Coupled high-latitude climate feedbacks and

1
Coupled high-latitude climate feedbacks and their impact on atmospheric
heat transport
2
3
Nicole Feldl⇤
4
Department of Earth and Planetary Sciences, University of California, Santa Cruz, California
5
Simona Bordoni
6
Environmental Science and Engineering, California Institute of Technology, Pasadena, California
7
Timothy M. Merlis
8
Department of Atmospheric and Oceanic Sciences, McGill University, Montreal, Quebec,
9
Canada
10
⇤ Corresponding
author address: University of California, Santa Cruz, 1156 High Street, Earth and
11
Planetary Sciences, Santa Cruz, CA 95064
12
E-mail: [email protected]
Generated using v4.3.2 of the AMS LATEX template
1
ABSTRACT
13
The response of atmospheric heat transport to anthropogenic warming is de-
14
termined by the anomalous meridional energy gradient. Feedback analysis
15
offers a characterization of that gradient and hence reveals how uncertainty
16
in physical processes translates into uncertainty in the circulation response.
17
However, individual feedbacks do not act in isolation. Anomalies associated
18
with one feedback may be compensated by another, as is the case for the pos-
19
itive water vapor and negative lapse rate feedbacks in the tropics. Here we
20
perform a set of idealized experiments in an aquaplanet model to evaluate the
21
coupling between the surface albedo feedback and other feedbacks, including
22
the impact on atmospheric heat transport. In the tropics, the dynamical re-
23
sponse manifests as changes in the intensity and structure of the overturning
24
Hadley circulation. Only half of the range of Hadley cell weakening exhibited
25
in these experiments is found to be attributable to systematic variations in the
26
surface albedo feedback. Changes in extratropical clouds that accompany the
27
albedo changes explain the remaining spread. This finding has implications
28
for the efficacy with which the climate system, including tropical circulation
29
and the hydrological cycle, adjusts to high latitude feedbacks, over climate
30
states that range from perennial or seasonal ice to ice-free conditions in the
31
Arctic.
2
32
1. Introduction
33
Climate feedbacks have long been recognized as a key piece to understanding the Earth’s cli-
34
mate sensitivity. Climate sensitivity is the amount of global-mean surface temperature change
35
for a given external forcing, typically defined as a doubling of CO2 . On time scales relevant
36
to anthropogenic warming, feedbacks include atmospheric processes, such as changes in clouds,
37
water vapor, atmospheric lapse rate, and sea ice, which in turn either amplify or damp the cli-
38
mate response to a forcing (Charney et al. 1979; Hansen et al. 1984; Schlesinger 1985). While
39
conventionally defined relative to the globally averaged case, these processes exhibit rich spatial
40
structures and are arguably activated by regional rather than global-mean warming (Armour et al.
41
2012). An emerging emphasis in the field of climate dynamics is to understand how the spatial
42
pattern of climate feedbacks controls the spatial pattern of climate change.
43
In addition to amplifying or damping the climate response, climate feedbacks exhibit two ad-
44
ditional characteristic behaviors: nonlinearities and remote impacts. Firstly, from a Taylor series
45
perspective, nonlinearities may be understood as higher-order terms in the energy balance (Colman
46
et al. 1997; Roe 2009). For instance, both longwave radiative fluxes (Stefan-Boltzmann law) and
47
atmospheric moisture content (Clausius-Clapeyron relation) are nonlinear functions of tempera-
48
ture. However, nonlinearities may also enter as interactions among feedbacks, which introduces
49
bias in the first-order linear approximation. Secondly, remote impacts are a consequence of pos-
50
itive feedbacks amplifying the local energy balance (or negative feedbacks damping it), which
51
the atmosphere then accommodates by diverging and converging energy flux meridionally (Feldl
52
and Roe 2013b; Roe et al. 2015; Zelinka and Hartmann 2012). Hence it is probable that regional
53
interactions between feedbacks affect remote climate responses in a nonlinear manner.
3
54
Perturbations to the energy balance, caused directly and indirectly by enhanced greenhouse gas
55
concentrations, affect the meridional energy flux and may manifest as changes in the strength or
56
position of the tropical Hadley circulation or the large-scale extratropical eddies. In a previous
57
study, we characterized the response of the tropical atmospheric circulation in CMIP5 in terms
58
of climate feedbacks, radiative forcing, ocean heat uptake, atmospheric eddies, and gross moist
59
stability (Feldl and Bordoni 2016). For the most part these effects are large and competing. At
60
the level of individual feedbacks, it becomes difficult to grasp intuitively how any single feedback
61
affects the circulation response given their interactive nature. For example, compensation between
62
lapse rate and water vapor feedbacks is well known (Bony et al. 2006; Cess 1975). Assuming
63
the feedbacks act in isolation, we can quantify the contribution of the temperature feedback as a
64
5-10%/K weakening of the Hadley cell and of the lapse rate feedback as a 5-10%/K strengthening.
65
But does that level of specificity matter when the net result is no change at all?
66
Naturally, such results are still useful for understanding uncertainty amongst climate change
67
projections, and moreover any particular coupled model is not required to cancel so neatly. Herein
68
we present a series of idealized modeling experiments designed to reveal the coupling between
69
regional climate feedbacks at a mechanistic level. This is conceptually similar to a perturbed
70
physics ensemble (e.g., Sanderson et al. 2008). In contrast to studies such as Kang et al. (2009),
71
which forces the high latitudes by applying an ocean heat source and sink, or Graversen and Wang
72
(2009), which suppresses the surface albedo feedback, we modify the sensitivity of sea ice to
73
warming. Specifically, we manipulate the strength of the surface albedo feedback in order to (1)
74
identify compensating behavior in other feedbacks and (2) assess the efficacy of the net impact on
75
the remote climate response.
4
76
2. Methods
77
We employ the Geophysical Fluid Dynamics Laboratory Atmospheric Model (GFDL AM2.1;
78
Delworth et al. 2006) in its aquaplanet configuration with daily mean solar zenith angle. Sea-
79
sonally varying insolation corresponds to modern day parameters in the control runs, except for
80
eccentricity, which is set to zero. The model is coupled to a slab ocean of fixed depth (30m) with no
81
oceanic heat transport. Sea-ice formation is enabled by introducing an ocean albedo dependence
82
on surface temperature; the surface albedo is increased where surface temperatures are less than
83
270 K. Elsewhere, ocean albedo remains a function of zenith angle (Taylor et al. 1996). Control
84
CO2 concentration is 330 ppm. The surface albedo a is systematically varied (0.3, 0.4, 0.45, 0.5)
85
in each CO2 quadrupling experiment to manipulate the strength of the surface albedo feedback.
