanied by lower tropospheric warming and enhanced precipitation at high latitudes. With ocean feedbacks, the response expands to cover the whole globe and exhibits a high degree of equatorial symmetry: the entire troposphere warms, the global hydrological cycle strengthens, and the intertropical convergence zones shift equatorward. Ocean dynamics are fundamental to producing this equatorially symmetric pattern of response to Arctic sea ice loss. Finally, the absence of a poleward shift of the wintertime Northern Hemisphere westerlies in CCSM4's response to greenhouse gas radiative forcing is shown to result from the competing effects of Arctic sea ice loss and greenhouse warming on the meridional temperature gradient in middle latitudes.
Article
Full-text available
A modified microphysics scheme is implemented in the Community AtmosphereModel, version 5 (CAM5). The new scheme features prognostic precipitation. The coupling between the microphysics and the rest of the model is modified to make it more flexible. Single-column tests show the new microphysics can simulate a constrained drizzling stratocumulus case. Substepping the cloud condensation (macrophysics) within a time step improves single-column results. Simulations of mixed-phase cases are strongly sensitive to ice nucleation. The new microphysics alters process rates in both single-column and global simulations, even at low (200 km) horizontal resolution. Thus, prognostic precipitation can be important, even in low-resolution simulations where advection of precipitation is not important. Accretion dominates as liquid water path increases in agreement with cloud-resolving model simulations and estimates from observations. The new microphysics with prognostic precipitation increases the ratio of accretion over autoconversion. The change in process rates appears to significantly reduce aerosol-cloud interactions and indirect radiative effects of anthropogenic aerosols by up to 33% (depending on substepping) to below 1Wm-2 of cooling between simulations with preindustrial (1850) and present-day (2000) aerosol emissions.
Article
Full-text available
We assess the uptake, transport and storage of oceanic anthropogenic carbon and heat over the period 1861 to 2005 in a new set of coupled carbon-climate Earth System models conducted for the fifth Coupled Model Intercomparison Project (CMIP5), with a particular focus on the Southern Ocean. Simulations show the Southern Ocean south of 30°S, occupying 30% of global surface ocean area, accounts for 43 ± 3% (42 ± 5 Pg C) of anthropogenic CO2 and 75 ± 22% (23 ± 9 •1022J) of heat uptake by the ocean over the historical period. Northward transport out of the Southern Ocean is vigorous, reducing the storage to 33 ± 6 Pg anthropogenic carbon and 12 ± 7•1022J heat in the region. The CMIP5 models as a class tend to underestimate the observation-based global anthropogenic carbon storage, but simulate trends in global ocean heat storage over the last fifty years within uncertainties of observation-based estimates. CMIP5 models suggest global and Southern Ocean CO2 uptake have been largely unaffected by recent climate variability and change. Anthropogenic carbon and heat storage show a common broad-scale pattern of change, but ocean heat storage is more structured than ocean carbon storage. Our results highlight the significance of the Southern Ocean for the global climate and as the region where models differ the most in representation of anthropogenic CO2 and in particular heat uptake.
Article
Spaceborne lidar observations from the Cloud-Aerosol Lidar and Infrared Pathfinder Satellite Observation (CALIPSO) satellite are used to evaluate cloud amount and cloud phase in the Community Atmosphere Model version 5 (CAM5), the atmospheric component of a widely used state-of-the-art global coupled climate model (Community Earth System Model). By embedding a lidar simulator within CAM5, the idiosyncrasies of spaceborne lidar cloud detection and phase assignment are replicated. As a result, this study makes scale-aware and definition-aware comparisons between model-simulated and observed cloud amount and cloud phase. In the global mean, CAM5 has insufficient liquid cloud and excessive ice cloud when compared to CALIPSO observations. Over the ice-covered Arctic Ocean, CAM5 has insufficient liquid cloud in all seasons. Having important implications for projections of future sea level rise, a liquid cloud deficit contributes to a cold bias of 2-3°C for summer daily maximum near-surface air temperatures at Summit, Greenland. Over the midlatitude storm tracks, CAM5 has excessive ice cloud and insufficient liquid cloud. Storm track cloud phase biases in CAM5 maximize over the Southern Ocean, which also has larger-than-observed seasonal variations in cloud phase. Physical parameter modifications reduce the Southern Ocean cloud phase and shortwave radiation biases in CAM5 and illustrate the power of the CALIPSO observations as an observational constraint. The results also highlight the importance of using a regime-based, as opposed to a geographic-based, model evaluation approach. More generally, the results demonstrate the importance and value of simulator-enabled comparisons of cloud phase in models used for future climate projection.