57
1 A Hybrid Dynamical-Statistical Downscaling 1 Technique, Part I: Development and Validation 2 of the Technique 3 4 5 Daniel B. Walton 1 6 Department of Atmospheric and Oceanic Sciences, 7 University of California, Los Angeles 8 9 Fengpeng Sun 10 Department of Atmospheric and Oceanic Sciences, 11 University of California, Los Angeles 12 13 Alex Hall 14 Department of Atmospheric and Oceanic Sciences, 15 University of California, Los Angeles 16 17 Scott Capps 18 Department of Atmospheric and Oceanic Sciences 19 University of California, Los Angeles 20 21 1 Corresponding author address: Daniel B Walton, 7229 Math Sciences Building, 405 Hilgard Ave, Los Angeles, CA 90095 E-mail: [email protected]

A Hybrid Dynamical-Statistical Downscaling … A Hybrid Dynamical-Statistical Downscaling 2 Technique, ... 88 Coordinated Regional Downscaling Experiment (CORDEX; ... In Part II, 141

  • Upload
    vuphuc

  • View
    225

  • Download
    0

Embed Size (px)

Citation preview

1

A Hybrid Dynamical-Statistical Downscaling 1

Technique, Part I: Development and Validation 2

of the Technique 3

4

5 Daniel B. Walton1 6

Department of Atmospheric and Oceanic Sciences, 7 University of California, Los Angeles 8

9 Fengpeng Sun 10

Department of Atmospheric and Oceanic Sciences, 11 University of California, Los Angeles 12

13 Alex Hall 14

Department of Atmospheric and Oceanic Sciences, 15 University of California, Los Angeles 16

17 Scott Capps 18

Department of Atmospheric and Oceanic Sciences 19 University of California, Los Angeles 20

21

1 Corresponding author address: Daniel B Walton, 7229 Math Sciences Building, 405 Hilgard Ave, Los Angeles, CA 90095 E-mail: [email protected]

2

21 Abstract 22

23 In Part I of this study, the mid-21st-century surface air temperature increase in the entire 24

CMIP5 ensemble is downscaled to very high resolution (2km) over the Los Angeles 25

region, using a new hybrid dynamical-statistical technique. This technique combines the 26

ability of dynamical downscaling to capture fine-scale dynamics with the computational 27

savings of a statistical model to downscale multiple GCMs. First, dynamical downscaling 28

is applied to five GCMs. Guided by an understanding of the underlying local dynamics, a 29

simple statistical model is built relating the GCM input and the dynamically downscaled 30

output. This statistical model is used to approximate the warming patterns of the 31

remaining GCMs, as if they had been dynamically downscaled. The full 32-member 32

ensemble allows for robust estimates of the most likely warming and uncertainty due to 33

inter-model differences. The warming averaged over the region has an ensemble mean of 34

2.3 ºC, with a 95% confidence interval ranging from 1.0 ºC to 3.6 ºC. Inland and high 35

elevation areas warm more than coastal areas year-round, and by as much as 60% in the 36

summer months. An assessment of the value added by the hybrid method shows that it 37

outperforms linear interpolation in approximating the dynamical warming patterns. These 38

improvements are attributed to the statistical model’s ability to capture the spatial 39

variations within the region more accurately. Additionally, this hybrid technique 40

incorporates an understanding of the physical mechanisms shaping the region’s warming 41

patterns, enhancing the credibility of the final results. 42

43

3

43

1. Introduction 44

To make informed adaptation and mitigation decisions, policymakers and other 45

stakeholders need future climate projections at the regional scale that provide robust 46

information about most likely outcomes and uncertainty estimates (Mearns et al. 2001, 47

Leung et al. 2003, Schiermeier 2010, Kerr 2011). The main tools available for such 48

projections are ensembles of global climate models (GCMs). However, GCMs have grid 49

box scales of 1° to 2.5° (~ 100 – 200 km), often too coarse to resolve important 50

topographical features and mesoscale processes that govern local climate (Giorgi and 51

Mearns 1991, Leung et al. 2003, Caldwell et al. 2009, Arritt and Rummukainen 2011). 52

The inability of GCMs to provide robust predictions at scales small enough for 53

stakeholder purposes has motivated numerous efforts to regionalize GCM climate change 54

signals through a variety of downscaling methods (e.g. Giorgi et al. 1994, Snyder et al. 55

2002, Timbal et al. 2003, Hayhoe et al. 2004, Leung et al. 2004, Tebaldi et al. 2005, 56

Duffy et al. 2006, Cabré et al. 2010, Salathé et al. 2010, Pierce et al. 2013). The aim of 57

this study is to develop downscaling techniques to recover the full complement of 58

warming signals in the greater Los Angeles region associated with the multi-model 59

ensemble from the World Climate Research Programme's Fifth Coupled Model 60

Intercomparison Project (CMIP5; Taylor et al. 2012; Table 1). 61

Regional downscaling attempts have been met with significant criticism (e.g., 62

Schiermeier 2010, Kerr 2011, 2013). One major critique is that the downscaled output is 63

constrained by the limitations of the GCM input. By itself, any single GCM may give a 64

4

misleading picture of the true state of knowledge about climate change, including in the 65

region of interest. Results from downscaling this single GCM will likewise be 66

misleading. Furthermore, the high resolution and realistic appearance of the downscaled 67

results may give a false impression of accuracy. This perception of accuracy at the 68

regional scale is especially problematic if a very small number of GCMs are downscaled, 69

since the uncertainty is dramatically undersampled. In this case, the downscaled output 70

may not reflect the most likely climate outcomes in the region, and it certainly does not 71

provide information about how the uncertainty associated with the GCM ensemble 72

manifests itself at the regional scale. Typically, previous studies have downscaled only 73

two global models (e.g., Hayhoe et al. 2004, Duffy et al. 2006, Cayan et al. 2008, Salathé 74

et al. 2010). This is too small an ensemble to obtain meaningful statistics about the most 75

likely (ensemble-mean) warming and uncertainty (inter-model spread). Instead, 76

information from a larger ensemble is preferred (Georgi and Mearns 2002, Kharin and 77

Zwiers 2002). The CMIP3 and CMIP5 ensembles (Meehl et al. 2007; Taylor et al. 2012), 78

with a few dozen ensemble members, are usually seen as large enough to compute a 79

meaningful ensemble-mean and span the climate change uncertainty space. 80

While downscaling of a large ensemble is desirable to compute most likely 81

outcomes and fully characterize uncertainty, this can be impractical because of the high 82

computational cost. Dynamical downscaling, in particular, is an expensive technique, and 83

most studies that perform it have only applied it to a few global models. For example, 84

Duffy et al. (2006) downscaled PCM and HadCM2 over the western United States, and 85

Pierce et al. (2013) downscaled GFDL CM2.1 and NCAR CCSM3 over California. There 86

are other examples of dynamical downscaling of multiple GCMs, such as the 87

5

Coordinated Regional Downscaling Experiment (CORDEX; Giorgi et al. 2009), but these 88

are very large undertakings that require coordination of multiple research groups. 89

Furthermore, they tend to span large geographic areas at lower resolutions (roughly 50 90

km) than needed for the region of interest here. Areas of intense topography and complex 91

coastlines typically need a model resolution finer than 10–15 km (Mass et al. 2002). The 92

greater Los Angeles region contains minor mountain complexes, such as the Santa 93

Monica Mountains, that have a significant role in shaping local climate gradients. These 94

mountain complexes have spatial scales of just a few kilometers, so even higher 95

resolution, with correspondingly higher computational costs, would be needed here. 96