86
The eight simulations are integrated for 40 years, with the exception of IcyAqua05 (i.e., a = 0.5),
87
which is run for 45 years (including 15 years of spin up), to serve as a run from which the others
88
are branched. Monthly climatologies are computed from 30-year periods.
89
Feedbacks are calculated using the radiative kernel technique (Shell et al. 2008; Soden and Held
90
2006; Soden et al. 2008). Individual feedback parameters are the product of the radiative kernel
91
for the relevant climate variable, Ki = ∂ R/∂ xi , and the climate change anomaly, Dxi , normalized
92
by the local (i.e., zonal-mean seasonal-mean) surface air temperature response DTs to give units of
93
Wm
2
K 1:
li =
∂ R Dxi
∂ xi DTs
(1)
94
where R is the net radiative flux at the top of the atmosphere. Note that the conventional approach
95
for global feedbacks is to instead normalize by the global-mean surface warming, however, the
96
regional feedbacks offer a number of advantages where spatial patterns of warming are of interest
97
(Armour et al. 2012; Feldl and Roe 2013a). We consider the Planck, lapse rate, surface albedo,
5
98
water vapor, and cloud feedbacks. The Planck feedback is associated with a vertically uniform
99
warming of the surface and troposphere (x = Ts ), the lapse rate feedback with tropospheric warm-
100
ing that deviates from the vertically uniform profile (x = T 0 ), and the surface albedo feedback with
101
changes in surface albedo (x = a). For the water vapor feedback, the specific humidity anomaly
102
(x = ln(q)) is given relative to the standard anomaly for a 1 K warming assuming no change in
103
relative humidity. Finally, we compute the cloud feedback from the change in cloud radiative ef-
104
fect, DCRE, with corrections for cloud masking of non-cloud feedbacks, following Soden et al.
105
(2008). Radiative kernels are those of Soden and Held (2006) for the GFDL model, but zon-
106
ally averaged, the northern hemisphere mirrored in the south, and with a six-month lag to create
107
pseudo-seasonal-cycle-aquaplanet kernels.
108
109
The sum of the energy adjustments due to individual feedbacks, along with the radiative forcing
of CO2 , R f , characterizes the anomalous atmospheric energy balance,
DR = Â li DTs + R f + O(DTs2 )
(2)
i
110
Higher-order terms account for the nonlinearity, which is estimated as a residual. In an aqua-
111
planet without ocean dynamics and for equilibrium climate states, anomalous surface fluxes are
112
negligible. To maintain balance, latitudes of amplified atmospheric energy anomalously diverge
113
atmospheric heat flux, and those of damped energy converge atmospheric heat flux (hereafter heat
114
indicates moist static energy). Thus we can write the northward energy flux as the zonal and
115
meridional integral of feedbacks and forcing, and further split the total atmospheric flux into con-
116
tributions due to stationary and transient eddies, DFe , and the mean meridional circulation, DFHC ,
DFHC = DFe +
Z f Z 2p
p
2
0
(Â li DTs 0 + R f 0 ) a2 cos f dl df .
(3)
i
117
Primes (0 ) denote deviations from the global mean; a uniform feedback or forcing does not alter
118
tropical transports.
6
119
The tropical mass flux ymax and energy flux can further be related via the gross moist stability,
120
H = FHC /ymax , i.e., the effective energy stratification of the tropics. Combining Equation 3 with
121
gross moist stability and assuming small perturbations, we arrive at the following expression for
122
the fractional change in mass flux by the Hadley cell:
Dymax
=
ymax
RR
RR
Âi (li DTs )0 +
FHC
Rf 0
DFe
DH
.
H
(4)
123
Integrals are as in Equation 3. For a lengthier derivation, see Feldl and Bordoni (2016). Changes in
124
Hadley cell strength are thus quantified in terms of contributions from feedbacks, radiative forcing,
125
atmospheric eddies, and gross moist stability. Since we focus on fractional changes, the response
126
at latitudes of zero energy flux by the Hadley circulation (i.e. the cell edges) is ill-defined. Instead,
127
this analysis best captures the region of the streamfunction extrema. In the following section,
128
we perform a detailed evaluation of changes in circulation strength for each albedo experiment,
129
following Equation 4.
130
3. Results
131
a. Climate feedbacks and atmospheric heat transport
132
The four experiments, in which the albedo magnitude is specified but areal extent freely interac-
133
tive, exhibit a wide range of polar amplification under 4⇥CO2 (Figure 1a). High-latitude surface
134
temperature change is 5 K in the low-albedo experiment (IcyAqua03) and up to 24 K in the high-
135
albedo experiment (IcyAqua05). All of the simulations reside in the same ice-free equilibrium
136
climate state at 4⇥CO2 , and hence differences in surface warming reflect differences in the initial
137
mean state. In particular, the high-albedo experiment starts from the coldest climatology, with
138
year-round sea-ice cover to nearly 50 degrees latitude, and must warm substantially to reach the
7
139
same end point as the low-albedo experiment, which only ever forms seasonal (not perennial) sea
140
ice.
141
The different climate responses are reflected in the different climate feedbacks among the ex-
142
periments (Figure 1b). In the tropics, the net feedbacks are more similar, though small differences
143
in positive subtropical feedbacks may have a substantial effect on divergence of atmospheric heat
144
flux. However, striking differences in the feedbacks occur at high latitudes: the net feedback is
145
weakly negative (i.e., stabilizing) in the high-albedo experiment and strongly negative in the low-
146
albedo experiment. Naively, one might expect this difference to equal the magnitude of the surface
147
albedo feedback, since it is the surface albedo that was perturbed as the experimental design.
