1

Using an AGCM to diagnose historical effective radiative forcing and

2

mechanisms of recent decadal climate change

3 4

Timothy Andrews1,*

5

1

Met Office Hadley Centre, Exeter, UK.

6 7 8

Submitted: 24th May 2013

9

Revised: 6th September 2013

10 11

* Corresponding author address:

12

Met Office Hadley Centre

13

FitzRoy Road

14

Exeter

15

EX1 3PB

16

United Kingdom

17 18

Email: [email protected]

1

19

Abstract

20 21

An atmospheric general circulation model is forced with observed monthly sea-

22

surface-temperature and sea-ice boundary conditions, as well as forcing agents that

23

vary in time, for the period 1979-2008. The simulations are then repeated with

24

various forcing agents, individually and in combination, fixed at pre-industrial levels.

25

The simple experimental design allows the diagnosis of the models global and

26

regional time-varying effective radiative forcing from 1979 to 2008 relative to pre-

27

industrial. Furthermore the design can be used to (i) calculate the atmospheric

28

model’s feedback/sensitivity parameters to observed changes in sea-surface-

29

temperature, and (ii) separate those aspects of climate change that are directly driven

30

by the forcing from those that are driven by large-scale changes in sea-surface-

31

temperature. It is shown that the atmospheric response to increased radiative forcing

32

over the last 3 decades has halved the global precipitation response to surface

33

warming. Trends in sea-surface-temperature and sea-ice are found to contribute only

34

~60% of the global-land, northern hemisphere and summer land warming trends.

35

Global effective radiative forcing is ~1.5 Wm-2 in this model, with anthropogenic and

36

natural contributions of ~1.3 Wm-2 and ~0.2 Wm-2 respectively. Forcing increases by

37

~0.5 Wm-2 dec-1 over the period 1979-2008, or ~0.4 Wm-2 dec-1 if years strongly

38

influenced by volcanic forcings – which are non-linear with time - are excluded from

39

the trend analysis. Aerosol forcing shows little global decadal trend due to offsetting

40

regional trends whereby negative aerosol forcing weakens in Europe and North

41

America but continues to strengthen in South East Asia.

2

42

1. Introduction

43 44

Radiative forcings have long been used to quantify and rank the drivers of climate

45

change (e.g. Hansen et al., 1997; Shine and Forster, 1999). In climate models,

46

radiative forcings can help us to understand why different models differ in their

47

simulations of the past and future. For example Forster et al. (2013) found the

48

intermodel spread in global surface temperature change across Coupled Model

49

Intercomparison Project phase 5 (CMIP5) (Taylor et al., 2012) historical simulations

50

to be primarily driven by differences in their present day forcings. Kiehl et al. (2007)

51

showed an inverse relationship between near-present day radiative forcing and climate

52

sensitivity across older generation models, suggesting that models can capture 20th

53

century warming trends with different combinations of forcings and feedbacks.

54

Diagnosing radiative forcings in transient model simulations is therefore important for

55

understanding their coupled responses, and is a required first step to calculating their

56

transient feedbacks.

57 58

Many different forcing definitions exist (Hansen et al., 2005) and diagnosing them in

59

models requires either offline radiative transfer calculations (e.g. Forster et al., 1997;

60

Myhre et al. 1998) and/or targeted climate model experiments and diagnostics (e.g.

61

Gregory et al., 2004; Hansen et al., 2005). Diagnosing the time-evolving forcing in

62

transient scenarios is particularly difficult because forcing and feedback evolve

63

together. It is currently most easily achieved using a global energy budget approach

64

which assumes a climate feedback parameter to the derive the forcing (Forster and

65

Taylor, 2006; Forster et al., 2013). This method however, does not isolate individual

3

66

forcings or provide regional information, and assumes an invariant climate sensitivity

67

parameter.

68 69

The main goal of this paper is to provide and utilize a simple and efficient

70

experimental design, set within a clear conceptual model/definition of radiative

71

forcing, that could be readily used to (i) diagnose a model’s historical (or at least

72

1979-2008) time-varying regional radiative forcing. Then, given the forcing, (ii)

73

calculate climate sensitivity/feedback parameters in a realistic transient scenario, and

74

then (iii) separate those aspects of recent historical climate change that could be

75

usefully considered as part of the forcing (i.e. an ‘adjustment’, see below and Section

76

2.3) from those that are driven by large-scale changes in sea-surface-temperature

77

(SST) and associated climate feedbacks.

78 79

It will be shown that a simple extension of the well established Atmospheric Model

80

Intercomparison Project (AMIP) (Gates, 1992; Gates et al., 1999) design can provide

81

this information. Two experiments (at least) are needed, both using 1979 to near

82

present observed monthly SST and sea-ice boundary conditions, but one with forcing

83

agents varying in time from 1979 to near present and one with forcing agents fixed at

84

pre-industrial levels. Note that previous studies have to a limited extent used similar

85

designs for various purposes and are discussed in Section 2.3.

86 87

Diagnosing radiative forcing in this way is analogous to the “fixed-SST” (e.g. Hansen

88

et al., 2002; Shine et al., 2003a) definition of forcing, sometimes referred to as a

89

“quasi-forcing” (Rotstayn and Penner, 2001) or “radiative flux perturbation” (e.g.

90

Lohmann et al., 2010; Ming and Ramaswamy, 2012). This definition of forcing

4

91

includes not only the instantaneous radiative effect, but also any atmospheric and/or

92

land surface adjustments that come about rapidly – days to weeks (Dong et al., 2009;

93

Cao et al., 2012) - that are independent of large-scale changes in SST. Well

94

established examples of adjustments that are important for forcings include the

95

stratospheric temperature adjustment, aerosol-cloud interactions (i.e. aerosol indirect

96

effects) (e.g. Lohmann et al., 2010) and cloud adjustments to CO2 (e.g. Andrews et

97

al., 2012a). Radiative forcings diagnosed this way have been shown to be one of the

98

most appropriate definitions of forcing for predicting long term climate change (Shine

99

et al., 2003a; Hansen et al., 2005) and are referred to here as the ‘effective radiative

100

forcing’ (ERF).

101 102

2. Experimental Design

103 104

2.1 Atmospheric Model

105 106

The model used, HadGEM2-A, is the atmospheric component of the Met Office

107

Hadley Centre Global Environmental Model 2. HadGEM2-A includes atmospheric,

108

land surface and hydrology processes. It has 38 vertical levels and a horizontal

109

resolution of 1.25° x 1.875° in latitude and longitude. A detailed description and

110

validation of HadGEM2-A is given in Martin et al. (2011).

111 112

2.2. Experiments Performed

113 114

The base experiment follows the AMIP protocol for CMIP5 (Taylor et al., 2012).

115

This involves running an AGCM with observed monthly SST and sea-ice fractions

5

116

from 1979-2008, as well as time-evolving forcing agents. Note that this differs from

117

the original AMIP experiments that used a constant CO2 level (or CO2 equivalents)

118

and solar constant representative of the period being simulated (Gates, 1992). The

119

base experiment is termed AMIPSST,Ice,F to indicate it includes time-varying SST, sea-

120

ice and forcing agent boundary conditions.

121 122

The various forcing agents (WMGHGs, aerosols, O3, volcanoes etc.) included are

123

shown in Table 1. The forcing datasets and how they are implemented in this model

124

configuration are described in detail in Jones et al. (2012). Note that the forcings and

125

model configuration is not the same as the fully coupled atmosphere-ocean historical

126

simulation that was part of the Met Office Hadley Centre submission to CMIP5 using

127

the Earth system model HadGEM2-ES. For example, HadGEM2-A has prescribed

128

land surface types and O3 concentrations, whereas HadGEM2-ES has a dynamic

129

vegetation and interactive tropospheric chemistry scheme.

130 131

The experiment is then repeated with various forcing agents, in combination and

132

individually, set back to pre-industrial (1860) levels. Table 1 indicates the various

133

different experiments. AMIPSST,Ice indicates that the only time-varying boundary

134

conditions are SST and sea-ice, all forcing agents are fixed at pre-industrial levels.

135

The ‘no CO2’ experiment indicates that only CO2 has been set back to pre-industrial

136

and so on. Land-use over the period 1979-2008 is included in the prescribed surface

137

types but its forcing effect on climate compared to no land-use is not investigated in

138

this study. Differencing these experiments with the AMIPSST,Ice,F experiment gives

139

the impact of including the time-varying forcing agent/agents on the diagnostic of

140

interest, independent of any impacts through SST.

6

141 142

All experiments had a 4 month spin up period from September 1978 through to

143

December 1978, and are then run for 30 yrs to cover the period of analysis, 1979-

144

2008. Each experiment was run 3 times starting from different atmospheric initial

145

conditions. The analysis was preformed on the average of the 3-member ensemble,

146

except where individual members are used to assess the spread across the ensemble

147

(as indicated in the text).

148 149

2.3 Connection to Previous Studies

150 151

The experimental design is similar (though not identical) to some previous literature.

152

In particular, a similar design has been used in detection and attribution studies

153

(Folland et al., 1998; Sexton et al., 2001) and the Climate of Twentieth Century

154

(C20C) project (Folland et al., 2002). The C20C project used an ensemble of AGCMs

155

forced with observed SST and sea-ice distributions from the 1950s onwards, with

156

various combinations of forcing agents, to study climate variability and predictability

157

on the timescales of seasons to decades (Folland et al., 2002). Anderson et al. (2010;

158

2012) utilitized the C20C design to diagnose radiative forcing in models in a manor

159

similar to that presented here, but was limited to a 1950 baseline (rather than pre-

160

industrial) and lacked the systematic set of experiments needed to diagnose the

161

contributions of individual forcing agents and rapid adjustments.

162 163

Whilst most studies based on the C20C project did not consider the ERF conceptual

164

framework, some of the C20C results can be interpreted in this way. For example

165

Scaife et al. (2009) noted that those models that only included observed time-varying

7

166

SSTs, not time-varying forcing agents, generally underestimated the rapid recent land

167

warming since 1970. Under the ERF framework, this implies a significant role for

168

land temperature adjustments - independent of SST changes - in recent land warming

169

trends.