Thus, for the purposes of this study, dynamical downscaling alone is an impractical 97

answer to the need for multi-model downscaling. 98

Due to its much lower computational cost, statistical downscaling is almost 99

always used for multi-model downscaling (e.g., Giorgi et al. 2001, Tebaldi et al. 2005, 100

Pierce et al. 2013). Unfortunately, statistical methods may not be able to capture 101

important fine-scale changes in climate shaped by topography and mesoscale dynamics. 102

Dynamical downscaling can capture such effects, provided the regional model resolution 103

is high enough (e.g. Caldwell et al. 2009, Salathé et al. 2010, Arritt and Rummukainen 104

2011, Pierce et al. 2013). Pierce et al. (2013) found that when a pair of global model was 105

dynamically downscaled, the average difference in the annual warming between the 106

Southern California mountains and coast was twice that of two common statistical 107

downscaling techniques. This suggests that statistical downscaling alone may be 108

insufficient in order to capture sharp gradients in temperature change in our region of 109

interest. 110

6

Here we provide a hybrid downscaling technique that allows us to fully sample 111

the GCM ensemble with the physical credibility of dynamical downscaling but without 112

the heavy computational burden of dynamically downscaling every GCM. In this 113

technique, dynamical downscaling is first performed on five GCMs. Then, the results 114

from dynamical downscaling are used to identify the common fine-scale warming 115

patterns and how they relate to the major GCM-scale warming features. Based on these 116

relationships, a simple statistical model is built to mimic the warming patterns produced 117

by the dynamical model. In the statistical model, the common fine-scale patterns are 118

dialed up or down to reflect the regional-scale warming found in the particular GCM 119

being downscaled. While scaling of regional climate change patterns has been around 120

since Mitchell et al. (1990) and Santer et al. (1990), the scaling has primarily been 121

relative to the global-mean warming and only within a single GCM (e.g. Cabré et al. 122

2010). The statistical model described here is more versatile because (1) it works for any 123

GCM, not just those dynamically downscaled; (2) the downscaled warming is dependent 124

on the GCM’s regional mean warming characteristics, not the global mean warming; and 125

(3) this dependence is allowed multiple degrees of freedom, based on the physical 126

processes at play in this particular region. 127

The construction of a statistical model that mimics dynamical model behavior 128

forces us to understand the physical mechanisms underpinning the regional patterns of 129

change, adding an additional layer of credibility to the results. This addresses another 130

concern about regional downscaling, namely that it is difficult to determine if the regional 131

climate change patterns are credible even if they appear realistic and visually appealing, 132

because the dynamics underpinning them are unclear, undiagnosed, or unknown. 133

7

After the statistical model undergoes a rigorous cross-validation procedure and 134

assessment of value added, it is applied to generate the warming patterns for the 135

remaining GCMs in the CMIP5 ensemble. These statistically generated warming patterns 136

represent our best estimate of what the warming would be if dynamical downscaling had 137

been performed on these remaining GCMs. The efficiency of the hybrid technique allows 138

us to downscale multiple emission scenarios and multiple time periods. In Part I, we 139

downscale 32 GCMs for the mid-century period (2041–2060) under RCP8.5. In Part II, 140

we expand this analysis by downscaling the full ensemble for RCP8.5 end-of-century 141

(2081–2100) and RCP2.6 mid-century and end-of-century. 142

2. Dynamical Downscaling 143

a. Model Configuration 144

Dynamical downscaling was performed using the Advanced Research Weather Research 145

and Forecasting Model version 3.2 (WRF; Skamarock et al. 2008). WRF has been 146

successfully applied to the California region in previous work (e.g. Caldwell 2009, Pierce 147

et al. 2013). For this study, we optimized it for the California region with sensitivity 148

experiments using various parameterizations, paying particular attention to the model’s 149

ability to simulate low cloud in the marine boundary layer off the California coast. The 150

following parameterization choices were made: Kain-Fritsch (new Eta) cumulus scheme 151

(Kain 2004); Yonsei University boundary layer scheme (Hong et al. 2006); Purdue Lin 152

microphysics scheme (Lin et al. 1983); Rapid Radiative Transfer Model longwave 153

radiation (Mlawer et al. 1997); Dudhia shortwave radiation schemes (Dudhia 1989). The 154

Noah land surface model (Chen and Dudhia 2001) was used to simulate land surface 155

8

processes including vegetation, soil, snowpack and exchange of energy, momentum and 156

moisture between the land and atmosphere. 157

The three nested domains for the simulations are shown in Fig. 1. The outermost 158

domain covers the entire state of California and the adjacent ocean at a horizontal 159

resolution of 18 km, the middle domain covers roughly the southern half of the state at a 160

horizontal resolution of 6 km, and the innermost domain encompasses Los Angeles 161

county and surrounding regions at a horizontal resolution of 2 km. In each domain, all 162

variables in grid cells closer than five cells from the lateral boundary in the horizontal 163

were relaxed toward the corresponding values at the lateral boundaries. This procedure 164

ensures smooth transitions from one domain to another. Each domain has 43 sigma-levels 165

in the vertical. To provide a better representation of surface and boundary layer 166

processes, the model’s vertical resolution is enhanced near the surface, with 30 sigma-167

levels below 3 km. 168

b. Baseline Simulation and Validation 169

Using this model configuration, we performed a baseline simulation whose purpose is 170

two-fold: (1) to validate the model’s ability to simulate regional climate, and (2) to 171

provide a baseline climate state against which a future climate simulation could be 172

compared, to quantify climate change. This simulation is a dynamical downscaling of the 173

National Centers for Environmental Prediction North America Regional Reanalysis 174

(NARR; Mesinger et al. 2006) over the period September 1981 to August 2001. This 175

dataset has 32-km resolution and provides lateral boundary conditions at the outer 176

boundaries of the outermost domain (Fig. 1). It also provides surface boundary conditions 177

over the ocean (i.e., sea surface temperature) in each of the three domains. The simulation 178

9

is designed to reconstruct the regional weather and climate variations that occurred in the 179

innermost domain during this time period, at 2-km resolution. The model was 180

reinitialized each year in August, and run from September to August. Because each year 181

was initialized separately, the time period could be divided into one-year runs performed 182

in parallel. 183

The regional model’s ability to reproduce climate variations during the baseline 184

period was assessed by comparing the output from the baseline climate simulation to the 185

available observational measurements from a network of 24 weather stations and buoys. 186

These quality-controlled, hourly, near-surface meteorological observations were obtained 187

from the National Climatic Data Center (NCDC; http://www.ncdc.noaa.gov/). The point 188

measurements are located in a variety of elevations and distances from the coast, and are 189

numerous enough to provide a sampling of the range of temperatures seen across the 190

region (Fig. 1). However, both the length and completeness of observational temperature 191

records vary by location. Most locations have reasonably complete records after 1995, so 192

validation is performed over the 1995–2001 period. 193

First, we check the realism of the spatial patterns seen in surface air temperature 194

climatology. Spatial patterns simulated by the model are highly consistent with 195

observations, as indicated by high correlations between observed and simulated 196

temperatures within each season (Fig. 2a). This confirms that for each season, the model 197

simulates spatial variations in climatological temperature reasonably well. The spatial 198

pattern is particularly well-represented in summer and winter (r > 0.9 in both seasons), 199

although the model exhibits a slight cold bias in the summer. During the transition 200

seasons, the model and observed spatial patterns are still in broad agreement, with 201