148
Figure 2 shows the annual-mean zonal-mean feedback parameters. The surface albedo feedback
149
varies in magnitude and location, with the high-albedo experiment extending the farthest equator-
150
ward (a signature of extensive sea ice retreat) and the low-albedo feedback being both weak and of
151
limited areal extent. The feedback magnitude ranges from 0-1.5 Wm 2 K 1 , consistent with cou-
152
pled models despite its idealized nature (Feldl and Bordoni 2016). Already, we see this difference
153
is not large enough to account for the spread in the net feedback in Figure 1b. The Planck feedback
154
is the same in all experiments, because DTs /DTs is unity so this is simply the temperature kernel.
155
The lapse rate feedback is positive in the high-albedo experiments and negative in the low-albedo
156
experiments, which we discuss in more detail later in the section. The well-known compensation
157
between water vapor and lapse rate feedbacks (Soden and Held 2006) leads to only small variabil-
158
ity in the combined temperature and water vapor feedback. This combined feedback is neutral for
159
the low-albedo experiment at high latitudes (not shown).
160
The cloud feedback is the only remaining feedback capable of contributing to the differences in
161
net feedbacks. Unlike the case for temperature and water vapor feedbacks, the combined effect
162
of surface albedo and cloud feedbacks increases rather than decreases the spread among the net
8
163
feedback (a positive covariance). This is evident in the high-albedo experiment having the most
164
positive surface albedo and net cloud feedback, whereas the low-albedo experiment has the most
165
negative of both feedbacks. At high latitudes, the SW cloud feedback is negative and the LW cloud
166
feedback is positive (not shown), resulting from complex interactions among clouds of different
167
thicknesses, heights, and optical properties (Ceppi et al. 2015; Zelinka et al. 2012).
168
Figure 3 shows the anomalous northward atmospheric energy fluxes implied by the spatial pat-
169
terns of feedbacks, radiative forcing, and atmospheric eddies, following Equation 3. The fluxes
170
are scaled by global-mean surface temperature change. Positive high latitude feedbacks pro-
171
duce anomalous energy divergence and an equatorward flux (lapse rate, surface albedo feedback)
172
whereas positive tropical and/or negative high latitude feedbacks (water vapor, net cloud feed-
173
backs) produce a poleward energy flux. Feldl and Bordoni (2016) show the same result for CMIP5.
174
As anticipated based on the structure of the feedbacks (Fig. 2), the surface albedo feedback and
175
the cloud feedback promote opposing tendencies in atmospheric heat flux, however, they do so in
176
a way that adds to rather than subtracts from the spread among feedbacks. At 45 N, for instance,
177
the spreads in the northward fluxes implied by the surface albedo and net cloud feedbacks are
178
comparable, 0.07 PW K
179
by temperature and water vapor feedbacks to 0.1 PW K 1 .
1
and 0.06 PW K
1
respectively. That combined spread is then reduced
180
The bottom panel of the figure shows the role of transient atmospheric eddies (stationary eddies
181
are nonexistent in the zonally symmetric aquaplanet) and radiative forcing. The anomalous flux
182
by eddies is poleward in all experiments in the tropics, but the extratropics show equatorward flux
183
in the high-albedo simulations, consistent with anomalous divergence from the sea-ice margin.
184
In not prescribing ocean heat flux divergence, the atmosphere bears the full brunt of transporting
185
heat meridionally. We anticipate the eddy heat flux to be an overestimate compared to coupled
186
atmosphere-ocean models. Finally, our choice of radiative forcing estimate contributes anomalous
9
187
divergence from the tropics and poleward heat flux, consistent with Huang and Zhang (2014),
188
which we discuss in more detail in Section 4.
189
In order to evaluate the impact of the coupled climate feedbacks on the strength of the mean
190
meridional atmospheric circulation, it is first helpful to understand the climatological mass and
191
energy fluxes in the aquaplanet simulations. Figure 4a shows the mass flux by the Hadley circu-
192
lation, ymax , in the 1⇥CO2 (solid lines) and 4⇥CO2 (dashed lines) climates. The peak mass flux
193
occurs in the subtropics, and the poleward edges of the Hadley cells are indicated by the latitudes
194
of zero mass flux (near 30 degrees). The high-albedo experiment has the strongest mean tropi-
195
cal circulation, and also weakens the most. The Hadley cell does expand under increased CO2 ,
196
consistent with theory and modeling (Held and Hou 1980; Korty and Schneider 2008; Levine and
197
Schneider 2015; Lu et al. 2007), however, the widening is the same in all four experiments (2 de-
198
grees). Hence the experimental setup induces variability in strength but not position of the Hadley
199
cell. The Intertropical Convergence Zone (ITCZ) remains at the equator in the annual-mean, on
200
account of the hemispherically symmetric model configuration. The northward energy flux by the
201
mean meridional circulation, FHC , is shown in solid lines in the lower panel.
202
A characteristic feature of the aquaplanet model is that, without the zonal asymmetries asso-
203
ciated with continents, stationary eddies do not form and cannot transport momentum, heat, or
204
moisture. In the absence of stationary eddies, transient eddy energy flux (dotted lines in Fig. 4b)
205
is calculated as the difference between the total atmospheric energy flux and the energy flux by the
206
mean meridional circulation.
207
Following Equation 4, the fractional changes in circulation strength are calculated as the ratio of
208
anomalous fluxes in Figure 3 to the climatological energy flux in Figure 4b (solid lines). Results
209
are averaged over the subtropical latitudes where the Hadley cell, and its changes, maximizes.
210
Figure 5a shows the circulation changes associated with feedbacks, radiative forcing, gross moist
10
211
stability, and atmospheric transient eddies. Total decreases (leftmost column) range from 1.1-1.9%
212
K 1 , with the high-albedo experiment exhibiting the largest changes in Hadley cell strength. The
213
greatest contributors to the circulation changes are feedbacks (strengthening tendency) and eddies
214
(strong weakening tendency). The other contributions are small and positive in most experiments.
215
The increase in eddy heat flux in the tropics (Fig. 3) is a consistent response to a positive net
216
feedback in the tropics. In previous work, the main compensation occurred instead between at-
217
mospheric eddies and ocean heat uptake (Feldl and Bordoni 2016). Here, in the absence of ocean
218
dynamics, that relationship is precluded.