170 171

Similar experimental frameworks (i.e. using AGCMs forced with SST changes and

172

various forcing agent combinations) have been used to investigate mechanisms of

173

atmospheric circulation trends (Deser and Philips, 2009; King et al., 2010), regional

174

African climate change (e.g. Patricola and Cook, 2011; Skinner et al., 2012), changes

175

in the diurnal temperature range (Caminade and Terray, 2006), perturbations to the

176

hydrological cycle (e.g. Bosilovich et al., 2005; Bichet et al., 2011, Allan et al., 2013)

177

variability of the Indian monsoon rainfall (Kucharski et al., 2009) and decreasing land

178

wind speeds (e.g. Bichet et al. 2012). Most of these studies, in various ways, point to

179

the importance of including time-varying forcing agents – not just SST changes – in

180

simulating certain climate trends, and hence motivates further examination of

181

separating adjustments from those aspects of climate change that are mediated

182

through large-scale changes in SST.

183 184

The design and systematic set of experiments used in this study differs somewhat

185

from the above studies in order to maximise the utility of the experiments. For

186

example Folland et al. (1998) used a control experiment that had time-varying

187

observed SSTs and sea-ice from 1949, with CO2 levels (or CO2 equivalent fixed)

188

fixed at 1949 levels. They then successively introduced different time-varying

189

forcing agents to investigate the direct impact of these forcings on simulated climate

190

trends. In the design proposed here, forcings are removed from the simulation (rather

8

191

than introduced), and setting the forcings back to pre-industrial additionally provides

192

information on the ERF relative to pre-industrial, rather than 1950 as in Anderson et

193

al. (2010; 2012). This baseline is more useful, as the difference to pre-industrial fully

194

defines the anthropogenic contribution to radiative forcing, and is also closer to

195

equilibrium conditions. In addition, forcing datasets and models have significantly

196

improved since the early Folland et al. (1998) studies, so it is worth re-visiting some

197

of these issues, especially with a comprehensive model (HadGEM2-A) that compares

198

very well against a range of observational metrics (Jiang et al., 2012).

199 200

Whilst earlier studies have to a limited extent done similar work, none have provided

201

a framework and systematic set of experiments to diagnose the recent time-varying

202

ERF, including individual forcing contributions and regional variation, as well as

203

diagnose the models climate feedback/sensitivity parameters and adjustments as

204

mechanisms of recent historical climate change. This paper will - for the first time -

205

bring together and extend these ideas in detail within a single model, all set within our

206

improved conceptual models/definitions of radiative forcing. It is not meant to be

207

exhaustive, it will highlight only certain aspects: i) diagnosing a model’s time-varying

208

ERF and its climate sensitivity parameters (Sections 3 and 4), ii) diagnosing

209

adjustments as mechanisms of atmospheric and surface temperature trends (Sections

210

5.1 and 5.2), and iii) the relationship between radiative forcing and perturbations to

211

the hydrological cycle (Section 5.3).

212 213

3. Effective Radiative Forcing

214 215

3.1 Method

9

216 217

Fig. 1a shows the globally annually averaged net top-of-atmosphere radiative flux

218

timeseries (positive downward) for the AMIPSST,Ice,F and AMIPSST,Ice experiments.

219

Shading equals the range (max to min) across the 3 member ensemble.

220

Corresponding satellite (CERES-EBAF, Loeb et al., 2009) measurements from 2000

221

are indicated in red purely for illustrative purposes. Stephens et al. (2012) give an

222

observed planetary imbalance of 0.6 ± 0.4 Wm-2 averaged over the decade 2000-2010.

223 224

When HadGEM2-A is forced with observed monthly SST, sea-ice and forcing agents

225

(AMIPSST,Ice,F) the simulated global radiative fluxes are consistent with observed

226

estimates (Fig. 1a). When the experiment is repeated with only the time-varying SST

227

and sea-ice (all forcing agents set to pre-industrial, AMIPSST,Ice) the global radiation

228

balance is strongly negative. This difference, AMIPSST,Ice,F – AMIPSST,Ice (Fig. 1b),

229

gives the direct impact of including time-varying forcing agents relative to pre-

230

industrial on the global radiation balance and measures the ERF (see also Anderson et

231

al., 2010; 2012).

232 233

ERF diagnosed in this way can be derived through a global linearised energy budget

234

approach as follows (see also Forster et al., 2013, their Section 4.1). The net TOA

235

radiative imbalance (N, units Wm-2) can be described by changes in forcing (F, units

236

Wm-2) and various climate feedback processes that are linearly proportional to global-

237

mean surface-air-temperature change (∆T), such that N = F – α∆T, where α is the

238

climate feedback parameter (units Wm-2 K-1). Firstly, let this equation represent the

239

heat balance in the AMIPSST,Ice,F experiment, and N′ = F′ – α∆T′ the heat balance in

240

the AMIPSST,Ice experiment (distinguished by the primes). Defining the ERF relative

10

241

to pre-industrial makes F′ = 0 by construction (since AMIPSST,Ice has pre-industrial

242

forcing levels). As both experiments have the same SSTs, ∆T ~ ∆T′ (ignoring small

243

changes from the land-surface, see Section 5.2), then substituting for ∆T gives F = N -

244

N′ (as is plotted in Fig. 1b). This method of diagnosing forcing is analogous to fixed-

245

climatological-SST methods (e.g. Hansen et al., 2002) but generalised to time-varying

246

forcing agents and an evolving base-state (see Discussion).

247 248

An analogous calculation can be made for individual forcings, for example

249

differencing AMIPSST,Ice,F against the same experiments with CO2 levels fixed at pre-

250

industrial gives the CO2 ERF and so on. The analysis can be done regionally and in

251

radiative component terms (e.g. longwave, shortwave, all-sky (i.e. with clouds if

252

present) and clear-sky (clouds artificially set to zero)). The difference between all-sky

253

and clear-sky fluxes is used as a measure of the impact of clouds on the radiation

254

balance, termed the cloud radiative effect (CRE) (sometimes referred to as cloud

255

radiative forcing (CRF) in the literature). Note that changes in CRE can come about

256

through changes in cloud masking of clear-sky fluxes as well as through changes in

257

cloud properties (e.g. Soden et al., 2004; 2008).

258 259

3.2 Global Effective Radiative Forcing

260 261

Fig. 1b and 1c show the global-annual-mean ERF timeseries relative to pre-industrial

262

and its major components (WMGHGs, aerosol, O3 and natural). ERF at 2005

263

(defined as the 2003-2005 average) and its 1979-2008 decadal trends for all

264

experiments are tabulated in Tables 2 and 3, and compared on the same scale in Fig.

265

2.

11

266 267

The largest term is a positive ERF from WMGHGs (2.51 Wm-2) partially offset by

268

aerosol forcing (-1.38 Wm-2). These forcings act on different parts of the Earth’s

269

energy flows. GHGs predominately act through altering LW clear-sky fluxes, while

270

aerosol forcing acts on both SW clear-sky and SW CRE fluxes (Table 2), the latter

271

being consistent with aerosol-cloud interactions (Section 3.3). Globally averaged,

272

only WMGHGs demonstrate a large decadal trend (0.43 Wm-2 dec-1), with CO2

273

forcing increasing by 0.27 Wm-2 dec-1 between 1979 and 2008. Using annual-mean

274

CO2 measurements from the Mauna Loa Observatory (Keeling et al., 1976) with a

275

logarithmic formula for CO2 radiative forcing (Myhre et al., 1998) gives a similar

276

decadal trend (~0.24 Wm-2 dec-1 from 1979-2008) to that simulated, relative to the

277

pre-industrial CO2 level of 286 ppmv that the model uses.

278 279

Natural forcing from changes in solar activity and volcanic eruptions is strongly

280

interrupted by large negative spikes (Fig. 1b and Fig. 1c). These coincide with the

281

large volcanic eruptions of El Chichon (1982) and Pinatubo (1991) which inject

282

sulphur gas into the stratosphere, subsequently forming aerosols that scatter SW

283

radiation. A linear trend analysis of the natural forcing gives a decadal trend of 0.19

284

Wm-2 dec-1 (Table 3), but the volcanic forcing is strongly non-linear (Fig. 1c).

285

Removing the years most affected by large volcanic forcing (defined, somewhat

286

arbitrarily, as years in which the natural forcing is more negative than -0.25 Wm-2, i.e.

287

years 1982-84 and 1991-94) from the natural and all forcing analysis gives a natural

288

forcing trend much closer to zero, 0.05 Wm-2 dec-1.

289

12

290

Forster et al. (2013) diagnosed CMIP5 historical ERF through regression techniques

291

and assumptions on the global energy balance. The ERF presented here using the

292

AMIP design with HadGEM2-A (~ 1.5 Wm-2 at 2005) falls close to their multi-model

293

mean at 2003 (~ 1.7 ± 0.9 Wm-2). However they report a net ERF of ~ 0.8 Wm-2 for

294

HadGEM2-ES. While HadGEM2-A and HadGEM2-ES share the same physical

295

atmosphere, they are not identical models (Section 2), and this study has not

296

considered some forcings used in the fully coupled simulation, such as land-use.

297

Without an explicit comparison of the two methodologies under the same

298

model/forcings it is difficult to conclude whether this difference is real or not.

299 300

The global-mean net aerosol ERF (-1.38 Wm-2) is close to the combined radiative

301

transfer estimates of the aerosol direct effect (-0.16 Wm-2) and 1st indirect effect (-1.3

302

Wm-2) in HadGEM2-ES reported by Bellouin et al. (2011). This implies a relatively

303

small role for the combined effect of other responses that may additionally be

304

included in the ERF (such as aerosol 2nd indirect and semi-direct effects).