10

correlations greater than 0.7. The model’s ability to simulate temporal variability on 202

monthly timescales and longer is also assessed. At each of the 24 locations, the 203

correlation was computed between the observed and modeled time series of monthly-204

mean temperature anomalies, after first removing a composite seasonal cycle (Fig. 2b). 205

Temporal variability is very well-simulated by the model, with high correlations at all 206

locations. 207

Fig. 2 demonstrates that the model gives approximately the right spatial and 208

temporal variations in surface air temperature at specific point locations where 209

trustworthy observational data are available. This gives a high degree of confidence that 210

the model is also producing the correct temperature variations in the rest of the region, 211

where observations are absent. And most importantly for this study, it gives confidence 212

that when it comes to surface air temperature, the model provides a realistic downscaling 213

of the regional pattern implicit in the coarser resolution forcing data set. Thus, the 214

dynamically downscaled climate change patterns presented here are very likely a true 215

reflection of how the atmosphere’s dynamics would distribute the warming across the 216

region if climate change signals seen in the global models occurred in the real world. 217

218

c. Future Simulations 219

With the same model configuration as in the baseline simulation, we performed a second 220

set of dynamical downscaling experiments designed to simulate the regional climate state 221

corresponding to the mid–21st century. We applied the pseudo-global warming method 222

(see Rasmussen et al. 2011 and references therein; also Sato et al. 2007, Kawase et al. 223

2009) to five global climate models in the CMIP5 ensemble corresponding to this time 224

period and the RCP8.5 emissions scenario (See Table 1). To simulate the future period, 225

11

we started by calculating the difference between future and baseline monthly 226

climatologies (2041–2060 minus 1981–2000) for each GCM. These differences are the 227

GCM climate change signals of interest. All model variables are included in the 228

calculation of the climate change signal (i.e., 3-dimensional atmospheric variables such 229

as temperature, relative humidity, zonal and meridional winds, and geopotential height 230

and 2-dimensional surface variables such as temperature, relative humidity, winds and 231

pressure). To produce the boundary conditions for the future period, we perturbed NARR 232

data corresponding to the baseline period (September 1981–August 2001) by adding the 233

change in monthly climatology. The resulting simulation can then be compared directly 234

with the baseline regional simulation to assess the effect of the GCM climate change 235

signals when they are included in the downscaling. Because we downscaled the mean 236

climate change signal in each GCM rather than the raw GCM data, we did not downscale 237

changes in GCM variability. Thus, any future changes in variability in the regional 238

simulations are solely the result of WRF’s dynamical response. In addition to imposing a 239

mean climate change perturbation at the lateral boundaries, CO2 concentrations were also 240

increased in WRF to match CO2-equivalent radiative forcing in the RCP8.5 scenario. 241

We first downscaled CCSM4 for a 20-year period and then performed sensitivity 242

testing to see if it was necessary to downscale such a long period to recover the regional 243

temperature change signal. (Using a shorter period when downscaling the other GCMs 244

conserves scarce computational resources.) Because we perturbed each year in the future 245

period with the same monthly-varying change signal from CCSM4, we expected the 246

warming patterns for each year to be relatively similar. In fact, the warming patterns were 247

nearly identical each year: We could have dynamically downscaled only three years and 248

12

recovered an average warming signal within 0.1 ºC of the 20-year value. Therefore, the 249

remaining four GCMs were only downscaled for three years. For each of these GCMs, 250

the boundary conditions for the future run were created by adding the mean climate 251

change signal (2041–2060 minus 1981–2000) from the GCM to the three-year period of 252

NARR corresponding to September 1998–August 2001. 253

d. Warming Patterns 254

In this section, we examine monthly-mean warming patterns (future minus baseline) 255

simulated from dynamical downscaling. Fig. 3 shows these warming patterns averaged 256

over the five dynamically downscaled GCMs. There are two prominent features that can 257

be understood through underlying physical processes. First, the warming is greater over 258

land than ocean. This is true for all months, but the effect is particularly evident in the 259

late spring, summer, and early fall. Differences between warming over the ocean and land 260

surfaces have been well-documented in GCMs (Manabe et al. 1991; Cubasch et al., 2001; 261

Braganza et al., 2003, 2004; Sutton et al. 2007; Lambert and Chiang 2007; Joshi et al. 262

2008; Dong et al. 2009; Fasullo et al. 2010) and the observational record (Sutton et al. 263

2007, Lambert and Chiang 2007, Drost et al. 2011). Greater warming over land is evident 264

on the continental scale in both transient and equilibrium climate change experiments due 265

to greater heat capacity and availability of moisture for evaporative heat loss over the 266

ocean compared to land (Manabe et al. 1991). Moisture availability is particularly low in 267

arid and semi-arid regions, including a large swath of western North America adjacent to 268

the greater Los Angeles region. 269

Land-sea contrast in the warming is present on large scales in each global model’s 270

climate change signal, but how is this contrast expressed on the regional scale? Local 271

13

topography and the circulation simulated by WRF govern which areas have warming that 272

is more ocean-like or land-like. The land-sea breeze brings marine air and its 273

characteristics to the coastal zone on a daily basis (Hughes et al. 2007) which suppresses 274

warming there, keeping it at or near ocean levels. This suppression is limited to the 275

coastal zone because marine air masses cannot easily penetrate the surrounding mountain 276

complexes. Meanwhile, the inland areas separated from the coast by a mountain complex 277

are not exposed to marine air and have similar warming as interior land areas in the 278

global models. 279

The second prominent feature is the enhanced warming at high elevations, which 280

can be seen by comparing the warming to the domain topography shown in Fig. 1. 281

During winter and spring months, snow-albedo feedback occurs in mountainous areas, a 282

feature also observed previously in California’s mountainous areas by Kim (2001). In a 283

warmer climate, reductions in snow cover result in an increase in absorbed solar 284

radiation, which are balanced, in part, by increased surface temperatures (Giorgi et al. 285

1997). Early in the year, snow cover at elevations near the snow line is more sensitive to 286

temperature changes than at higher elevations. The decreased snow cover near the 287

approximate snow line results in rings of enhanced warming in March and April. In May 288

and June, snow cover at all elevations may be sensitive to temperature change, leading to 289

larger warming extending all the way up to the mountain peaks. 290

3. Statistical Downscaling 291

We constructed a statistical model to accurately and efficiently approximate the warming 292

patterns that would have been produced had dynamical downscaling been performed on 293

the remaining GCMs. The statistical model scales the dominant spatial pattern (identified 294

14

through EOF analysis of the dynamical warming patterns) and the regional mean so they 295

are consistent with the regionally averaged warming over the Los Angeles region as well 296

as the land-sea contrast in the warming. 297

a. Empirical Orthogonal Function Analysis of Spatial Patterns 298

Empirical orthogonal function (EOF) analysis was performed on the 60 monthly patterns 299

(five models, each with 12 monthly warming patterns) with their regional means removed 300

(Fig. 4). Although EOF analysis is typically applied to temporal anomalies to identify 301

common modes of variability relative to temporal mean, instead we perform EOF 302

analysis on the spatial anomalies to find how warming at in the region differs from the 303

regional average. The leading EOF explains 74% of the variance. It is referred to as the 304

Coastal-Inland Pattern (CIP) henceforth because of its strong positive loadings inland and 305

negative loadings over the coastal zone and ocean. The second and third EOFs (13%, 5% 306

variance explained) may also represent important physical phenomena, but their roles in 307

shaping the warming patterns are much smaller, and we ignore them for the remainder of 308

this paper. 309

The CIP arises from local dynamics modulating the basic contrast in climate 310

between the land and ocean. These dynamics are apparent in other basic variables 311

shaping the region’s climate. For example, there is a very strong negative correlation (r = 312