219
Evaluating the response of the tropical circulation demonstrates the remote impact of coupled
220
climate feedbacks. Given our four experiments are differentiated only by their sea ice albedo
221
formulation, a null hypothesis would have been that the experiments differ only in their surface
222
albedo feedbacks and consequently that the 2.9% K
223
net feedback may be accounted for by the spread in circulation change due to the surface albedo
224
feedback alone. However, in Figure 5b we see that is not at all the case. The combined temperature
225
and water vapor feedback collapses to a strengthening tendency with minimal contribution to the
226
spread. The surface albedo feedback is a weakening tendency on the tropical circulation, consistent
227
with the equatorward anomalous energy flux evidenced in Figure 3. Notably, the spread in the
228
surface albedo feedback contribution is only 1.4% K
229
of circulation response. The remaining half is made up by the cloud feedback, a strengthening
230
tendency of 1.2-3.6% K 1 . In other words, the spread–or uncertainty–in surface albedo produces
231
changes in polar clouds that are additive with respect to the uncertainty in the tropical circulation
232
response.
233
234
Nonlinear processes account for 3 Wm
2
1
1
spread in circulation change due to the
and insufficient to explain the total range
of the global energy balance in the high-albedo exper-
iment. As a result, the effective climate sensitivity from the sum of the feedbacks underestimates
11
235
actual global-mean surface warming by 3 degrees (9.4 compared to 12.4 K). Largest nonlinearities
236
are found in the IcyAqua05 experiment. Since this experiment is most comparable to the climate
237
used to generate the radiative kernels, it is unlikely that the nonlinearity stems from a kernel-
238
simulation mismatch. Rather, we attribute the nonlinear term to interactions between feedbacks,
239
discussed in more detail below, as well as to processes that are nonlinear functions of surface tem-
240
perature. The nonlinearity is zero in IcyAqua03, which in addition to not having an appreciable
241
surface albedo feedback also exhibits the least warming.
242
b. Polar clouds and sea ice albedo
243
A documented relationship exists between clouds and sea ice, but the mechanisms involved
244
remain ambiguous. In an intermodel comparison, Huybers (2010) reports a negative covariance
245
between surface albedo and cloud feedback, though the cloud feedback is estimated as a residual
246
and may be biased. Mauritsen et al. (2013) similarly find that cloud and water vapor changes
247
dampen the effect of the surface albedo feedback. One interpretation of these previous studies is
248
that increased cloud fraction masks surface albedo changes, such that the positive surface albedo
249
feedback is weakened relative to clear-sky conditions. One might then anticipate the opposite
250
relationship (i.e., a positive covariance) to develop if the clouds are instead reduced.
251
Under a quadrupling of CO2 , we find that polar boundary layer clouds decrease strongly in
252
the high-albedo experiments (Figure 6). Climatologically in this region, the surface is cold and
253
the atmosphere, stable. Low cloud fraction is high (black lines). By contrast, the low-albedo
254
simulations are relatively warm, lack atmospheric temperature inversions, and have much weaker
255
cloud decks in the initial mean state. The decrease in low clouds, which becomes more pronounced
256
from the low- to high-albedo experiments, is a positive shortwave cloud radiative effect. Enhanced
12
257
liquid water path in all four experiments ensures an overall negative shortwave cloud feedback,
258
consistent with Ceppi et al. (2015).
259
The lower tropospheric stability, estimated as the difference in potential temperature between
260
700 hPa and the surface, further elucidates this response (Figure 7). As previously mentioned, the
261
mean 4⇥CO2 climate is identical among simulations (dashed lines). The low-albedo experiments
262
start from a weakly stable atmosphere that increases in stability to reach equilibrium. The high-
263
albedo experiments start from a strongly stable atmosphere that decreases in stability to reach the
264
final state. That decrease in stability is marked by a strong cloud response as the capping inver-
265
sion at the top of the boundary layer is eroded. The correlation between cloud fraction and lower
266
tropospheric stability is consistent with common cloud parameterizations (Rasch and Kristjánsson
267
1998; Slingo 1987). For reference, the AM2.1 model uses the cloud and convective parameteriza-
268
tions of Tiedtke (1993) and Moorthi and Suarez (1992), respectively.
269
The decrease in planetary albedo associated with these cloud changes is analogous to, and acts
270
in concert with, the decrease in surface albedo due to sea ice retreat. In both cases, the resulting
271
increase in absorbed shortwave radiation warms the boundary layer, contributing to the charac-
272
teristic bottom-heavy polar warming profile. In both cases, where clouds and sea ice are initially
273
more extensive, potential decreases may also be larger. And in both cases, uncertainty in the mag-
274
nitude of the response contributes to the total uncertainty in atmospheric heat transport. Critically,
275
while the SW cloud feedback is negative, the mechanisms described above render it less nega-
276
tive in the high-albedo experiment relative to the low-albedo experiment. Polar cloud and sea ice
277
changes thus amplify the energy imbalance at high latitudes, requiring less poleward heat transport
278
under enhanced CO2 . In the high-albedo experiments, that decrease is accomplished by decreases
279
in energy flux by both transient eddies and the mean meridional circulation. In the low-albedo
280
experiments, the weakening of the Hadley cell alone is sufficient.
13
281
4. Summary and Discussion
282
Feedbacks do not act in isolation. When we turn a positive feedback off, the net feedback be-
283
comes more negative than can be accounted for based on the strength of the missing feedback
284
alone. In the experiments presented herein, the effect of manipulating the surface albedo feed-
285
back is investigated. By restricting the formation of sea ice—in actuality a simple ocean albedo
286
dependence on surface temperature—we simulate the transition from perennial or seasonal ice
287
to ice-free conditions. For a warm initial mean state, the characteristic bottom-heavy structure of
288
Arctic warming is inhibited, and boundary layer clouds show slight increases rather than the strong
289
decreases that occur when the lower troposphere is destabilized. Importantly, while the coupled
290
climate feedbacks explored are extratropical in nature, the impacts are global. The net feedback
291
has a stronger meridional gradient in the low-albedo experiment, which reinforces the existing
292
pole-to-equator temperature gradient and promotes an anomalous poleward flux of energy. The
293
high-albedo experiment of IcyAqua05 behaves in the opposite manner. The strong surface albedo
294
feedback provokes an anomalous divergence of heat flux from the sea ice margin, the signature of
295
which is a decrease in energy flux by extratropical eddies as well as by the tropical Hadley cell.