305 306

3.3 Regional Forcing

307 308

The geographical distribution of the time-averaged (1979-2008) ERF (Fig. 1d) is

309

dominated by positive forcing from WMGHGs over much of the globe. There exist

310

large negative regions which are attributable to aerosol forcing (mostly SO4) (Fig. 3).

311

Changes in aerosol optical depth (AOD) at 0.55µm (Fig. 3a) relative to pre-industrial

312

are dominated by increases over the industrialised regions of Europe, South East Asia

313

and North America. Biomass burning contributes significantly to changes over

314

Africa.

13

315 316

Aerosol ERF can be approximately split into direct aerosol-radiation interactions (e.g.

317

scattering and absorption of radiation) and aerosol-cloud interactions (e.g. aerosol

318

indirect effects, such as changes in cloud albedo and lifetime). The clear-sky and

319

CRE split approximates these processes, noting that it cannot, and is not meant to, be

320

exact (Lohmann et al., 2010). The increased AOD enhances SW scattering (and to a

321

lesser extent absorption), as seen in the negative SW clear-sky forcing (Table 3 and

322

Fig. 3c). Off the eastern coast of continents there exist large oceanic low cloud decks.

323

These are the regions of largest aerosol forcing (dominated by SO4 forcing) in

324

HadGEM2-A (Fig. 3b) and is dominated by changes in SW CRE (Fig. 3d), consistent

325

with aerosol-cloud interactions (namely the 1st indirect effect) which increases the

326

clouds’ albedo.

327 328

Despite a large aerosol forcing relative to pre-industrial, there is little to no global

329

decadal trend over the period 1979-2008. Regional analysis (Fig. 4) reveals large

330

offsetting regional trends. In South East Asia, AOD and forcing continue to

331

strengthen but it weakens in Europe and North America (Fig. 4a and 4b), as seen in

332

previous modelling studies (e.g. Shindell et al., 2013) and consistent with detailed

333

regional emission inventories derived from fuel use data and regional variations in

334

technology (Streets et al., 2009). Murphy (2013) used recent satellite measurements

335

and a radiative transfer model to show that a regional redistribution of aerosols have

336

little impact on clear-sky radiative forcing. Consistent with this the model produces

337

no decadal trend in global clear-sky aerosol ERF from 1979-2008 (Table 3) despite

338

some large regional trends (Fig. 4c).

339

14

340

4. Climate Resistance and Feedback Parameters

341 342

With the forcing now known it is possible to calculate the climate sensitivity

343

parameters of the model. These are usually diagnosed from idealised coupled model

344

experiments, such as step CO2 experiments and/or idealised AGCM experiments such

345

as +4K SST simulations used by the Cloud Feedback Model Intercomparison Project

346

2 (CFMIP2) (e.g. Bony et al., 2011). Diagnosing sensitivity parameters in more

347

realistic transient scenarios is desirable if we want these parameters to be more

348

directly comparable with observed estimates, and so would complement the idealised

349

approach. Note that with the AGCM design the coupled response is necessarily

350

omitted, so here the feedbacks represent the atmospheric response of the model to

351

‘perfect’ SSTs.

352 353

4.1. Climate Resistance

354 355

Gregory and Forster (2008) showed that under recent historical climate change and

356

projections of the 21st century F = ρ∆T is a good approximation, where ρ is the

357

‘climate resistance’ (units Wm-2 K-1). This relationship is tested in Fig. 5. Years

358

strongly influenced by volcanic forcing are given in red and are excluded from the

359

regression analysis. As in Gregory and Forster (2008) linearity between F and ∆T on

360

a multi-decadal timescale is a good approximation (r = 0.78) for years not strongly

361

influenced by volcanic forcing. The slope of the ordinary least square (OLS)

362

regression line gives an estimate of the climate resistance, ρ = 1.7 ± 0.29 Wm-2 K-1

363

(1σ uncertainty from regression), which is within the observed estimates covering a

364

similar period by Gregory and Forster (2008). As the coupled response is necessarily

15

365

omitted here, i.e. the model has been forced with observed ∆T, this agreement implies

366

that the derived estimate of F varies in a similar manor to that of Gregory and Forster

367

(2008).

368 369

4.2 Climate Feedback Parameters

370 371

Diagnosing feedbacks in transient simulations - or observations – is difficult because

372

forcing/adjustments and feedback evolve together. Forcings first need to be removed

373

so they do not contaminate the feedback signal. Returning to the global linearised

374

energy budget equation gives N – F = –α∆T. Hence, now that F has been derived, the

375

feedback parameter, -α, can be diagnosed by regressing (N – F) against ∆T.

376

Alternatively, in the AGCM framework, it is trivial to estimate the components of the

377

feedback parameter from the AMIPSST,Ice experiment since F = 0 by definition. In

378

some ways this is a reverse approach to the Forster and Taylor (2006) method, who

379

first assumed α to derive F. Here F is first determined in order to derive α.

380 381

Fig. 6 diagnoses a net feedback parameter of -α = -1.79 ± 0.35 Wm-2 K-1 from decadal

382

variations in the simulated radiation balance and ∆T in the AMIPSST,Ice experiment.

383

Andrews et al. (2012b) examined these relationships in coupled atmosphere-ocean

384

CMIP5 models on longer timescales using abrupt 4xCO2 simulations. Unfortunately

385

a long term climate sensitivity experiment has not been run using the exact model

386

physics used here, but a comparison of the response of HadGEM2-A and HadGEM2-

387

ES might still provide some insight because they share the same physical atmospheric

388

model. The net feedback parameter is substantially more positive (i.e. the climate

389

more sensitive) in the long term coupled atmosphere-ocean HadGEM2-ES simulation

16

390

(-α = -0.64 Wm-2 K-1). Comparing the individual feedback components (LW clear-

391

sky ~ -2 (-1.7) Wm-2 K-1, SW clear-sky ~ +0.6 (+0.7) Wm-2 K-1 and net CRE ~ -0.3

392

(+0.4) Wm-2 K-1, HadGEM2-ES in parenthesis) reveals that this mostly arises because

393

of a larger (more positive) CRE feedback in the long term coupled simulation.

394 395

It cannot be ruled out that the HadGEM2-A model setup is simply a ‘lower

396

sensitivity’ model, or that the coupled response has a large impact on the CRE

397

feedback in the coupled model, but another consideration is that recent decadal

398

climate change is not a good analogue for long term (multi-decadal to centennial)

399

climate change. Fig. 5b shows the decadal trend in surface warming from 1979-2008.

400

As the model is forced with observed SSTs this warming pattern closely resembles

401

those observed (at least over the ocean). There exist large parts of the Pacific where

402

cooling has occurred during this period, consistent with the interdecadal Pacific

403

oscillation (e.g. Meehl et al., 2013). In contrast, tropical warming patterns show no

404

cooling in long-term climate model simulations (e.g. Knutti and Sedlacek, 2013) from

405

which feedback parameters are usually defined. Hence it would not be surprising if

406

the atmospheric response, especially cloud feedback, was different. This highlights

407

the limitations of defining feedbacks relative to global-mean temperature change, as it

408

does not account for differences in warming patterns (Senior and Mitchell, 2000;

409

Armour et al., 2013). Similar points have been made of studies that extrapolate

410

radiative dampening rates diagnosed from inter-annual variability (the pattern of

411

which is dominated by the El Nino-Southern Oscillation) to long term externally

412

forced change, especially for cloud feedback (e.g. Dessler, 2010).

413

17

414

Still, linearity between N – F against ∆T is a good approximation on the timescales of

415

this study (Fig. 6) and would be an appropriate test for evaluating the sensitivity of

416

atmospheric models against similarly derived observed estimates using satellite

417

measurements of the global radiation balance (e.g. Forster and Gregory, 2006;

418

Murphy et al., 2009). To what extent linearity holds to longer timescales, giving the

419

relationship more predictive power, or the influence of the coupled response, should

420

be a matter of further investigation.

421 422

5. Mechanisms of Recent Historical Climate Change

423 424

The experimental design is now utilized to separate the components of climate

425

response due to adjustments (i.e. the component driven directly by the forcing agent)

426

and SST changes. Whilst many of the below results have been studied in greater

427

detail elsewhere (see below and Section 2.3), the purpose here is to provide insight

428

and highlight the utility of the experimental design for investigating such changes

429

under the improved conceptual framework/forcing-definition and a set of systematic

430

experiments.

431 432

5.1 Decadal Trends in Atmospheric Temperatures

433 434

Observed and modelled cooling trends in stratospheric temperatures over recent

435

decades are well established (e.g. Seidel et al., 2011), despite recent uncertainties in

436

the magnitude of observed cooling trends (Thompson et al., 2012). Trends in global

437

stratospheric temperatures are almost entirely driven by the radiative effect of changes

438

in stratospheric composition (i.e. stratospheric adjustments), predominantly from

18

439

increased WMGHGs (mostly CO2) and O3 depletion. Stratospheric temperatures

440

therefore provide an important fingerprint of anthropogenic climate change due to the

441

different ways in which forcing agents alter the stratospheric temperature profile (e.g.

442

Folland et al., 1998).

443 444

The experimental design and conceptual framework presented here provides a simple

445

way of diagnosing this and could provide a test for AGCM evaluation,

446

complementing evaluations of radiation schemes (e.g. Forster et al., 2011), and has

447

been used in a similar way for detection and attribution (Folland et al., 1998; Sexton

448

et al., 2001). The global 1979-2008 decadal temperature trend profiles for the various

449

experiments are shown in Fig. 7a. The AMIPSST,Ice,F experiment (blue line) shows

450

tropospheric warming which is amplified aloft (see below) and large stratospheric

451

cooling trends. The stratospheric cooling trends are absent when changes in

452

atmospheric composition are not included (AMIPSST,Ice thick black line), as expected

453

if the cooling trends are dominated by external changes in stratospheric composition

454

(i.e. adjustments). The stratospheric cooling trend declines from ~0.5 K dec-1 at

455

~20km in the all forcings run to ~0.4 K dec-1 when WMGHGs are fixed, and ~0.5 K

456

dec-1 to ~0.0 K dec-1 just above 30km. In other words, at ~30 km, changes in

457

WMGHGs are entirely responsible for the global stratospheric cooling trend of ~0.5 K

458

dec-1 in this model. O3 depletion contributes to the cooling trend at around ~20km,

459

declines to zero at ~30km, and then increases again higher up (Fig. 7a).