-0.97) between the CIP and the baseline period annual-mean specific humidity (Fig. 5), a 313

climate variable that also exhibits a significant land-sea contrast in this region. This 314

relationship arises because the ocean is by far the most consistent source of water for 315

evaporation in this region. Air masses over the ocean are rapidly and continuously 316

resupplied with water vapor as necessary to maintain high relative humidity levels. 317

15

Meanwhile, dry air masses over the desert interior remain cut off from moisture sources. 318

In the coastal zone, land-sea breezes and synoptically driven alternations of the onshore 319

and offshore flow pattern (Conil and Hall 2006) generally lead to intermediate moisture 320

levels. Very similar dynamics mediate the warming distribution, as described in Section 321

2d, with relatively small warming over the ocean, intermediate warming over the coastal 322

zone, and larger warming inland. Thus the CIP is an expression of local atmospheric 323

circulation patterns endemic to the region. Because the mechanisms that create the CIP 324

are independent of the particular GCMs we have chosen, we are confident that the CIP 325

can be used to downscale other GCMs. 326

The CIP and the regional mean can be linearly combined to closely approximate 327

the dynamically downscaled warming patterns for each month and for each GCM. When 328

linear regression is used to calculate the combination of the regional mean and the CIP 329

that is closest to the dynamical warming pattern, the resulting approximate warming 330

patterns are within 0.19 ºC of their dynamical counterparts, on average. (When we 331

repeated this calculation omitting the contribution of the CIP, the error more than doubled 332

to 0.39 ºC, indicating the importance of including spatial variations.) Furthermore, at 333

each point in the domain, we calculated the correlation between the dynamically 334

downscaled warming and the linearly approximated warming. The domain average of 335

these correlations is 0.98. This confirms that we can capture nearly all variations in 336

warming just by combining appropriate scalings of the regional mean and the CIP. 337

Therefore, we use these two factors as the basis for our statistical model, as discussed in 338

the next section. 339

16

b. Finding Optimal Sample Locations 340

In order to statistically downscale each of the remaining GCMs in the CMIP5 ensemble, 341

we need to obtain approximate values of the regional mean and land-sea contrast, which 342

we do by sampling the large-scale warming. To find the optimal sample locations, we 343

examined the five GCMs we have dynamically downscaled and identified the points in 344

the large-scale domain that are best correlated with the dynamically downscaled regional 345

mean and land-sea contrast. Since the GCMs have different resolutions, we first 346

interpolated the GCM monthly warming patterns to a common grid (our outermost WRF 347

grid, with 18 km resolution, Fig. 1). The highest correlations between the large-scale 348

warming and the regional mean are found over the adjacent ocean and along the coast 349

(Fig. 6a). Since these correlations were calculated using the monthly averages from each 350

of the five GCMs, they indicate the degree to which sampling at that location would 351

capture both inter-monthly and inter-model variations in the regional mean. If this 352

exercise could be undertaken for all 32 GCMs in the ensemble, the location of the 353

optimal sampling point might be slightly different, due variations in resolution and grid 354

placement between the GCMs. To build in a tolerance for such ensemble-size effects, we 355

sampled over a region encompassing the highest correlated points, rather than just the 356

best-correlated point. The GCM regional mean, RgMean(gcm), is calculated as the 357

average over all the points a rectangular region with longitude bounds [120.5º W, 117.5º 358

W] and latitude bounds [32º N, 34.5º N] shown in Fig. 6a (black box). 359

A similar procedure was used to find the optimal locations to sample the land and 360

the ocean warming for calculation of the GCM land-sea contrast. First, the exact values 361

of the land-sea contrast from the dynamically downscaled warming patterns were 362

17

calculated by taking the dot product of the monthly-mean warming patterns with the CIP. 363

These values were then correlated with the GCM warming interpolated to the common 364

18-km grid (Fig. 6b). The correlations are highest over the high desert of Southern 365

California and Southern Nevada, northeast of our 2-km domain. The GCM inland 366

warming is calculated as the average warming over the rectangular area with longitude 367

bounds [118º W, 113º W] and latitude bounds [34º N, 37.5º N]. To find the location to 368

sample the ocean warming, we repeated this procedure, but using partial correlations with 369

the effect of inland warming removed (Fig. 6c). These partial correlations identify the 370

optimal ocean sampling location to use in conjunction with our previously selected inland 371

location. The GCM ocean warming is calculated as the warming averaged over a 372

rectangular area with longitude bounds [120.5º W, 117.5º W] and latitude bounds [32º N, 373

34º N]. The GCM land-sea contrast, LandSeaContrast(gcm), is calculated as the GCM 374

inland warming minus the GCM ocean warming. If the procedure is reversed, and the 375

optimal ocean location is selected before the optimal inland location, they still end up in 376

nearly identical spots. 377

c. The Prediction Equation 378

The statistical model approximates the dynamically downscaled warming as a linear 379

combination of the scaled regional mean warming in the GCM and the product of the 380

GCM’s land-sea contrast with the coastal-inland pattern. The prediction equation for the 381

statistically downscaled warming is 382

dT(gcm,month,i, j) =α + β ⋅ RgMean(gcm,month)+ γ ⋅ LandSeaContrast(gcm,month)

383

18

where (i,j) are coordinates in the 2km grid and α, β, and γ are coefficients determined by 384

linear regression (Fig. 7). The values of these coefficients are α = 0.14 °C, β = 1.10, γ = 385

1.03. Since β is close to one, the dynamically downscaled regional mean warming varies 386

nearly equally with sampled regional mean warming in the GCM. However, it is shifted 387

up by 0.14 °C, which reflects the fact that the predictor (the warming over the coast and 388

adjacent ocean in the GCMs) must be shifted to a slightly greater value to match the 389

dynamically downscaled regional mean, which encompasses inland areas as well. The 390

dynamically downscaled and GCM-sampled land-sea contrasts are nearly the same, as 391

their ratio is approximately one (γ = 1.03). 392

d. Validation of the Statistical Model 393

Cross-validation was performed to assess how accurately the statistical model replicates 394

the warming patterns produced by the dynamical model. The entire statistical model was 395

rebuilt using only four of the five GCMs, and then used to predict the warming of the 396

remaining GCM. This involved first redoing the EOF analysis to find the CIP. (These 397

alternative patterns are nearly identical no matter which model is left out: The correlation 398

between any two is greater than 0.98. This is additional evidence for the robustness of 399

this pattern in regional warming.) Next, the optimal sampling locations were recalculated. 400

They were similarly located in each case. Finally, linear regression was performed to 401

recalculate the parameters α, β, and γ. Once the model was rebuilt, it was applied to the 402

remaining GCM. This procedure was performed five times in all, with each GCM taking 403

a turn being omitted from calibration and used for testing. This cross-validation technique 404

gives us five sets of predicted warming patterns that are compared to their dynamical 405

19

counterparts. These warming patterns are also used later to assess value added (Section 406

3e). 407

The statistical model consistently reproduces the dynamically downscaled 408

warming pattern for the omitted GCM with a reasonable degree of accuracy (Fig. 8, 409

rightmost columns). The average spatial is correlation between the dynamically and 410

statistically generated annual-mean patterns is 0.95. The average root mean squared error 411

(RMSE) in the annual-mean warming patterns is 0.28 ºC over the five models. This error 412

has to be viewed in the context of the variations the statistical model is intended to 413

capture. The range of the five annual means averaged over the whole domain is 2.1 ºC, 414

about an order of magnitude larger than the error. This error is small enough that 415

substituting the statistical model output for that of the dynamical model does not 416

significantly affect the mean or spread of the ensemble, two of the most important 417

outcomes of a multi-model climate change study like this one. The statistical model is 418

slightly less accurate at reproducing the monthly warming patterns (average RMSE is 419