296
Taken as a whole, the range in tropical circulation responses can thus be understood in terms of
297
the spatial pattern of coupled climate feedbacks.
298
An important result from the analysis is added insight into the interactive nature of surface
299
albedo and cloud feedbacks. Temperature and water vapor feedbacks are affected as well, though
300
to a lesser degree. In particular, we find that the lapse rate feedback is positive in the high-albedo
301
experiments, contributing to the evident polar amplification as other studies have also shown (Feldl
302
and Roe 2013b; Graversen et al. 2014; Pithan and Mauritsen 2014), and neutral/negative in the
303
low-albedo experiments. The sign change of the lapse rate feedback is consistent with work by
14
304
Cronin and Jansen (2016), in which the lapse rate feedback was found to be positive in the presence
305
of surface forcing and negative in the presence of warming by atmospheric heat transport. In our
306
case, boundary-layer warming is provided by the surface albedo feedback; in the absence of a
307
strong albedo feedback, the polar warming structure is instead dominated by meridional transport.
308
Ongoing research is aimed at uncovering the mechanisms responsible for the varying amounts of
309
polar amplification in the aquaplanet experiments.
310
Previous work by the authors has demonstrated that polar amplification persists (albeit in a
311
reduced form) even in the absence of a surface albedo feedback. In contrast, here we show that
312
warming in the low-albedo experiment is relatively uniform, with maxima at the equator and poles
313
(Fig. 1b). It is tempting to declare the discrepancy a consequence of the inclusion of a seasonal
314
cycle in the present study, however, the feedback structures and mean climates are different enough
315
to render a direct comparison challenging. Regardless, the combination of an extreme feedback
316
gradient and uniform warming pattern in Figure 1 is evidence of an atmosphere highly effective
317
at redistributing heat between latitudes. If this were not the case, the dominant response would be
318
one of tropical amplification, i.e., greatest warming where feedbacks are most positive.
319
The inclusion of a seasonal cycle also impacts the radiative forcing estimate (Fig. 8). Merlis
320
(2015a,b) has shown that time-independent and seasonally varying radiative forcing can have
321
different impacts on the annual-mean circulation response. Here we estimate the CO2 radiative
322
forcing as the ensemble-mean, zonal-mean, symmetrized and six-month-lagged forcing from the
323
CMIP5 sstClim4⇥CO2 minus sstClim experiments. This fixed-SST forcing includes seasonality,
324
which results in heat flux divergence from the equator. Were a time-invariant forcing instead used,
325
a localized region of convergence due to cloud masking would develop along the equator. Since
326
discussion of tropical circulation within the present study focuses on the regions of streamfunction
327
extrema, this distinction between time-invariant or seasonally varying forcing largely has little ef15
328
fect. For instance, the partial circulation changes due to radiative forcing are 0.2-0.5%/K for the
329
time-invariant forcing, compared to 0.3-0.9%/K for seasonally varying forcing.
330
We have emphasized the coupling between high-latitude cloud and surface albedo feedbacks.
331
Resulting differences in atmospheric heat transport are felt globally, including by the tropical
332
circulation, which must in turn mediate tropical clouds (Seo et al. 2014; Voigt and Shaw 2015).
333
At the edge of the tropics, the decrease in poleward heat flux implied by the extratropical cloud
334
feedback is 0.1 PW K
335
(Fig. 3). The edge of the tropics (i.e., 30 degrees) also exhibits differences in low-latitude cloud
336
feedbacks. Since the Hadley cell widens by the same amount in all four albedo experiments, the
337
spread is not due to a differential expansion of the subtropics. Rather, decreases in subsidence
338
are consistent with decreases in tropospheric cloud fraction, and both changes are greatest in the
339
high-albedo experiments.
1
and 0.06 PW K
1
in the low- and high-albedo experiments, respectively
340
Interactions among climate feedbacks have been implied based on documented nonlinearities.
341
Here, we demonstrate in a controlled modeling environment how perturbing one feedback, surface
342
albedo, affects the other feedbacks and atmospheric heat transport. Although nonlinear feedback
343
interactions and feedbacks that are nonlinear functions of temperature enter the energy balance at
344
the same order, we may be confident that the former dominates here: (1) The increase is longwave
345
emission with temperature, following Stefan-Boltzmann law, is a negative nonlinear feedback,
346
and our nonlinearity is positive. (2) The structure of the nonlinearity reveals a peak at approxi-
347
mately 60 degrees latitude, in a region of active sea ice and atmospheric processes. We express
348
the degree of coupling between feedbacks in terms of uncertainty in circulation response. For
349
instance, variability in the surface albedo feedback does indeed contribute to the weakening of the
350
Hadley cell, however the impact is larger than would be expected based on the magnitude of the
351
albedo feedback alone. The implication is that the strength of a feedback cannot be accurately
16
352
quantified in isolation. Hence decompositions that emphasize the climate response to individual
353
feedbacks must account for the energy balance residual, which captures the effect of the nonlinear
354
interactions among feedbacks.
355
Acknowledgments. NF was supported by the National Science Foundation (AGS-1524569).
356
References
357
Armour, K. C., C. M. Bitz, and G. H. Roe, 2012: Time-varying climate sensitivity from regional
358
359
360
361
362
363
364
365
366
367
368
369
feedbacks. Journal of Climate, 26 (13), 4518–4534, doi:10.1175/JCLI-D-12-00544.1.
Bony, S., and Coauthors, 2006: How Well Do We Understand and Evaluate Climate Change
Feedback Processes? Journal of Climate, 19 (15), 3445–3482, doi:10.1175/JCLI3819.1.
Ceppi, P., D. L. Hartmann, and M. J. Webb, 2015: Mechanisms of the negative shortwave cloud
feedback in high latitudes. Journal of Climate, 139–157, doi:10.1175/JCLI-D-15-0327.1.
Cess, R. D., 1975: Global climate change: an investigation of atmospheric feedback mechanisms.
Tellus, 27 (3), 193–198, doi:10.1111/j.2153-3490.1975.tb01672.x.
Charney, J. G., A. Arakawa, D. J. Baker, B. Bolin, and R. E. Dickinson, 1979: Carbon dioxide and
climate: A scientific assessment. Tech. rep., National Academy of Sciences.