460 461

The zonal-mean temperature trend profile in the AMIPSST,Ice,F experiment is shown in

462

Fig. 7b. This trend is separated into a component driven through decadal increases in

463

SST (from the AMIPSST,Ice experiment) (Fig. 7c) and a direct forcing driven

19

464

component which is estimated as the residual (assuming linearity) between this and

465

the trend diagnosed from the standard AMIPSST,Ice,F experiment (Fig. 7d). It confirms

466

that the stratospheric temperature trends are largely driven by the radiative effect of

467

changes in atmospheric composition, independent of SST and tropospheric warming,

468

even in the zonal-mean. The forcing component is further separated into its

469

WMGHGs (which is mostly dominated by CO2) and O3 contributions (Fig. 7e and 7f

470

respectively). The component driven by multi-decadal SST and tropospheric

471

warming is shown in Fig. 7c and shows a cooling trend in tropical stratospheric

472

temperatures and a warming trend in high latitude stratospheric temperatures. This is

473

expected from a strengthened Brewer-Dobson circulation with climate change (e.g.

474

Butchart et al., 2006; McLandress and Shepherd, 2009), but in the global-mean these

475

SST driven stratospheric temperature trends cancel at all levels (Fig. 7a; Shine et al.,

476

2003b; Seidel et al., 2011; Thompson et al., 2012). Note that Folland et al. (1998) and

477

Sexton et al. (2001) showed significant stratospheric warming in response to

478

increased SST using an older generation model (HadAM2a), even in the global-mean,

479

but they attributed this to a model artefact.

480 481

Note that these model results clearly depend on the forcing datasets and model used,

482

and no evaluation against observed estimates have been made. Changes in

483

stratospheric water vapour could also have an impact on stratospheric temperature

484

trends, especially in the lower stratosphere (Shine et al., 2003; Seidel et al., 2011)

485

where some models exhibit large trends in water vapour over recent decades

486

(Gettelman et al., 2009), but is not examined here.

487

20

488

In the troposphere, large warming is seen at high latitudes and in the upper tropical

489

troposphere (Fig. 7b). The amplified warming aloft in the tropics is expected from

490

moist convective processes (e.g. Manabe and Wetherald, 1975), but comparisons with

491

observations can be controversial due to large uncertainties in trends (Thorne et al.,

492

2011). Fig. 7a and 7c shows that most of the tropospheric temperature trends are

493

captured by the AMIPSST,Ice experiment. Hence most tropospheric temperature trends

494

are determined solely by SST changes and associated climate feedbacks; the trends in

495

tropospheric temperature adjustments (Fig. 7d) are small compared to this term.

496 497

5.2 Surface Temperatures

498 499

Folland et al. (1998) (amongst others) showed that models forced only with time

500

varying SSTs and sea-ice underestimate land warming compared to models

501

additionally forced with time-varying forcing agents. This implies a significant role

502

for forcing agents in directly increasing land temperatures independent of SST

503

changes.

504 505

This result is examined in Fig. 8. When the model is forced with time-varying forcing

506

agents as well as observed SSTs and sea-ice the linear trend in global temperature

507

anomaly from 1979 to 2008 is 0.172 K dec-1 (Fig. 8a, blue line). This is close to the

508

observed estimate (0.17 K dec-1) reported by Morice et al. (2012) for the period 1979-

509

2010 using HadCRUT4 data. When the model is forced only with changes in SST

510

and sea-ice the trend drops to 0.135 K dec-1 (Fig. 8a, black line). Hence, assuming

511

linearity, ~80% of the recent decadal global temperature trend in this model is driven

512

by changes in SST and ~20% directly driven by adjustments to forcings. As the SSTs

21

513

are identically prescribed in these experiments the temperature adjustment arises

514

predominantly from over land (Fig. 8b); the global land warming trend of 0.304 K

515

dec-1 has contributions of ~60% due to SST changes and ~40% due directly to forcing

516

agents. Focusing over the northern hemisphere (NH) land (Fig. 8c), the warming

517

trend in AMIPSST,Ice (0.216 K dec-1) is again ~60% of the total trend in AMIPSST,Ice,F

518

(0.361 K dec-1), implying ~40% contribution directly from forcing agents. Winter NH

519

land temperatures trends are difficult to isolate due to internal atmospheric variability

520

(not shown), but trends in NH summer (June, July and August) land temperatures

521

again drop by ~40% when forcing agents are not considered, from 0.405 K dec-1 in

522

AMIPSST,Ice,F to 0.254 K dec-1 in AMIPSST,Ice (Fig. 8d). A direct contribution from

523

forcing agents of ~40% is slightly smaller than that reported by Folland et al. (1998),

524

who found it contributed as equally as SST changes to the global land temperature

525

change between 1950 and 1994. This difference could be a difference in model

526

response, but it could also be due to the different time periods considered (1950 to

527

1994 compared to 1979-2008).

528 529

Given that ∆T is not identical between AMIPSST,Ice,F and AMIPSST,Ice due to the direct

530

effects of forcing agents on land temperatures (Fig. 8) it raises questions about the

531

physical interpretation of ERF and its separation from climate feedbacks. The land

532

warming and consequent effects on the atmosphere (i.e. circulation, clouds etc.) are

533

included in the ERF calculation, but could be described as a climate feedback.

534

Hansen et al. (2005) and Shine et al. (2003a) discuss an analogous effect when

535

considering fixed-surface or fixed-SST forcing definitions. Gregory and Webb (2008)

536

define the separation of forcing and feedback according to timescale. In their

22

537

approach direct land warming acts on a similar timescale as atmospheric adjustments

538

and thus can be usefully considered part of the forcing as it is here.

539 540

5.3 Precipitation

541 542

The Earth’s atmospheric energy balance provides a useful constraint for

543

understanding global (e.g. Mitchell et al., 1987) and regional (Muller and O’Gorman,

544

2011) precipitation changes. For example, increased CO2 and black carbon directly

545

reduce atmospheric radiative cooling and this is largely balanced by changes in

546

condensational heating (precipitation), independent of any surface warming which

547

then tends to increase precipitation (e.g. Andrews et al., 2010; Ming et al., 2010,

548

amongst others). A few studies (e.g. Andrews, 2009; Bichet et al., 2011; Frieler et al.,

549

2011; Allan et al., 2013; Bony et al., 2013) have looked at the direct atmospheric

550

impact of forcing agents on precipitation in realistic transient scenarios, but most

551

studies use idealised experiments. It is still unclear whether precipitation adjustments

552

could be an important contributor to recent precipitation trends.

553 554

Fig. 9a shows the global timeseries of precipitation in the AMIPSST,Ice,F and

555

AMIPSST,Ice experiments. The increase in global precipitation when forcing agents are

556

allowed to vary (0.003 mm day-1 dec-1) is substantially muted compared to when only

557

SST and sea-ice changes are allowed to vary (0.009 mm day-1 dec-1), consistent with

558

the direct effect of forcing agents reducing precipitation.

559 560

A commonly defined metric is the ‘hydrological sensitivity’, defined as global-mean

561

precipitation change per global-mean surface temperature change (dP/dT). Models

23

562

suggest this to be ~ 2-3 % K-1 (e.g. Held and Soden, 2006; Lambert and Webb, 2008).

563

Fig. 11c and 11d shows the hydrological sensitivity (determined from anomalies in

564

global precipitation and surface temperature) in the AMIPSST,Ice,F and AMIPSST,Ice

565

experiment respectively. When only forced with increasing SST and variations in

566

sea-ice the hydrological sensitivity is ~ 2 % K-1 in this model, but this is muted to

567

only ~ 1 % K-1 when the direct effects of forcing agents on precipitation are

568

considered.

569 570

Over land (Fig. 9b) the trend in precipitation is more variable and negative. In

571

contrast to the global results, the direct influence of forcing agents on precipitation

572

makes this negative trend less negative (i.e. it enhances precipitation). This is

573

consistent with the direct effect of forcing agents causing land surface warming

574

(Section 5.2) which tends to enhance evaporation and precipitation, in contrast to

575

precipitation over the ocean where the atmospheric heating increases atmospheric

576

stability and suppresses convection, moistening the boundary layer and reducing

577

precipitation/evaporation (Cao et al., 2012).

578 579

6. Discussion

580 581

The experimental design covered the period 1979-2008 in order to exploit the already

582

well established AMIP experimental design that is routinely performed by modelling

583

centres. If information on the entire historical ERF and adjustments timeseries was

584

needed, then one could readily extend the design beyond the years covered, as per

585

Held et al. (2010) and Anderson et al. (2010; 2012).

586

24

587

Computing transient ERFs in future scenarios would be more problematic under the

588

AMIP design since it is not obvious what SST and sea-ice boundary conditions should

589

be used. One solution would be to use the model’s own SST and sea-ice fractions

590

from a coupled simulation. Alternatively, periodic boundary conditions could also be

591

used. For example, Rotstayn and Penner (2001), Hansen et al., (2002) and Ming and

592

Ramaswamy (2012) all use observed climatological boundary conditions to derive

593

ERFs. The CMIP5 design includes a suite of AGCM experiments that use periodic

594

pre-industrial SST and sea-ice climatologies representative of each models fully-

595

coupled pre-industrial control (the sstClim, sstClim4xCO2 and sstClimAerosol

596

experiments). Adding similar AGCM experiments, whether using observed or

597

modelled climatological SST and sea-ice boundary conditions, forced with transient

598

historical and future forcings would allow a simple diagnosis of each model’s

599

transient ERF and adjustments.