0.39 ºC) due to greater variety in the monthly patterns. Still, the error is an order of 420

magnitude smaller than the range of the monthly-mean regional-mean warming (3.5 ºC). 421

This gives additional confidence that the statistical model can capture even the monthly 422

warming patterns to a reasonable level of accuracy. 423

e. Value of Incorporating Dynamical Information 424

The goal of this study is to provide an ensemble of projections, as if all 32 GCMs had 425

been dynamically downscaled. Due to computational limitations, only five GCMs were 426

dynamically downscaled and the remaining 27 GCMs were statistically downscaled using 427

our statistical model that incorporates the dynamically downscaled output. A reasonable 428

20

question is whether incorporating the dynamically downscaled output into the statistical 429

model was helpful or if a simple statistical method would have sufficed. We answered 430

this question by comparing the statistically generated warming patterns—generated via 431

cross-validation—and the GCM warming patterns interpolated to the 2-km resolution 432

grid, to see which method produced closer results to the five dynamically downscaled 433

GCMs. Fig. 8 gives a comparison of warming patterns produced by dynamical 434

downscaling, our statistical downscaling technique, and linear interpolation. A 435

comparison to the warming at the nearest GCM grid point, is also included to give an 436

idea of the result if raw GCM data are used, with no downscaling whatsoever. 437

The statistically downscaled warming patterns are clearly the most visually 438

similar to the dynamically generated warming patterns. However, it is important to verify 439

this observation using objective measures of model skill. We used two metrics: spatial 440

correlation and RMSE (divided into errors in the regional mean and errors in the spatial 441

pattern), shown in Table 2. The average spatial correlation between the statistically 442

downscaled annual-mean warming patterns and the dynamically downscaled patterns is 443

0.95, compared to 0.79 and 0.64 for the linear interpolated and raw GCM warming 444

patterns, respectively. This demonstrates that the statistical method is superior to linear 445

interpolation at predicting the spatial variations and sharp gradients in warming. For the 446

monthly average patterns, the statistical model also provides added value over linear 447

interpolation. The value added is somewhat smaller because the statistical model dials up 448

or down only one spatially varying pattern (the CIP), while each month has a slightly 449

different characteristic spatial pattern (Fig. 3). 450

21

The second metric is root mean squared error (RMSE). For the annual warming, 451

the statistical model adds value by capturing the spatial variations. The statistical model’s 452

spatial error is 0.14 ºC, which is a substantial improvement over linear interpolation and 453

the raw GCM values of 0.20 ºC and 0.26 ºC, respectively. The statistical model is also an 454

improvement over interpolation of the monthly patterns, though the improvement is 455

somewhat smaller. Again, this is likely due to the simplicity of using a single spatial 456

pattern for all calendar months. We experimented with using different spatial patterns for 457

each month. However, the gains in accuracy were minimal and were accompanied by 458

problems arising from small sample sizes. (We had only five dynamically downscaled 459

warming patterns each month to calibrate each monthly-varying model, rather than the 60 460

patterns used for the original model.) The statistical model provides no added value in 461

predicting the regional mean warming because the GCM warming averaged over the 462

innermost domain is already a good predictor of the dynamical downscaled regional 463

mean. 464

The biggest advantage of the statistical model comes when we consider the 465

ensemble-mean annual-mean warming. As we have seen, the statistically downscaled, 466

linearly interpolated, and raw GCM warming patterns all have biases relative to 467

dynamical downscaling. However, when we aggregate the approximate warming patterns 468

into a five-model ensemble, the statistical model’s errors cancel out, while those from the 469

other methods do not (Fig. 9). In fact the statistically downscaled ensemble-mean is 470

nearly an unbiased estimator of the dynamically downscaled ensemble mean. The only 471

bias is a slight one at the highest elevations. In contrast, the other two methods have 472

systematic biases as large as 1 ºC in magnitude. These methods give overly smoothed 473

22

land-sea contrasts that fail to resolve the sharp gradients in the warming over the 474

mountains, along the coastline, and in the western part of the domain. Thus, there are 475

large swathes of the region where the statistical model is necessary to provide an accurate 476

characterization of the most likely warming outcome. 477

We note that the error estimates in Table 2 and the patterns in Figs. 8 and 9 are 478

based on the statistical model built on only four GCMs and their associated regional 479

warming patterns. Since each GCM has a unique combination of regional mean and land-480

sea contrast (Fig. 10), when one is left out, there is a large region of the parameter space 481

that goes unrepresented in the calibration of the statistical model. Thus the final statistical 482

model, calibrated using all five GCMs as described in Section 3c, produces results of 483

even higher quality. In fact, due to use of linear regression, which ensures that 484

statistically and dynamically downscaled mean warming match, the biases in the 485

ensemble-mean annual-mean warming are negligible. The final statistical model is used 486

to generate the results discussed from Section 4 onward. 487

4. Ensemble-Mean Warming and Uncertainty 488

The final statistical model (calibrated using all five GCMs) was applied to all 32 CMIP5 489

GCMs with output available for the RCP8.5 scenario. The GCMs have widely varying 490

values of the regional mean and land-sea contrast (Fig. 10). The regional mean values 491

range from 1.4 to 3.3 °C, and land-sea contrast ranges from 0.3 to 1.3 °C. Notably, these 492

two parameters are also uncorrelated, so pattern-scaling using only a single of degree of 493

freedom would be misleading here. The dynamically downscaled GCMs (Fig. 10, 494

highlighted in green) approximately span the range of both parameters, confirming that 495

23

the statistical model has been validated in the same parameter range in which it is 496

applied. The annual-mean warming patterns that result from plugging these parameters 497

into the statistical model are shown in Fig. 11. There is considerable variation among 498

these warming patterns, underscoring the importance of considering multiple global 499

models when doing regional downscaling. 500

The ensemble-mean annual-mean warming pattern, as well as upper and lower 501

bounds of the 95% confidence interval, are shown in Fig. 12. The regional mean warming 502

is 2.3 ºC, with a lower bound of 1.0 ºC and an upper bound of 3.6 ºC. This large inter-503

model spread indicates that the models disagree considerably on the magnitude of 504

warming, even when using the same scenario. However, the global models share the 505

characteristic of more warming inland than over the ocean. The difference in ensemble-506

mean warming between coastal and inland areas is especially dramatic in the summertime 507

(Fig. 13). The average August difference between the inland and coastal areas is 0.6 °C, 508

with certain locations showing warming elevated above the coastal values by as much as 509

1.2 °C (+62%). 510

The winter and spring warming that would occur in the mountains would likely be 511

somewhat larger if we had done dynamical downscaling for all the global models 512

(compare Figs. 3 and 13), because the statistical model underestimates some warming 513

due to snow-albedo feedback. Based on comparisons between the dynamically and 514

statistically downscaled warming patterns for spring (MAM), the springtime ensemble-515

mean warming would be as much as 0.5 ºC or larger in the San Bernardino and San 516