Colman, R. A., S. B. Power, and B. J. McAvaney, 1997: Non-linear climate feedback analysis in
an atmospheric general circulation model. Climate Dynamics, 13 (10), 717–731.
Cronin, T. W., and M. F. Jansen, 2016:
Analytic radiative-advective equilibrium as a
370
model for high-latitude climate. Geophysical Research Letters, 43, 449–457, doi:10.1002/
371
2015GL067172.
17
372
Delworth, T. L., and Coauthors, 2006: GFDL’s CM2 global coupled climate models. Part
373
I: Formulation and simulation characteristics. Journal of Climate, 19 (5), 643–674, doi:
374
10.1175/JCLI3629.1.
375
376
377
378
379
380
Feldl, N., and S. Bordoni, 2016: Characterizing the Hadley circulation response through regional
climate feedbacks. Journal of Climate, 29, 613–622, doi:10.1175/JCLI-D-15-0424.1.
Feldl, N., and G. H. Roe, 2013a: Four perspectives on climate feedbacks. Geophysical Research
Letters, 40, doi:10.1002/grl.50711.
Feldl, N., and G. H. Roe, 2013b: The nonlinear and nonlocal nature of climate feedbacks. Journal
of Climate, 26, 8289–8304, doi:10.1175/JCLI-D-12-00631.1.
381
Graversen, R. G., P. L. Langen, and T. Mauritsen, 2014: Polar Amplification in CCSM4: Con-
382
tributions from the Lapse Rate and Surface Albedo Feedbacks. Journal of Climate, 27 (12),
383
4433–4450, doi:10.1175/JCLI-D-13-00551.1.
384
385
Graversen, R. G., and M. Wang, 2009: Polar amplification in a coupled climate model with locked
albedo. Climate Dynamics, doi:10.1007/s00382-009-0535-6.
386
Hansen, J., A. Lacis, D. Rind, G. Russell, P. Stone, I. Fung, R. Ruedy, and J. Lerner, 1984: Climate
387
sensitivity: Analysis of feedback mechanisms. Climate Processes and Climate Sensitivity, J. E.
388
Hansen, and T. Takahashi, Eds., AGU Geophysical Monograph 29, Maurice Ewing Vol. 5,
389
American Geophysical Union, Washington, D.C., 130–163.
390
391
392
393
Hansen, J., and Coauthors, 2005: Efficacy of climate forcings. Journal of Geophysical Research,
110 (D18).
Held, I. M., and A. Y. Hou, 1980: Nonlinear Axially Symmetric Circulations in a Nearly Inviscid
Atmosphere. Journal of the Atmospheric Sciences, 37 (3), 515–533.
18
394
Huang, Y., and M. Zhang, 2014: The implication of radiative forcing and feedback for merid-
395
ional energy transport. Geophysical Research Letters, 41 (2014), 1665–1672, doi:10.1002/
396
2013GL058954.
397
398
Huybers, P., 2010: Compensation between model feedbacks and curtailment of climate sensitivity.
Journal of Climate, 23, 3009–3018, doi:10.1175/2010JCLI3380.1.
399
Kang, S. M., D. M. W. Frierson, and I. M. Held, 2009: The tropical response to extrat-
400
ropical thermal forcing in an idealized GCM: The importance of radiative feedbacks and
401
convective parameterization. Journal of the Atmospheric Sciences, 66 (9), 2812–2827, doi:
402
10.1175/2009JAS2924.1.
403
404
Korty, R. L., and T. Schneider, 2008: Extent of Hadley circulations in dry atmospheres. Geophysical Research Letters, 35 (23), L23 803, doi:10.1029/2008GL035847.
405
Levine, X. J., and T. Schneider, 2015: Baroclinic Eddies and the Extent of the Hadley Circulation:
406
An Idealized GCM Study. Journal of the Atmospheric Sciences, 72 (7), 2744–2761, doi:10.
407
1175/JAS-D-14-0152.1.
408
409
Lu, J., G. A. Vecchi, and T. Reichler, 2007: Expansion of the Hadley cell under global warming.
Geophysical Research Letters, 34 (6), doi:10.1029/2006GL028443.
410
Mauritsen, T., R. G. Graversen, D. Klocke, P. L. Langen, B. Stevens, and L. Tomassini, 2013:
411
Climate feedback efficiency and synergy. Climate Dynamics, 41 (9-10), 2539—-2554, doi:10.
412
1007/s00382-013-1808-7.
413
414
Merlis, T. M., 2015a: Direct weakening of tropical circulations from masked CO2 radiative forcing. Proceedings of the National Academy of Sciences, 12 (43), 13 167–13 171.
19
415
Merlis, T. M., 2015b: Does humidity’s seasonal cycle affect the annual-mean tropical pre-
416
cipitation response to sulfate aerosol forcing?
417
10.1175/JCLI-D-15-0388.1.
Journal of Climate, 29, 1451–1460, doi:
418
Moorthi, S., and M. J. Suarez, 1992: Relaxed Arakawa-Schubert. A Parameterization of Moist
419
Convection for General Circulation Models. Monthly Weather Review, 120 (6), 978–1002, doi:
420
10.1175/1520-0493(1992)120h0978:RASAPOi2.0.CO;2.
421
422
Pithan, F., and T. Mauritsen, 2014: Arctic amplification dominated by temperature feedbacks in
contemporary climate models. Nature Geoscience, 7 (3), 181–184, doi:10.1038/ngeo2071.
423
Rasch, P. J., and J. E. Kristjánsson, 1998: A Comparison of the CCM3 Model Climate Using
424
Diagnosed and Predicted Condensate Parameterizations. Journal of Climate, 11 (7), 1587–1614,
425
doi:10.1175/1520-0442(1998)011h1587:ACOTCMi2.0.CO;2.
426
427
Roe, G. H., 2009: Feedbacks, Timescales, and Seeing Red. Annual Review of Earth and Planetary
Sciences, 37, 93–115, doi:10.1146/annurev.earth.061008.134734.
428
Roe, G. H., N. Feldl, K. C. Armour, Y.-T. Hwang, and D. M. W. Frierson, 2015: The remote
429
impacts of climate feedbacks on regional climate predictability. Nature Geosci, 8 (2), 135–139,
430
doi:10.1038/ngeo2346.