600 601

Using climatological, rather than time-evolving, SST and sea-ice boundary conditions

602

might help to reduce noise. For example Fig. 1a and 1b indicates there is significant

603

inter-annual variability across the 3 member ensemble which would limit analysing

604

year-to-year variations in ERF if one only had a single run (though large volcanic

605

forcings are clearly seen above the variability in all ensemble members, Fig 1b).

606

Without large ensembles the methodology is better defined for decadal variations in

607

ERF; the decadal trend in net ERF in Fig. 3b of all three ensemble members fall

608

within ±0.03 Wm-2 dec-1 of the ensemble-mean trend of 0.54 Wm-2 dec-1.

609 610

A conceptually important point is that the AGCM design with evolving SST boundary

611

conditions is calculating the ERF relative to an evolving base state. This may be

25

612

desirable if there is a dependence of forcing and climate state; it will give a forcing

613

closer to what the model actually felt at the time the forcing was applied. On the

614

other hand, it will mean forcing can no longer be assumed constant as the climate

615

evolves.

616 617

7. Summary

618 619

An atmospheric generation circulation model was forced with time-varying (1979-

620

2008) changes in radiatively active constituents, such as greenhouse gases and

621

atmospheric aerosols, and monthly observed sea-surface-temperature and sea-ice

622

boundary conditions. A simple experimental design is then proposed to diagnose the

623

global and regional effective radiative forcing for the period 1979-2008 relative to

624

1860. This is done by repeating the simulations with various forcing agents,

625

individually and in combination, fixed at 1860 levels.

626 627

Whilst earlier studies have to a limited extent done similar work, this is the first study

628

to show how a framework and systematic set of experiments can be used to (i)

629

diagnose the recent time-varying regional ERF for individual forcing mechanisms, (ii)

630

compute climate sensitivity parameters of an AGCM forced by observed SST

631

changes, and (iii) examine adjustments as mechanisms of recent historical climate

632

change, all set within our improved conceptual models/definitions of radiative

633

forcing.

634 635

If adopted by other modelling groups the simple experimental design used here would

636

allow us to diagnose recent historical radiative forcing in models, identify and address

26

637

potential uncertainties in modelled radiative forcing, compare components of radiative

638

feedbacks under a realistic transient scenarios, and examine the relationship between

639

radiative forcing and mechanisms of recent decadal climate change. Repeating these

640

calculations with a model’s internally generated SSTs as boundary conditions

641

(derived from an equivalent historical coupled atmosphere-ocean simulation) may

642

provide a useful approach for testing the potential impact of model biases in SST on

643

effective radiative forcing, feedbacks, climate sensitivity and recent decadal climate

644

trends.

645 646

Acknowledgements

647

I am grateful for useful discussions with Jonathan Gregory, William Ingram,

648

Mark Ringer, Mark Webb, Karl Taylor and Piers Forster. I thank two anonymous

649

reviewers for helping to improve the clarity of the manuscript.

650

supported by the Joint DECC/Defra Met Office Hadley Centre Climate Programme

651

(GA01101).

27

This work was

652

References

653 654

Allan, R.P, C. Liu, M. Zahn, D. A. Lavers, E. Koukouvagias and A. Bodas-Salcedo

655

(2013), Physically consistent responses of the global atmospheric hydrological cycle

656

in models and observations, Surv. Geophys., doi:10.1007/s10712-012-9213-z.

657 658

Anderson, B.T., J.R. Knight, M.A. Ringer, J. Yoon, A. Cherchi (2012), Testing for

659

the Possible Influence of Unknown Climate Forcings upon Global Temperature

660

Increases from 1950 to 2000. J. Climate, 25, 7163–7172.

661 662

Anderson, B.T., J.R. Knight, M.A. Ringer, C. Deser, A.S. Phillips, J.-H. Yoon, and A.

663

Cherchi (2010), Climate forcings and climate sensitivities diagnosed from

664

atmospheric global circulation models. Clim. Dyn., 35, 1461-1475,

665

doi:10.1007/s00382-010-0798-y.

666 667

Andrews, T., J.M. Gregory, P.M. Forster, and M.J. Webb (2012a), Cloud adjustment

668

and its role in CO2 radiative forcing and climate sensitivity: A review, Surv.

669

Geophys., 33, 619–635, doi:10.1007/s10712-011-9152-0.

670 671

Andrews, T., J. M. Gregory, M. J. Webb, and K. E. Taylor (2012b), Forcing,

672

feedbacks and climate sensitivity in CMIP5 coupled atmosphere–ocean climate

673

models, Geophys. Res. Lett., 39, L09712.

674

28

675

Andrews, T., P.M. Forster, O. Boucher, N. Bellouin and A. Jones (2010),

676

Precipitation, radiative forcing and global temperature change. Geophys. Res. Lett.,

677

27, L14701, doi:10.1029/2010GL043991.

678 679

Andrews, T. (2009), Forcing and response in simulated 20th and 21st century surface

680

energy and precipitation trends. J. Geophys. Res., 114, D17110,

681

doi:10.1029/2009JD011749.

682 683

Armour, Kyle C., Cecilia M. Bitz, Gerard H. Roe (2013), Time-varying climate

684

sensitivity from regional feedbacks. J. Climate, 26, 4518–4534,

685

doi:10.1175/JCLI-D-12-00544.1.

686 687

Bala, G., K. Caldeira and R. Nemani (2010), Fast versus slow response in climate

688

change: implications for the global hydrological cycle. Clim. Dyn., 35:423-434,

689

doi:10.1007/s00382-009-0583-y.

690 691

Bellouin, N., J. Rae, A. Jones, C. Johnson, J. Haywood, and O. Boucher (2011),

692

Aerosol forcing in the Climate Model Intercomparison Project (CMIP5) simulations

693

by HadGEM2-ES and the role of ammonium nitrate, J. Geophys. Res., 116, D20206,

694

doi:10.1029/2011JD016074.

695 696

Bichet, A., M. Wild, D. Folini, and C. Schär (2012), Causes for decadal variations of

697

wind speed over land: Sensitivity studies with a global climate model, Geophys. Res.

698

Lett., 39, L11701, doi:10.1029/2012GL051685.

699

29

700

Bichet, A., M. Wild, D. Folini and C. Schar (2011), Global precipitation response to

701

changing forcings since 1870. Atmos. Chem. Phys., 11, 9961-9970.

702 703

Bony, S., G. Bellon, D. Klocke, S. Sherwood, S. Fermepin and S. Denvil (2013),

704

Robust direct effect of carbon dioxide on tropical circulation and regional

705

precipitation. Nature Geoscience, doi:10.1038/ngeo1799.

706 707

Bony, S., M. Webb, C. Bretherton, S Klein, P. Siebesma, G. Tselioudis and M. Zhang

708

(2011), CFMIP: Towards a better evaluation and understanding of clouds and cloud

709

feedbacks in CMIP5 models. Clivar Exchanges, 56, 16, 2.

710 711

Bosilovich, M.G., S.D. Schubert and G.K. Walker (2005), Global changes of the

712

water cycle intensity. J. Climate, 18, 1591–1608, doi: /10.1175/JCLI3357..

713 714

Butchart, N., and Coauthors, (2006), Simulations of anthropogenic change in the

715

strength of the Brewer–Dobson circulation. Climate Dyn., 27, 727–741.

716

doi:10.1007/s00382-006-0162-4.

717 718

Caminade, C., and L. Terray (2006), Influence of increased greenhouse gases and

719

sulphate aerosols concentration upon diurnal temperature range over Africa at the end

720

of the 20th century. Geophys. Res. Lett., 33, L15703, doi:10.1029/2006GL026381.

721 722

Cao, L., G. Bala and K. Caldeira (2012), Climate response to changes in atmospheric

723

carbon dioxide and solar irradiance on the time scale of days to weeks. Environ. Res.

724

Lett., 7 034015 doi:10.1088/1748-9326/7/3/034015.

30

725 726

Deser, C., A.S. Phillips (2009), Atmospheric Circulation Trends, 1950–2000: The

727

Relative Roles of Sea Surface Temperature Forcing and Direct Atmospheric Radiative

728

Forcing. J. Climate, 22, 396–413.

729 730

Dessler, A. E. (2010), A determination of the cloud feedback from climate variations

731

over the past decade, Science, 330, 1523–1527, doi:10.1126/science.1192546.

732 733

Dong, B., J.M. Gregory and R.T. Sutton (2009), Understanding land-sea warming

734

contrast in response to increasing greenhouse gases. Part I: Transient adjustment. J.

735

Climate, 22, 3079-3097.

736 737

Folland, C.K., J. Shukla, J. Kinter and M.J. Rodwell (2002), The climate of the

738

twentieth century project. CLIVAR Exchanges 7(2):37-39.

739 740

Folland, C.K., D.M.H. Sexton, D.J. Karoly, C.E. Johnson, D.P. Rowell and D.E.

741

Parker (1999): Influences of anthropogenic and oceanic forcing on recent climate

742

change. Science, 25, 352-356.

743 744

Forster, P.M., T. Andrews, P. Good, J.M. Gregory, L.S. Jackson and M. Zelinka

745

(2013), Evaluating adjusted forcing and model spread for historical and future

746

scenarios in the CMIP5 generation of climate models. J. Geophys. Res., 118,

747

doi:10.1002/jgrd.50174.

748

31

749

Forster, P.M., et al. (2011), Evaluation of radiation scheme performance within

750

chemistry climate models, J. Geophys. Res., 116, D10302,

751

doi:10.1029/2010JD015361.

752 753

Forster, P.M., and K. Taylor (2006), Climate forcings and climate sensitivities

754

diagnosed from coupled climate model integrations, J. Clim., 6181–6194.

755 756

Forster, P.M., and J.M. Gregory (2006), The Climate Sensitivity and Its Components

757

Diagnosed from Earth Radiation Budget Data. J. Climate, 19, 39–52.

758 759

Forster, P.M., R.S. Freckleton and K.P. Shine (1997), On aspects of the concept of

760

radiative forcing. Clim. Dyn., 13, 547-560.