Gabriel Ranges. This is also consistent with the larger errors in the statistical model seen 517

at the highest elevations in Fig. 9g. 518

24

5. Discussion 519

In this paper, we present a hybrid dynamical-statistical approach to downscale the mid-520

century warming signal in 32 CMIP5 GCMs. First, we used dynamical downscaling to 521

produce warming patterns associated with five GCMs. Then, to save computational 522

resources, a statistical model was built that scales the characteristic dynamically derived 523

patterns according to the regional warming sampled from the global model. This 524

statistical model was then used to approximate the warming that would result if the 525

remaining global models were dynamically downscaled. The ensemble-mean regional-526

mean warming was projected to be approximately 2.3 °C, with 95% confidence that the 527

warming is between 1.0 °C and 3.6 °C. Thus, the inter-model differences in the GCM 528

outcomes create significant uncertainty in projections of warming over Southern 529

California. 530

In this hybrid method, statistical downscaling is employed a unique way. First, 531

while statistical models typically relate large-scale GCM output to observations, ours 532

relates GCM output to dynamically downscaled output. This is because our statistical 533

model is designed to be an approximate dynamical model. As we showed in Section 3d, 534

the differences between the dynamically and statistically downscaled patterns are an 535

order of magnitude smaller than inter-model variations in the warming. This means that 536

our statistical downscaling projections for the ensemble-mean warming and spread are 537

reasonable approximations to the projections that would have resulted from dynamically 538

downscaling all 32 models. 539

The second difference is that our statistical model was built to directly predict the 540

temperature change, as opposed to predicting the future period temperatures and then 541

25

differencing them with baseline temperatures, as is typically done. Normally, the 542

empirical relationship employed by a statistical model is derived from the historical time 543

period and then applied to a future time period. This leads to stationarity concerns 544

because the relationship between predictor and predictand may not hold in the future 545

period (Wilby and Wigley 1997). In contrast, our statistical model uses a mathematical 546

relationship between the temperature change in the GCM and the temperature change 547

produced by dynamically downscaling. Therefore, we have a different stationarity 548

assumption—one that is easier to satisfy—that the remaining GCMs have values of mid-549

century regional-mean warming and land-sea contrast within the range of the five we 550

dynamically downscaled. Since this condition is satisfied, we have confidence that the 551

statistical relationships hold for all the GCMs that we downscale. 552

The statistical model adds value by capturing the fine-scale spatial variations in 553

the warming. Inland and mountain locations are expected to warm up considerably more 554

than coastal areas, especially during the summer months. When we compared the 555

statistically downscaled patterns to the raw and linearly interpolated GCM patterns, the 556

statistical model captured these spatial variations much more accurately. Furthermore, 557

when we take an ensemble average of the warming patterns, the errors in the statistically 558

downscaled patterns nearly cancel out, revealing only minor biases. In contrast, the raw 559

and linear interpolated GCM warming patterns have large systematic biases, especially 560

along the coastline and in topographically complex regions that are not resolved well in 561

the GCMs. The statistical model does not improve upon the GCM estimates of regional 562

mean warming estimates, because the GCM warming averaged over our innermost 563

domain is already a good predictor of the dynamically downscaled regional mean. 564

26

Another advantage of our hybrid method is that it reflects our understanding of 565

regional climate dynamics. Some types of statistical models, like those based on artificial 566

neural networks, have the effect of being “black boxes,” where the mathematical 567

relationships have no clear physical interpretation. Unlike those techniques, our method 568

first employs dynamical downscaling, which allows us to identify two important physical 569

mechanisms controlling the warming. The first is that the local atmospheric circulation 570

leads to warming over the coastal ocean similar to that seen over the ocean in GCMs, 571

warming over the coastal zone that is slightly elevated above the ocean values, and much 572

higher warming over inland areas separated from the coast by mountain complexes. The 573

second mechanism, smaller in spatial scope, is snow-albedo feedback, which leads to 574

enhanced warming in the mountains. With this knowledge, we built a statistical model 575

that scales the characteristic spatial pattern (which contains signatures of both 576

mechanisms) to fit with the large-scale land-sea contrast and regional mean warming. 577

Because the warming patterns produced by the hybrid approach reflect physical 578

understanding of the region’s climate, they have an extra layer of credibility. Suppose, for 579

instance, that the real climate does warm more over the interior of western North 580

America than over the northeast Pacific Ocean over the coming decades, as is likely if 581

GCM projections are correct. Given the realistic behavior of the WRF model in 582

distributing humidity and temperature across the landscape, it seems very likely that the 583

associated warming pattern in the greater Los Angeles region would be characterized by 584

sharp gradients separating the desert interior and coastal ocean, and that these gradients 585

would be distributed across the landscape in a way very similar to the regional warming 586

patterns we present here. 587

27

In Part II of this study, we apply the hybrid technique developed here to other 588

scenarios and time periods. We examine the differences between mid-century (2041–589

2060) and end-of-century (2081–2100) warming and demonstrate how emission scenario 590

has a much larger effect at end of the century. We also explore how warming effects the 591

diurnal cycle and the number of extreme heat days. In a separate study, Berg et al. (in 592

preparation) use a similar hybrid dynamical-statistical approach to downscale the CMIP5 593

ensemble’s mid-century precipitation projections to the greater Los Angeles region. 594

595

Acknowledgments 596

Support for this work was provided by the City of Los Angeles and the US Department of 597

Energy as part of the American Recovery and Reinvestment Act of 2009. Additional 598

funding was provided by the National Science Foundation (Grant #1065864, 599

"Collaborative Research: Do Microenvironments Govern Macroecology?") and the 600

Southwest Climate Science Center. The authors would like to thank Dan Cayan for 601

reviewing an early draft of this work. 602

603

References 604

Arritt, R. W., and M. Rummukainen, 2011: Challenges in regional-scale climate 605

modeling. Bull. Amer. Meteor. Soc., 92, 365–368. doi:10.1175/2010BAMS2971.1. 606

607

28

Berg N., A. Hall, F. Sun, S. C. Capps, D. Walton, D. Neelin, and B. Langenbrunner, 608

2014: Mid 21st-century precipitation changes over the Los Angeles region, in 609

preparation. 610

611

Braganza, K., D. J. Karoly, A. C. Hirst, M.E. Mann, P. Stott, R.J. Stouffer, and S. F. B. 612

Tett, 2003: Simple indices of global climate variability and change: Part I—Variability 613

and correlation structure. Climate Dyn., 20, 491– 502, doi:10.007/s00382-002-0286-0. 614

615

Braganza, K., D. J. Karoly, A. C. Hirst, P. Stott, R. J. Stouffer, and S. F. B. Tett, 2004: 616

Simple indices of global climate variability and change: Part II: Attribution of climate 617

change during the twentieth century, Climate Dyn., 22, 823– 838, doi:10.007/s00382-618

004-0413-1. 619

620

Cabré, M., S. A. Solman, and M. N. Nuñez, 2010: Creating regional climate change 621

scenarios over southern South America for the 2020’s and 2050’s using the pattern 622

scaling technique: validity and limitations. Climatic Change, 98, 449–469. 623

doi:10.1007/s10584-009-9737-5. 624

Caldwell, P. M., H.-N. S. Chin, D. C. Bader, and G. Bala, 2009: Evaluation of a WRF 625

based dynamical downscaling simulation over California. Climatic Change, 95, 499–521, 626

doi:10.1007/s10584-009-9583-5. 627

628

29

Cayan, D., E. P. Maurer, M. D. Dettinger, M. Tyree, K. Hayhoe, 2008: Climate change 629

scenarios for the California region. Climatic Change, 87, (Suppl 1):S21–S42, 630

doi:10.1007/s10584-007-9377-6. 631

632

Conil, S., and A. Hall, 2006: Local regimes of atmospheric variability: A case study of 633

Southern California. J. Climate, 19, 4308–4325. doi:10.1175/JCLI3837.1. 634

635

Cubasch, U., and Coauthors, 2001: Projections of future climate change. Climate Change 636