431
Sanderson, B. M., C. Piani, W. J. Ingram, D. A. Stone, and M. R. Allen, 2008: Towards constrain-
432
ing climate sensitivity by linear analysis of feedback patterns in thousands of perturbed-physics
433
GCM simulations. Climate Dynamics, 30, 175–190.
434
Schlesinger, M. E., 1985: Feedback analysis of results from energy balance and radiative-
435
convective models. The Potential Climatic Effects of Increasing Carbon Dioxide, M. C. Mac20
436
Cracken, and F. M. Luther, Eds., DOE/ER-0237, U. S. Department of Energy, Washington, DC,
437
280–319.
438
Seo, J., S. M. Kang, and D. M. W. Frierson, 2014: Sensitivity of intertropical convergence zone
439
movement to the latitudinal position of thermal forcing. Journal of Climate, 27 (8), 3035–3042,
440
doi:10.1175/JCLI-D-13-00691.1.
441
Shell, K. M., J. T. Kiehl, and C. A. Shields, 2008: Using the radiative kernel technique to calculate
442
climate feedbacks in NCAR’s Community Atmospheric Model. Journal of Climate, 21 (10),
443
2269–2282, doi:10.1175/2007JCLI2044.1.
444
Slingo, J. M., 1987: The Development and Verification of A Cloud Prediction Scheme For the
445
ECMWF Model. Quarterly Journal of the Royal Meteorological Society, 113 (477), 899–927,
446
doi:10.1002/qj.49711347710.
447
448
Soden, B. J., and I. M. Held, 2006: An assessment of climate feedbacks in coupled oceanatmosphere models. Journal of Climate, 19 (14), 3354–3360, doi:10.1175/JCLI3799.1.
449
Soden, B. J., I. M. Held, R. Colman, K. M. Shell, J. T. Kiehl, and C. A. Shields, 2008: Quantifying
450
climate feedbacks using radiative kernels. Journal of Climate, 21 (14), 3504–3520, doi:10.1175/
451
2007JCLI2110.1.
452
Taylor, J. P., J. M. Edwards, M. D. Glew, P. Hignett, and A. Slingo, 1996: Studies with a flexible
453
new radiation code. II: Comparisons with aircraft short-wave observations. Quarterly Journal
454
of the Royal Meteorological Society, 122 (532), 839–861, doi:10.1002/qj.49712253204.
455
456
Tiedtke, M., 1993: Representation of Clouds in Large-Scale Models. Monthly Weather Review,
121 (11), 3040–3061, doi:10.1175/1520-0493(1993)121h3040:ROCILSi2.0.CO;2.
21
457
458
Voigt, A., and T. A. Shaw, 2015: Circulation response to warming shaped by radiative changes of
clouds and water vapour. Nature Geoscience, 8, doi:10.1038/ngeo2345.
459
Zelinka, M. D., and D. L. Hartmann, 2012: Climate feedbacks and their implications for
460
poleward energy flux changes in a warming climate. Journal of Climate, 25, 608–624, doi:
461
10.1175/JCLI-D-11-00096.1.
462
Zelinka, M. D., S. A. Klein, and D. L. Hartmann, 2012: Computing and Partitioning Cloud Feed-
463
backs using Cloud Property Histograms. Part II: Attribution to Changes in Cloud Amount, Alti-
464
tude, and Optical Depth. Journal of Climate, 25, 3736–3754, doi:10.1175/JCLI-D-11-00249.1.
22
465
LIST OF FIGURES
466
Fig. 1.
467
468
469
Fig. 2.
470
471
472
Fig. 3.
473
474
475
Fig. 4.
476
477
478
479
480
Fig. 5.
481
482
483
484
485
486
487
488
489
490
491
498
499
500
501
502
.
25
Anomalous northward atmospheric energy fluxes, integrated from the south to north pole.
Each experiment is normalized by its global-mean surface temperature change to give units
of PW K 1 . Simulations are color-coded as in the legend of Fig. 1. . . . . . . .
.
26
(a) Mass flux in the 1⇥CO2 (solid lines) and 4⇥CO2 (dashed lines) climates. Mass flux is
calculated as the signed maximum magnitude of the meridional mass streamfunction. (b)
Energy flux by the mean meridional circulation (solid lines) and transient eddies (dotted
lines) in the 1⇥CO2 climate. Due to the absence of zonal asymmetries, stationary eddies do
not occur in the aquaplanet. Simulations are color-coded as in the legend of Fig. 1. . . .
. 27
(a) Changes in circulation strength implied by net feedback (fdbk), radiative forcing (forc),
gross moist stability (gms), and transient atmospheric eddies (eddy) averaged between 1525 degrees in both hemispheres. The total change (tot) is in the leftmost column and the
portion of the net feedback due to nonlinearities (nonl) in the rightmost column. The nonlinear contribution is estimated as the residual
RR between the anomalous TOA radiative flux
and the sum of feedbacks and forcing, FHC 1 DR0 (Si li DTs 0 + R f 0 ). Each experiment is
normalized by its global-mean surface temperature change to give units of %K 1 . CMIP5
results reproduced from Feldl and Bordoni (2016) are in grey. (b) The same but for individual climate feedbacks: temperature (T), water vapor (WV), combined temperature and water
vapor (T+WV), surface albedo (A), net cloud (C), and the longwave (LW C) and shortwave
(SW C) components of the cloud feedback. . . . . . . . . . . . . .
.
28
.
29
Fig. 7.
Annual-mean zonal-mean lower tropospheric stability calculated as the difference in potential temperature between 700 hPa and the surface in the 1⇥CO2 (solid lines) and 4⇥CO2
(dashed lines) climates. Simulations are color-coded as in the legend of Fig. 1. . . . .
.
30
Annual-mean zonal-mean radiative forcing for 4⇥CO2 . Stratosphere-adjusted radiative
forcing (dashed line) calculated as 2 ⇥ R f from the GFDL radiative transfer model based on
the perpetual equinox aquaplanet simulations of Feldl and Roe (2013b). Fixed-SST radiative forcing (solid line) calculated from the seasonally varying CMIP5 sstClim4⇥CO2 and
sstClim experiments. Symmetrized fixed-SST radiative forcing used in the present study
(dotted line); the global mean is 7.6 W m 2 . See Hansen et al. (2005) for more on forcing
definitions. . . . . . . . . . . . . . . . . . . . . . .