761 762

Frieler, K., M. Meinshausen, T. Schneider von Deimling, T. Andrews and P. Forster

763

(2011), Changes in global-mean precipitation in response to warming, greenhouse gas

764

forcing and black carbon. Geophys. Res. Lett., 38, L04702,

765

doi:10.1029/2010GL045953.

766 767

Gates, W.L., et al., (1999) An Overview of the Results of the Atmospheric Model

768

Intercomparison Project (AMIP I). Bull. Amer. Meteor. Soc., 80, 29–55.

769 770

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

771

Amer. Meteor. Soc., 73, 1962–1970.

772

32

773

Gregory, J. M., and P. Forster (2008), Transient climate response estimated from

774

radiative forcing and observed temperature change, J. Geophys. Res., 113, D23105.

775 776

Gregory, J.M. and M.J. Webb (2008), Tropospheric adjustment induces a cloud

777

component in CO2 forcing. J. Climate, 21, 58-71, doi :10.1175/2007JCLI1834.1.

778 779

Gregory, J.M., et al. (2004), A new method for diagnosing radiative forcing and

780

climate sensitivity. Geophys. Res. Lett., 31, L03205, doi:10.1029/2003GL018747.

781 782

Hansen, J., et al. (2005), Efficacy of climate forcings. J. Geophys. Res., 110, D18104,

783

doi:10.1029/2005JD005776.

784 785

Hansen, J., et al (2002), Climate forcings in Goddard Institute for Space Studies

786

SI2000 simulations, J. Geophys. Res., 107(D18), 4347, doi:10.1029/2001JD001143.

787 788

Hansen, J., M. Sato, and R. Ruedy (1997), Radiative forcing and climate response, J.

789

Geophys. Res., 102(D6), 6831–6864, doi:10.1029/96JD03436.

790 791

Held, I.M., et al. (2010), Probing the fast and slow components of global warming by

792

returning abruptly to preindustrial forcing. J. Climate, 23, 2418-2427.

793 794

Held, I.M., and B.J. Soden (2006), Robust Responses of the Hydrological Cycle to

795

Global Warming. J. Climate, 19, 5686–5699, doi:10.1175/JCLI3990.1.

796

33

797

Jiang, J. H., et al. (2012), Evaluation of cloud and water vapor simulations in CMIP5

798

climate models using NASA “A-Train” satellite observations, J. Geophys. Res., 117,

799

D14105, doi:10.1029/2011JD017237.

800 801

Jones, C.D., et al., (2011), The HadGEM2-ES implementation of the CMIP5

802

centennial simulations. Geosci. Model Dev ., 4, 543-570, doi:10.5194/gmd-4-543-

803

2011.

804 805

Keeling, C.D., Bacastow, R. B., Bainbridge, A. E., Ekdahl, C. A., Guenther, P. R.,

806

Waterman, L. S. and Chin, J. F. S. (1976), Atmospheric carbon dioxide variations at

807

Mauna Loa Observatory, Hawaii. Tellus, 28: 538–551.

808 809

Kiehl, J.T., (2007), Twentieth century climate model response and climate sensitivity.

810

Geophys. Res. Lett., 34, L22710, doi:10.1029/2007GL031383.

811 812

King, M.P., F. Kucharski and F. Molteni (2010), The roles of external forcings and

813

internal variabilities in the northern hemisphere atmospheric circulation change from

814

the 1960s to the 1990s. J. Climate, 23, 6200–6220.

815 816

Knutti, R., and J. Sedlacek (2013), Robustness and uncertainties in the new CMIP5

817

climate model projections. Nature Climate Change, 3, 369-373,

818

doi:10.1038/nclimate1716.

819

34

820

Kucharski, F., et al., (2009), The CLIVAR C20C project: skill of simulating Indian

821

monsoon rainfall on interannual to decadal timescales. Does GHG forcing play a role?

822

Clim. Dyn., 33, 615-627, doi:10.1007/s00382-008-0462-y.

823 824

Lambert, F.H., and M.J. Webb (2008), Dependence of global mean precipitation on

825

surface temperature. Geophys. Res. Lett., 35, L16706, doi:10.1029/2008GL034838.

826 827

Loeb, N.G., B.A. Wielicki, D.R. Doelling, G. Louis Smith, D.F. Keyes, S. Kato, N.

828

Manalo-Smith and T. Wong (2009), Toward Optimal Closure of the Earth's Top-of-

829

Atmosphere Radiation Budget. J. Climate, 22, 748–766.

830 831

Lohmann, U., et al., (2010), Total aerosol effect: radiative forcing or radiative flux

832

perturbation? Atmos. Chem. Phys. 10, 3235-3246.

833 834

Manabe, S. and R.T. Wetherald (1975), The Effects of Doubling the CO2

835

Concentration on the climate of a General Circulation Model. J. Atmos. Sci., 32, 3–15.

836 837

Martin, G.M., et al. (2011), The HadGEM2 family of Met Office Unified Model

838

Climate configurations. Geosci. Model Dev ., 4, 723-757, doi:10.5194/gmd-4-723-

839

2011.

840 841

McLandress, C. and T.G. Shepherd (2009), Simulated Anthropogenic Changes in the

842

Brewer–Dobson Circulation, Including Its Extension to High Latitudes. J. Climate,

843

22, 1516–1540.

844

35

845

Meehl, G A., A. Hu, J.M. Arblaster, J. Fasullo, and K.E.Trenberth (2013), Externally

846

forced and internally generated decadal climate variability associated with the

847

Interdecadal Pacific Oscillation, J. Climate, In press.

848 849

Ming, Y., and V. Ramaswamy (2012), Nonlocal component of radiative flux

850

perturbation, Geophys. Res. Lett., 39, L22706, doi:10.1029/2012GL054050.

851 852

Ming, Y., V. Ramaswamy, and G. Persad (2010), Two opposing effects of absorbing

853

aerosols on global-mean precipitation, Geophys. Res. Lett., 37, L13701,

854

doi:10.1029/2010GL042895.

855 856

Mitchell, J.F.B., C.A. Wilson and W.M. Cunnington (1987), On CO2 climate

857

sensitivity and model dependence of results. Q. J. R. Meteorol. Soc., 113, 293-322.

858 859

Morice, C. P., J. J. Kennedy, N. A. Rayner, and P. D. Jones (2012), Quantifying

860

uncertainties in global and regional temperature change using an ensemble of

861

observational estimates: The HadCRUT4 data set, J. Geophys. Res., 117, D08101,

862

doi:10.1029/2011JD017187.

863 864

Muller, C. J. and P.A. O'Gorman (2011), An energetic perspective on the regional

865

response of precipitation to climate change. Nature Climate Change, 1, 266-271.

866 867

Murphy, D.M., (2013), Little net clear-sky radiative forcing from recent regional

868

redistribution of aerosols. Nature Geoscience, 6, 258–262, doi:10.1038/NGEO1740.

869

36

870

Murphy, D. M., S. Solomon, R. W. Portmann, K. H. Rosenlof, P. M. Forster, and T.

871

Wong (2009), An observationally based energy balance for the Earth since 1950, J.

872

Geophys. Res., 114, D17107, doi:10.1029/2009JD012105.

873 874

Myhre, G., E.J. Highwood, K.P. Shine and F. Stordal (1998), New estimates of

875

radiative forcing due to well mixed greenhouse gases. Geophys. Res. Lett., 25, 14,

876

2715-2718.

877 878

Patricola, C.M., K.H. Cook (2011), Sub-Saharan African climate at the end of the

879

twenty-first century: forcing factors and climate change processes. Clim. Dyn.,

880

37:1165-1188, doi:10.1007/s00382-010-0907-y.

881 882

Rotstayn, L.D., and J. E. Penner (2001), Indirect Aerosol Forcing, Quasi Forcing, and

883

Climate Response. J. Climate, 14, 2960–2975.

884 885

Scaife, A., et al., (2009), The CLIVAR C20C project: selected twentieth century

886

climate events. Clim. Dyn., 33:603-614, doi:10.1007/s00382-008-0451-1.

887 888

Seidel, D.J., N.P. Gillett, J.R. Lanzante, K.P Shine and P.W. Thorne (2011),

889

Stratospheric temperature trends: our evolving understanding. WIREs Clim Change, 2:

890

592-616. doi: 10.1002/wcc.125.

891 892

Senior, C.A., and J.F.B Mitchell (2000), The time-dependence of climate sensitivity.

893

Geophys. Res. Lett., 21(17):2685-2688, doi:10.1029/2000GL011373.

894

37

895

Sexton, D.M.H, D.P. Rowell, C.K. Folland and D.J. Karoly (2001), Detection of

896

anthropogenic climate change using an atmospheric GCM. Clim. Dyn., 17:669-685.

897 898

Shindell, D. T., et. al, (2013), Radiative forcing in the ACCMIP historical and future

899

climate simulations. Atmos. Chem. Phys., 13, 2939-2974, doi:10.5194/acp-13-2939-

900

2013.

901 902

Shine, K.P., J. Cook, E.J. Highwood and M.J. Joshi (2003a), An alternative to

903

radiative forcing for estimating the relative importance of climate change

904

mechanisms. Geophys. Res. Lett., 30(20), 2047, doi:10.1029/2003GL018141.

905 906

Shine K.P., et al., (2003b), A comparison of model-predicted trends in stratospheric

907

temperatures. Q. J. R. Meteorl. Soc., 129:1565–1588, doi:10.1256/qJ02.186.

908 909

Shine, K.P., and P.M. Forster (1999), The effect of human activity on radiative

910

forcing of climate: a review of recent development. Global Planet. Change, 202-225.

911 912

Skinner, C.B., M. Ashfaq and N.S. Diffenbaugh (2012), Influence of Twenty-First-

913

Century Atmospheric and Sea Surface Temperature Forcing on West African Climate.

914

J. Climate, 25, 527–542.

915 916

Soden, B.J., I.M. Held, R. Colman, K.M. Shell, J.T. Kiehl and C.A. Shields (2008),

917

Quantifying climate feedbacks using radiative kernels, J. Climate, 21, 3504-3520,

918

doi:10.1175/2007JCLI2110.1.