2001: The Scientific Basis, J. T.Houghton et al., Eds., Cambridge University Press, 525–637

582. 638

639

Chen, F., and J. Dudhia, 2001: Coupling an advanced land surface–hydrology model with 640

the Penn State–NCAR MM5 modeling system. Part I: Model implementation and 641

sensitivity. Mon. Wea. Rev., 129, 569–585. 642

643

Dong, B., J. M. Gregory, and R. T. Sutton, 2009: Understanding land-sea warming 644

contrast in response to increasing greenhouse gases. Part I: transient adjustment. J. 645

Climate, 22, 3079–3097. 646

647

Drost, F., D. Karoly, and K. Braganza, 2011: Communicating global climate change 648

using simple indices: An update. Climate Dyn., 39, 989–999, doi:10.1007/s00382-011-649

1227-6. 650

651

30

Duffy, P. B., R. W. Arritt, J. Coquard, W. Gutowski, and Coauthors, 2006: Simulations 652

of present and future climates in the western United States with four nested regional 653

climate models. J. Climate, 19, 873–895. 654

655

Dudhia, J., 1989: Numerical study of convection observed during the winter monsoon 656

experiment using a mesoscale two-dimensional model. J. Atmos. Sci., 46, 3077–3107. 657

658

Fasullo, J. T., 2010: Robust land–ocean contrasts in energy and water cycle feedbacks. J. 659

Climate, 23, 4677–4693. 660

661

Fowler, H. J., S. Blenkinsop, C. Tebaldi, 2007: Linking climate change modeling to 662

impacts studies: Recent advances in downscaling techniques for hydrological modeling. 663

Int. J. Climatol., 27, 1547–1578. 664

665

Giorgi, F., C. S. Brodeur, G. T. Bates, 1994: Regional climate change scenarios over the 666

United States produced with a nested regional climate model. J. Climate, 7, 375–399. 667

doi:http://dx.doi.org/10.1175/1520-0442(1994)007<0375:RCCSOT>2.0.CO;2 668

669

Giorgi, F., J. W. Hurrell, M. R. Marinucci, and M. Beniston, 1997: Elevation dependency 670

of the surface climate change signal: A model study. J. Climate, 10, 288–296. 671

doi:http://dx.doi.org/10.1175/1520-0442(1997)010<0288:EDOTSC>2.0.CO;2 672

673

31

Giorgi, F., et al., 2001: Regional climate information—evaluation and projections. In: 674

Climate Change 2001: The Scientific Basis. Contribution of Working Group I to the 675

Third Assessment Report of the Intergovernmental Panel on Climate Change (J.T. 676

Houghton et al., Eds). Cambridge University Press, Cambridge, United Kingdom and 677

New York, NY, USA, 583–638. 678

679

Giorgi, F., C. Jones, and G. R. Asrar, 2009: Addressing climate information needs at the 680

regional level: the CORDEX framework. World Meteorological Organization (WMO) 681

Bulletin, 58, 175–183. 682

683

Giorgi, F., and L. O. Mearns, 1991: Approaches to regional climate change simulation: A 684

review. Rev. Geophys., 29, 191–216. 685

686

Giorgi, F., L. O. Mearns, 2002: Calculation of average, uncertainty range, and reliability 687

of regional climate changes from AOGCM simulations via the “reliability ensemble 688

averaging” (REA) Method. J. Climate, 15, 1141–1158. 689

doi:http://dx.doi.org/10.1175/1520-0442(2002)015<1141:COAURA>2.0.CO;2 690

691

Hayhoe, K., D. R. Cayan, and Coauthors, 2004: Emissions pathways, climate change, and 692

impacts on California. Proceedings of the National Academy of Sciences of the United 693

States of America, 101, 12422–12427. 694

695

32

Hong, S. Y., and J. O. J. Lim, 2006: The WRF single-moment 6-class microphysics 696

scheme (WSM6). J. Korean Meteor. Soc, 42, 129–151. 697

698

Hughes, M., A. Hall, and R. G. Fovell, 2007: Dynamical controls on the diurnal cycle of 699

temperature in complex topography. Climate Dyn., 29, 277–292. 700

701

Joshi, M. M., J. M. Gregory, M. J. Webb, D. M. Sexton, and T. C. Johns, 2008: 702

Mechanisms for the land/sea warming contrast exhibited by simulations of climate 703

change. Climate Dyn., 30, 455–465. 704

705

Kain, J, S., 2004: The Kain–Fritsch convective parameterization: An update. J. Appl. 706

Meteor. Soc., 43, 170–181. doi:http://dx.doi.org/10.1175/1520-707

0450(2004)043<0170:TKCPAU>2.0.CO;2 708

709

Kawase, H., T. Yoshikane, M. Hara, F. Kimura, T. Yasunari, B. Ailikun, H. Ueda, and T. 710

Inoue, 2009: Intermodel variability of future changes in the Baiu rainband estimated by 711

the pseudo global warming downscaling method. J. Geophys. Res., 114, D24110, 712

doi:10.1029/2009JD011803. 713

714

Kharin, V. V., & F.W. Zwiers, 2002: Climate predictions with multimodel ensembles. J. 715

Climate, 15(7), 793. 716

717

33

Kerr, R. A., 2011: Vital details of global warming are eluding forecasters. Science, 334, 718

173–174. 719

720

Kerr, R.A., 2013: Forecasting regional climate change flunks its first test. Science, 339, 721

638. 722

723

Kim, J., 2001: A nested modeling study of elevation-dependent climate change signals in 724

California induced by increased atmospheric CO2. Geophysical Research Letters, 28, 725

2951–2954. 726

727

Lambert, F. H., & Chiang, J. C., 2007: Control of land‐ocean temperature contrast by 728

ocean heat uptake. Geophysical Research Letters, 34(13), L13704. 729

730

Leung, L. R., L. O. Mearns, F. Giorgi, and R. L. Wilby, 2003: Regional climate research: 731

Needs and opportunities. Bull. Amer. Meteor. Soc. January 2003, 89–95. 732

733

Leung, L. R., Y. Qian, X. Bian, W. M., Washington, J. Han, and J. O. Roads, 2004: Mid-734

century ensemble regional climate change scenarios for the western United States. 735

Climatic Change, 62(1-3), 75–113. 736

737

Lin, Y.-L., R. D. Farley, and H. D. Orville, 1983: Bulk parameterization of the snow field 738

in a cloud model. J. Climate Appl. Meteor., 22, 1065–1092. 739

doi:http://dx.doi.org/10.1175/1520-0450(1983)022<1065:BPOTSF>2.0.CO;2 740

34

741

Manabe, S., R. J., Stouffer, M. J. Spelman, and K. Bryan, 1991: Transient responses of a 742

coupled ocean-atmosphere model to gradual changes of atmospheric CO2. Part I. Annual 743

mean response. J. Climate, 4(8), 785–818. 744

745

Mass, C. F., D. Ovens, K. Westrick, and B.A. Colle, 2002: Does increasing horizontal 746

resolution produce more skillful forecasts? Bull. Amer. Meteor. Soc., 83(3), 407–430. 747

748

Mearns, L. O., I. Bogardi, F. Giorgi, I. Matyasovszky, and M. Palecki, 1999: Comparison 749

of climate change scenarios generated from regional climate model experiments and 750

statistical downscaling, J. Geophys. Res., 104(D6), 6603–6621, 751

doi:10.1029/1998JD200042. 752

753

Meehl, G. A., C. Covey, K.E., Taylor, T. Delworth, R. J. Stouffer, M. Latif, and J.F. 754