.
31
495
497
Annual-mean, zonal-mean regional feedbacks (Wm 2 K 1 ) for the four aquaplanet experiments. Simulations are color-coded as in the legend of Fig. 1. See Feldl and Bordoni (2016)
for the CMIP5 version of this figure. . . . . . . . . . . . . . . .
Annual-mean, zonal-mean change in cloud fraction (shading) and 1⇥CO2 climatology (lines
at 0.1 intervals) for the four aquaplanet experiments. . . . . . . . . . . .
494
496
. 24
Fig. 6.
492
493
a) Annual-mean zonal-mean change in near-surface temperature (K) under 4⇥CO2 for the
four aquaplanet experiments. Surface albedo values are 30% (purple), 40% (blue), 45%
(green), and 50% (yellow). (b) Zonal-mean net feedback in Wm 2 K 1 . . . . . . .
Fig. 8.
23
Surface temperature change
a
K
20
15
10
5
-60
-30
0
Latitude
30
60
Net feedback
1
b
W m -2 K -1
0
-1
-2
IcyAqua03
IcyAqua04
IcyAqua045
IcyAqua05
-3
-4
-5
-60
-30
0
Latitude
30
60
503
F IG . 1. a) Annual-mean zonal-mean change in near-surface temperature (K) under 4⇥CO2 for the four aqua-
504
planet experiments. Surface albedo values are 30% (purple), 40% (blue), 45% (green), and 50% (yellow). (b)
505
Zonal-mean net feedback in Wm 2 K 1 .
24
Planck
Temperature plus water vapor
0
W m -2 K -1
W m -2 K -1
0
-2
-4
-6
-4
-6
-60 -30
2
-2
0
30
60
-60 -30
Lapse rate
0
30
60
Surface albedo
W m -2 K -1
W m -2 K -1
4
0
-2
2
0
-4
-60 -30
30
60
Water vapor
-60 -30 0 30
Latitude
Net cloud
60
-60 -30
60
2
W m -2 K -1
W m -2 K -1
6
0
4
2
0
-2
0
-60 -30
0
30
60
0
30
506
F IG . 2. Annual-mean, zonal-mean regional feedbacks (Wm 2 K 1 ) for the four aquaplanet experiments.
507
Simulations are color-coded as in the legend of Fig. 1. See Feldl and Bordoni (2016) for the CMIP5 version of
508
this figure.
25
Planck
Temperature plus water vapor
0.05
PW/K
PW/K
0.1
0
-0.1
-0.05
PW/K
0
30
Latitude
Lapse rate
60
0.1
PW/K
-60 -30
0.1
0
0
-0.1
0
30
Latitude
Surface albedo
60
-60 -30
60
0
-0.1
-60 -30
0
30
Latitude
Water vapor
60
0
30
Latitude
Net cloud
0.1
PW/K
0.1
PW/K
-60 -30
0
-0.1
0
-0.1
-60 -30
0.2
0
30
Latitude
Eddy
60
-60 -30
0
30
Latitude
Forcing
60
-60 -30
0
30
Latitude
60
PW/K
PW/K
0.02
0
0
-0.02
-0.2
-60 -30
0
30
Latitude
60
509
F IG . 3. Anomalous northward atmospheric energy fluxes, integrated from the south to north pole. Each
510
experiment is normalized by its global-mean surface temperature change to give units of PW K 1 . Simulations
511
are color-coded as in the legend of Fig. 1.
26
Northward mass flux (10
9
kg s -1 )
a
200
100
0
-100
-200
-60
-30
0
Latitude
30
60
Northward energy flux (PW)
b
5
0
-5
-60
-30
0
Latitude
30
60
512
F IG . 4. (a) Mass flux in the 1⇥CO2 (solid lines) and 4⇥CO2 (dashed lines) climates. Mass flux is calculated as
513
the signed maximum magnitude of the meridional mass streamfunction. (b) Energy flux by the mean meridional
514
circulation (solid lines) and transient eddies (dotted lines) in the 1⇥CO2 climate. Due to the absence of zonal
515
asymmetries, stationary eddies do not occur in the aquaplanet. Simulations are color-coded as in the legend of
516
Fig. 1.
27
Partial circulation changes
12
a
10
8
6
4
% K -1
2
0
-2
-4
-6
-8
-10
-12
tot
fdbk
forc
gms eddy
ohu
nonl
IcyAqua03
IcyAqua04
IcyAqua045
IcyAqua05
cmip5
10
b
8
6
4
2
% K -1
0
-2
-4
-6
-8
-10
-12
-14
T
WV T+WV
A
C
LW C SW C
28
526
527
F IG . 6. Annual-mean, zonal-mean change in cloud fraction (shading) and 1⇥CO2 climatology (lines at 0.1
intervals) for the four aquaplanet experiments.
29
Lower tropospheric stability
22
20
18
K
16
14
12
10
8
-90
-60
-30
0
Latitude
30
60
90
528
F IG . 7. Annual-mean zonal-mean lower tropospheric stability calculated as the difference in potential tem-
529
perature between 700 hPa and the surface in the 1⇥CO2 (solid lines) and 4⇥CO2 (dashed lines) climates. Sim-
530
ulations are color-coded as in the legend of Fig. 1.
30
Radiative forcing
10
W m -2
8
6
4
2
-90
-60
-30
0
Latitude
30
60
90
531
F IG . 8. Annual-mean zonal-mean radiative forcing for 4⇥CO2 . Stratosphere-adjusted radiative forcing
532
(dashed line) calculated as 2 ⇥ R f from the GFDL radiative transfer model based on the perpetual equinox
533
aquaplanet simulations of Feldl and Roe (2013b). Fixed-SST radiative forcing (solid line) calculated from the
534
seasonally varying CMIP5 sstClim4⇥CO2 and sstClim experiments. Symmetrized fixed-SST radiative forcing
535
used in the present study (dotted line); the global mean is 7.6 W m 2 . See Hansen et al. (2005) for more on
536
forcing definitions.
31