919

38

920

Soden, B.J., A.J. Broccoli and R.S. Hemler (2004), On the use of cloud forcing to

921

estimate cloud feedback. J. Climate, 17, 19, 3661-3665.

922 923

Stephens, G.L., J. Li, M. Wild, C.A. Clayson, N. Loeb, S. Kato, T. L'Ecuyer, P.

924

Stackhouse, M. Lebsock and T. Andrews (2012), An update on Earth's energy balance

925

in light of the latest global observations. Nature Geoscience, 5, 691-696,

926

doi:10.1038/ngeo1580.

927 928

Streets, D. G., F. Yan, M. Chin, T. Diehl, N. Mahowald, M. Schultz, M. Wild, Y. Wu,

929

and C. Yu (2009), Anthropogenic and natural contributions to regional trends in

930

aerosol optical depth, 1980–2006, J. Geophys. Res., 114, D00D18,

931

doi:10.1029/2008JD011624.

932 933

Taylor, K.E., R.J. Stouffer and G.A. Meehl (2012), An overview of CMIP5 and the

934

experiment design. Bull. Amer. Meteor. Soc., doi:10.1175/BAMS-D-11-00094.1.

935 936

Thompson, D.W.J, D.J. Seidel, W.J. Randel, C. Zou, A.H. Butler, C. Mears, A. Osso,

937

C. Long, R. Lin (2012), The mystery of recent stratospheric temperature trends,

938

Nature, 491, 7426, 692.

939 940

Thorne, P. W., Lanzante, J. R., Peterson, T. C., Seidel, D. J. and Shine, K. P. (2011),

941

Tropospheric temperature trends: history of an ongoing controversy. WIREs Clim.

942

Change, 2: 66–88. doi: 10.1002/wcc.80.

39

943

Table 1: Summary of experiments performed. All experiments use observed

944

monthly (1979-2008) SST and sea-ice fraction boundary conditions and include a

945

3 member ensemble starting from different atmospheric states. The experiments

946

include a 4 month spin up period from September 1978 and are then run for 30

947

yrs to cover the period of analysis, 1979-2008. A tick ( ) indicates that a 1979-

948

2009 time-varying forcing species is included. A cross (x) indicates a constant

949

pre-industrial forcing level.

950 1979-2008 forcing relative to pre-industrial*

Experiment

CO2 CH4 N2O Halo O3 SO4 BC OC Bio Sol Vol AMIPSST,Ice,F























AMIPSST,Ice

x

x

x

x

x

x

x

x

x

x

x

No CO2

x





















No CH4



x



















No WMGHGs

x

x

x

x















No O3









x













No SO4











x











No BC













x









No Aerosol











x

x

x

x





No Natural



















x

x

951 952

*

953

Sulphate aerosol; BC = Black carbon aerosol; OC = Organic carbon aerosol; Sol =

954

Solar; Vol = Volcanic. Details of the forcing dataset and how they are implemented

955

in the model are given in Jones et al., (2011).

CO2 = Carbon dioxide; CH4 = Methane; Halo = Halocarbons; O3 = Ozone; SO4 =

40

956

Table 2: Effective radiative forcing of climate in 2005 (2003 – 2007 average)

957

relative to 1860 for various forcing agents and radiative components: longwave

958

(LW) and shortwave (SW) clear-sky (CS) and cloud radiative effect (CRE).

959 Forcing agent

Effective radiative forcing of climate between 1860 and 2005

(Wm-2)

NET

LWCS

SWCS

LWCRE

SWCRE

NETCS

NETCRE

WMGHGs

2.51

2.91

-0.14

-0.75

0.48

2.78

-0.27

CO2

1.50

1.59

-0.07

-0.27

0.25

1.52

-0.02

CH4

0.50

0.74

-0.13

-0.23

0.12

0.61

-0.11

N2O+Halocarbons*

0.50

0.58

0.06

-0.25

0.11

0.64

-0.14

O3

0.02

0.27

-0.37

0.09

0.02

-0.09

0.11

Aerosol

-1.38

0.14

-0.67

-0.03

-0.81

-0.53

-0.84

SO4

-1.33

0.17

-0.72

0.09

-0.87

-0.55

-0.79

Black Carbon

0.20

0.03

0.09

-0.12

0.20

0.12

0.08

BB+OC*

-0.25

-0.07

-0.04

0.00

-0.14

-0.11

-0.14

Anthropogenic*

1.29

3.19

-0.84

-0.64

-0.42

2.35

-1.06

Natural

0.18

-0.09

0.34

0.03

-0.10

0.25

-0.07

All

1.47

3.11

-0.50

-0.61

-0.52

2.60

-1.13

960 961

*

962

N2O+Halocarbons = WMGHGs – CO2 – CH4; BB+OC = Aerosol – SO4 – Black

963

Carbon; Anthropogenic = All – Natural.

These

forcing

agents

are

estimated

41

as

a

residual

assuming

linearity:

964

Table 3: 1979-2008 decadal trend in effective radiative forcing for various

965

forcing agents and radiative components: longwave (LW) and shortwave (SW)

966

clear-sky (CS) and cloud radiative effect (CRE). Trends are determined from

967

ordinary least square linear regression of the 1979-2008 global-annual-mean

968

timeseries.

969

volcanic forcings (see Fig. 1c), so these trends are additionally calculated after

970

removing years strongly influenced by volcanic forcing (1982-84 and 1991-93).

Natural and All forcings are strongly influenced by non-linear

971 Forcing agent (Wm-2 Dec-1)

Decadal trend in effective radiative forcing from 1979-2008 NET

LWCS

SWCS

LWCRE

SWCRE

NETCS

NETCRE

WMGHGs

0.43

0.43

0.01

-0.10

0.09

0.44

-0.01

CO2

0.27

0.24

0.03

-0.05

0.06

0.26

0.01

CH4

0.05

0.03

0.01

-0.02

0.03

0.04

0.00

N2O+Halocarbons*

0.12

0.17

-0.02

-0.03

0.01

0.14

-0.03

O3

-0.03

0.03

-0.08

0.02

0.01

-0.05

0.02

Aerosol

-0.02

-0.01

0.00

-0.02

0.00

-0.01

-0.01

SO4

0.02

-0.01

0.02

0.00

0.02

0.01

0.02

Black Carbon

0.05

0.01

0.02

-0.01

0.04

0.02

0.03

-0.10

0.00

-0.04

-0.01

-0.05

-0.04

-0.06

Anthropogenic*

0.35

0.42

-0.05

-0.09

0.06

0.37

-0.03

Natural

0.19

-0.14

0.42

0.04

-0.13

0.27

-0.08

Natural (excl. volc. yrs)

0.05

-0.03

0.09

0.01

-0.03

0.06

-0.01

All

0.54

0.27

0.37

-0.04

-0.07

0.65

-0.11

All (excl. volc. yrs)

0.40

0.40

0.04

-0.07

0.03

0.44

-0.04

BB+OC*

42

972

*These

forcing

agents

are

estimated

973

N2O+Halocarbons = WMGHGs – CO2 – CH4; BB+OC = Aerosol – SO4 – Black

974

Carbon; Anthropogenic = All – Natural.

975 976

43

as

a

residual

assuming

linearity:

977

Figure Captions

978 979

Figure 1: (a) Globally annually averaged net top-of-atmosphere (TOA) radiative flux

980

timeseries in the AMIPSST,Ice,F experiment (i.e. with 1979-2008 time-varying forcing

981

agents, as well as monthly observed SST and sea-ice boundary conditions) and the

982

AMIPSST,Ice experiment (i.e. same as above but with fixed pre-industrial forcing

983

levels). Shading equals the range across the 3 member ensemble. Red line indicates

984

CERES-EBAF satellite measured variations in net TOA radiative flux. (b) Time

985

series of the effective radiative forcing (ERF) relative to pre-indsutrial as diagnosed

986

from the difference in radiative flux between the AMIPSST,Ice,F and AMIPSST,Ice

987

experiments. (c) Components of the ERF (WMGHGs, Aerosol, O3 and Natural). (d)

988

Geographical distribution of the ERF averaged over the entire 1979-2008 time period.

989 990

Figure 2: Comparison of the effective radiative forcing of climate at 2005 (2003 –

991

2007 average) relative to pre-industrial (top) and the 1979-2008 decadal trends

992

(bottom). Further description and numbers are given in Tables 1 and 2.

993 994

Figure 3: (a) 1979-2008 average change in aerosol optical depth (AOD) at 0.55µm

995

relative to pre-industrial, diagnosed from an AMIP experiment with 1979-2008

996

aerosol emissions and one with pre-industrial aerosol emissions. (b) Change in net

997

TOA radiative flux, giving the aerosol ERF. (c) Clear-sky and (d) CRE contributions

998

to the aerosol ERF.

999 1000

Figure 4: 1979-2008 decadal trends in (a) aerosol AOD,

1001

radiative forcing (ERF), and (c) the clear-sky and (d) CRE contribution to the aerosol

44

(b) aerosol effective

1002

ERF trend. Quantities are calculated from a linear trend of the global-annual mean

1003

1979-2008 timeseries of the difference between an AMIP experiment with 1979-2008

1004

aerosol emissions and one with pre-industrial aerosol emissions.

1005 1006

Figure 5: (a) Relationship between the effective radiative forcing and global

1007

temperature anomaly (relative to the 1979-2008 average) in the AMIPSST,Ice,F

1008

experiment. The slope, excluding years influenced by volcanic forcing (red points),

1009

gives the climate resistance ρ = 1.7 Wm-2 K-1. All points are global-annual-mean data

1010

covering 1979-2008.

1011

determined from the 1979-2008 decadal trend in surface-air-temperature.

(b) The pattern of warming/cooling during this period,

1012 1013

Figure 6: Variations between (a) net TOA radiation, (b) longwave clear-sky radiation,

1014

(c) shortwave clear-sky radiation, (d) net (longwave+shortwave) cloud radiative effect

1015

and global temperature anomaly (relative to the 1979-2008 average) in the

1016

AMIPSST,Ice experiment. In this experiment there is no external forcing, so the

1017

variation of radiative flux with ∆T measures the feedback parameter. All fluxes are

1018

defined as positive down. The correlation coefficient, r, is shown. All points are

1019

global-annual-mean data from 1979-2008.

1020

uncertainties from the OLS regression.

Errors in the slope represent 1σ

1021 1022

Figure 7: (a) Decadal trend in the global atmospheric temperature profile from 1979-

1023

2008 for AMIP experiments with and without various time-varying forcing agents as

1024

indicated.

1025

AMIPSST,Ice,F experiment and its separation into (c) an SST driven component (from

1026

the AMIPSST,Ice experiment) and (d) a direct atmospheric response to the change in

(b) Decadal trend in the zonal-mean temperature profile for the

45

1027

forcing agent (estimated from (b) – (c)). The forcing driven component is then further

1028

separated into its two largest components: (e) WMGHGs and (f) O3 (estimated from

1029

decadal trends in AMIPSST,Ice,F minus an equivalent experiment with fixed WMGHGs

1030

and O3).

1031 1032

Figure 8: Annual-mean timeseries of (a) global, (b) land, (c) northern hemisphere land

1033

and (d) northern hemisphere land summer (June-August) surface-air-temperature

1034

anomalies (relative to the 1979-2008 average) for the AMIPSST,Ice,F (i.e. with time-

1035

varying forcing agents) and AMIPSST,Ice (i.e. with fixed forcing agents) experiments.

1036

Shading equals the range across the 3 member ensemble.

1037

indicated. Note that panel (a) has a different scale.

Decadal trends are

1038 1039

Figure 9: Annual-mean timeseries of (a) global and (b) land precipitation rate

1040

anomalies (relative to the 1979-2008 average) for the AMIPSST,Ice,F and AMIPSST,Ice

1041

experiments. Shading equals the range across the 3 member ensemble. Decadal

1042

trends are indicated. Global hydrological sensitivities (dP/dT) diagnosed from (c)

1043

AMIPSST,Ice,F and (d) AMIPSST,Ice. Points are global-annual-means for the period

1044

1979-2008. Note that (a) and (b) have different scales.

46

1045 1046

Figure 1: (a) Globally annually averaged net top-of-atmosphere (TOA) radiative

1047

flux timeseries in the AMIPSST,Ice,F experiment (i.e. with 1979-2008 time-varying

1048

forcing agents, as well as monthly observed SST and sea-ice boundary

1049

conditions) and the AMIPSST,Ice experiment (i.e. same as above but with fixed

1050

pre-industrial forcing levels). Shading equals the range across the 3 member

1051

ensemble. Red line indicates CERES-EBAF satellite measured variations in net

1052

TOA radiative flux. (b) Time series of the effective radiative forcing (ERF)

1053

relative to pre-indsutrial as diagnosed from the difference in radiative flux

1054

between the AMIPSST,Ice,F and AMIPSST,Ice experiments. (c) Components of the

1055

ERF (WMGHGs, Aerosol, O3 and Natural). (d) Geographical distribution of the

1056

ERF averaged over the entire 1979-2008 time period.

47

1057 1058

Figure 2: Comparison of the effective radiative forcing of climate at 2005 (2003 –

1059

2007 average) relative to pre-industrial (top) and the 1979-2008 decadal trends

1060

(bottom). Further description and numbers are given in Tables 1 and 2.

48

1061 1062

Figure 3: (a) 1979-2008 average change in aerosol optical depth (AOD) at 0.55µm

1063

relative to pre-industrial, diagnosed from an AMIP experiment with 1979-2008

1064

aerosol emissions and one with pre-industrial aerosol emissions. (b) Change in

1065

net TOA radiative flux, giving the aerosol ERF. (c) Clear-sky and (d) CRE

1066

contributions to the aerosol ERF.

49

1067 1068

Figure 4: 1979-2008 decadal trends in (a) aerosol AOD, (b) aerosol effective

1069

radiative forcing (ERF), and (c) the clear-sky and (d) CRE contribution to the

1070

aerosol ERF trend. Quantities are calculated from a linear trend of the global-

1071

annual mean 1979-2008 timeseries of the difference between an AMIP

1072

experiment with 1979-2008 aerosol emissions and one with pre-industrial aerosol

1073

emissions.

50

1074 1075

Figure 5: (a) Relationship between the effective radiative forcing and global

1076

temperature anomaly (relative to the 1979-2008 average) in the AMIPSST,Ice,F

1077

experiment.

1078

points), gives the climate resistance ρ = 1.7 Wm-2 K-1. All points are global-

1079

annual-mean data covering 1979-2008.

1080

during this period, determined from the 1979-2008 decadal trend in surface-air-

1081

temperature.

The slope, excluding years influenced by volcanic forcing (red

51

(b) The pattern of warming/cooling

1082 1083

Figure 6: Variations between (a) net TOA radiation, (b) longwave clear-sky

1084

radiation, (c) shortwave clear-sky radiation, (d) net (longwave+shortwave) cloud

1085

radiative effect and global temperature anomaly (relative to the 1979-2008

1086

average) in the AMIPSST,Ice experiment. In this experiment there is no external

1087

forcing, so the variation of radiative flux with ∆T measures the feedback

1088

parameter. All fluxes are defined as positive down. The correlation coefficient,

1089

r, is shown. All points are global-annual-mean data from 1979-2008. Errors in

1090

the slope represent 1σ uncertainties from the OLS regression.

52

1091 1092

Figure 7: (a) Decadal trend in the global atmospheric temperature profile from

1093

1979-2008 for AMIP experiments with and without various time-varying forcing

1094

agents as indicated. (b) Decadal trend in the zonal-mean temperature profile for

1095

the AMIPSST,Ice,F experiment and its separation into (c) an SST driven component

1096

(from the AMIPSST,Ice experiment) and (d) a direct atmospheric response to the

1097

change in forcing agent (estimated from (b) – (c)).

1098

component is then further separated into its two largest components: (e)

1099

WMGHGs and (f) O3 (estimated from decadal trends in AMIPSST,Ice,F minus an

1100

equivalent experiment with fixed WMGHGs and O3).

53

The forcing driven

1101 1102

Figure 8: Annual-mean timeseries of (a) global, (b) land, (c) northern

1103

hemisphere land and (d) northern hemisphere land summer (June-August)

1104

surface-air-temperature anomalies (relative to the 1979-2008 average) for the

1105

AMIPSST,Ice,F (i.e. with time-varying forcing agents) and AMIPSST,Ice (i.e. with

1106

fixed forcing agents) experiments.

1107

member ensemble. Decadal trends are indicated. Note that panel (a) has a

1108

different scale.

Shading equals the range across the 3

54

1109 1110

Figure 9: Annual-mean timeseries of (a) global and (b) land precipitation rate

1111

anomalies (relative to the 1979-2008 average) for the AMIPSST,Ice,F and

1112

AMIPSST,Ice experiments.

1113

ensemble.

1114

(dP/dT) diagnosed from (c) AMIPSST,Ice,F and (d) AMIPSST,Ice. Points are global-

1115

annual-means for the period 1979-2008. Note that (a) and (b) have different

1116

scales.

Shading equals the range across the 3 member

Decadal trends are indicated.

55

Global hydrological sensitivities

1 Using an AGCM to diagnose historical effective radiative forcing and ...

May 24, 2013 - It is currently most easily achieved using a global energy budget approach. 63 ... and solar constant representative of the period being simulated (Gates, 1992). The ...... 2008) changes in radiatively active constituents, such as ...

3MB Sizes 2 Downloads 132 Views

Recommend Documents

Optical Properties and Radiative Forcing of Southern ...
Jul 7, 2006 - for wavelengths spanning the solar spectrum. Using the retrieved aerosol optical properties as input to the radiative transfer model, we estimate the radiative forcing of southern African biomass burning aerosol. The diurnally-averaged

Global radiative forcing from contrail cirrus.pdf
Global radiative forcing from contrail cirrus.pdf. Global radiative forcing from contrail cirrus.pdf. Open. Extract. Open with. Sign In. Main menu. Displaying Global ...

The Tropospheric Jet Response to Prescribed Zonal Forcing in an ...
followed by similar-signed anomalies in the tropo- spheric annular mode that ..... The dashed–dotted line is the tropopause level in the control simulation. 2258.

Forcing Husband To Get Separated.pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Main menu.

Forcing Husband To Get Separated.pdf
Digitally signed by. NARENDRA PRASAD. Date: 2016.10.06. 16:23:11 IST. Reason: Signature Not Verified. WWW.LIVELAW.IN. Page 1 of 14 ...

Breakpoint Forcing Revisited: Phase Between Forcing ...
shown to depend on beach geometry and wave group parameters. The breakpoint-forced incoming long wave lags behind the wave group, by a phase smaller than p/2. The phase lag decreases as the beach slope decreases and the group frequency increases, app

an approach to lossy image compression using 1 ... - Semantic Scholar
In this paper, an approach to lossy image compression using 1-D wavelet transforms is proposed. The analyzed image is divided in little sub- images and each one is decomposed in vectors following a fractal Hilbert curve. A Wavelet Transform is thus a

an approach to lossy image compression using 1 ... - Semantic Scholar
images are composed by 256 grayscale levels (8 bits- per-pixel resolution), so an analysis for color images can be implemented using this method for each of ...

INCREASING 1 2 AND NAMBA-STYLE FORCING §1 ...
forcing in the model L[U] with one measurable cardinal introducing a mouse which iterates to any length given ..... MIAMI UNIVERSITY. UNIVERSITY OF FLORIDA.

Online PDF How to Diagnose and Fix Everything ...
... Second Edition, All Ebook How to Diagnose and Fix Everything Electronic, .... panel TVs, laptops, headsets, and mobile devices are also included in this ...