Mitchell, 2007: The WCRP CMIP3 multimodel dataset: A new era in climate change 755

research. Bull. Amer. Meteor. Soc. 88(9), 1383–1394. 756

757

Mesinger, F., G. DiMego, E. Kalnay, K. Mitchell, P.C. Shafran, W. Ebisuzaki, and W. 758

Shi, 2006: North American regional reanalysis. Bull. Amer. Meteor. Soc., 87(3), 343–360. 759

760

Mitchell, J. F. B., S. Manabe, T. Tokioka, and V. Meleshko, , 1990: “Equilibrium 761

Climate Change,” in Houghton, J. T., Jenkins, G. J., and Ephraums, J. J. (Eds.), Climate 762

Change: The IPCC Scientific Assessment, Cambridge University Press. 763

35

764

Mlawer, E. J., S. J., Taubman, P. D. Brown, M. J. Iacono, and S. A. Clough, 1997: 765

Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated‐k 766

model for the longwave. Journal of Geophysical Research: Atmospheres (1984–2012), 767

102(D14), 16663–16682. 768

769

Pierce, D. W., T. Das, D. R. Cayan, E. P. Maurer, N. Miller, Y. Bao, and M. Tyree, 2013: 770

Probabilistic estimates of future changes in California temperature and precipitation using 771

statistical and dynamical downscaling. Climate Dyn., 40(3-4), 839–856. 772

773

Rasmussen, R., C. Liu, K. Ikeda, D. Gochis, D. Yates, F. Chen, and E. Gutmann, 2011: 774

High-resolution coupled climate runoff simulations of seasonal snowfall over Colorado: 775

A process study of current and warmer climate. J. Climate, 24(12), 3015–3048. 776

777

Salathé Jr, E. P., L. R. Leung, Y. Qian, and Y. Zhang, 2010: Regional climate model 778

projections for the State of Washington. Climatic Change, 102(1-2), 51–75. 779

780

Santer, B. D., T. M. L. Wigley, M. E. Schlesinger, and J. F. B. Mitchell, 1990: 781

Developing climate scenarios from equilibrium GCM results, MPI Report Number 47, 782

Hamburg. 783

784

Sato T., F. Kimura, and A. Kitoh, 2007: Projection of global warming onto regional 785

precipitation over Mongolia using a regional climate model. J. Hydrology, 333, 144–154. 786

36

787

Schiermeier, Q., 2010: The real holes in climate science. Nature, 463(7279), 284–287. 788

789

Skamarock, W. C., J. B. Klemp, J. Dudhia, D. O. Gill, D. M. Barker, M. G. Duda, X.-Y. 790

Huang, W. Wang, and J. G. Powers, 2008: A description of the Advanced Research WRF 791

version 3. NCAR Tech. Note, NCAR/TN-4751+STR. 792

793

Snyder, M. A., J. L. Bell, L. C. Sloan, P. B. Duffy, and B. Govindasamy, 2002: Climate 794

responses to a doubling of atmospheric carbon dioxide for a climatically vulnerable 795

region. Geophysical Research Letters, 29(11), 9-1. 796

797

Sutton, R. T., B. Dong, and J. M. Gregory, 2007: Land/sea warming ratio in response to 798

climate change: IPCC AR4 model results and comparison with observations. Geophysical 799

Research Letters, 34(2), L02701. 800

801

Taylor, K. E., R. J. Stouffer, and G. A. Meehl, 2012: An overview of CMIP5 and the 802

experiment design. Bull. Amer. Meteor. Soc., 93(4), 485–498. 803

804

Tebaldi, C., R. L. Smith, D. Nychka, and L. O. Mearns, 2005: Quantifying uncertainty in 805

projections of regional climate change: A Bayesian approach to the analysis of 806

multimodel ensembles. J. Climate, 18(10), 1524–1540. 807

808

37

Timbal, B., A. Dufour, and B. McAvaney, 2003: An estimate of future climate change for 809

western France using a statistical downscaling technique. Climate Dyn., 20(7-8), 807–810

823. 811

812

Wilby, R. L., and T. M. L. Wigley, 1997: Downscaling general circulation model output: 813

a review of methods and limitations. Progress in Physical Geography, 21(4), 530–548.814

38

List of Figures 815

FIG. 1: (a) Model setup, with three nested WRF domains at resolutions of 18, 6, and 2 816

km. Topography (in meters) is shown in color. (b) The innermost domain of the regional 817

simulation, with 2 km resolution. Topography is shown in meters. Black dots indicate the 818

locations of the 24 stations used for surface air temperature validation. 819

820

FIG. 2: Validation of WRF dynamical downscaling with against a network of 24 stations, 821

for the period 1995-2001. (a) Simulated versus observed seasonal average climatological 822

temperatures, for each of station. (b) Correlations at each station between the time series 823

of simulated and observed monthly temperature anomalies. Anomalies are relative to the 824

monthly climatology. 825

826

FIG. 3: Monthly-mean surface air temperature change (°C) for the mid-century period 827

(2041-2060) relative to the baseline period (1981-2000) averaged over the five 828

dynamically downscaled GCMs. 829

830

FIG. 4: The spatial patterns associated with the three largest EOFs, in descending order of 831

size. EOF analysis was performed on the monthly warming patterns from the five 832

dynamically downscaled models, with the monthly domain averages removed. EOF 1 833

accounts for 74% of the variance. This EOF is referred to as the Coastal-Inland Pattern 834

(CIP) because of its negative loadings over the coastal land areas and positive loadings 835

inland. EOFs 2 and 3 account for 13% and 5% of the variance, respectively. 836

39

FIG 5: Coastal-Inland Pattern (left) and surface specific humidity climatology (g kg-1) of 837

the baseline period (right). The two spatial patterns are highly correlated (r = -0.97). 838

839

FIG. 6: Correlations between GCM warming (interpolated to an 18-km grid) and the 840

dynamically downscaled (a) regional mean warming and (b) land-sea contrast in the 841

warming. The sampled regional mean warming and inland warming are calculated as 842

averages over the warming in the black boxes in panels (a) and (b), respectively. Panel 843

(c) shows partial correlations between the interpolated GCM warming and the 844

dynamically downscaled land-sea contrast with the effect of the sampled inland warming 845

removed. The ocean warming is calculated as the average over the black box in (c). 846

847

FIG. 7: Scatter plots of dynamically downscaled regional mean warming versus GCM-848

sampled regional mean warming (left), and dynamically downscaled land-sea contrast 849

versus the GCM-sampled land-sea contrast (right). For each GCM (colors), the twelve 850

monthly-mean warming values are shown. Approximations used by the statistical model 851

are shown as black dashed lines. 852

853

FIG.8: Annual-mean warming projections (°C) for five GCMs produced by four different 854

methods: nearest GCM grid box (first column), linear interpolation of GCM (second 855

column), statistical downscaling with the hybrid technique (third column), and dynamical 856

downscaling (fourth column). Projections are for mid-century (2041-2060) relative to the 857

baseline period (1981-2000) under the RCP8.5 scenario. The statistically generated 858

40

pattern for a particular GCM was creating using the hybrid technique applied to the other 859

four GCMs. 860

861

FIG. 9: Annual-mean warming (°C) averaged over five GCMs downscaled using four 862

different methods: (a) nearest GCM grid box, (b) linear interpolation of GCM, (c) 863

statistical downscaling with hybrid technique, (d) dynamical downscaling. Bias of first 864

three methods relative to dynamical downscaling (°C) shown in (e)-(g). 865

866

FIG.10: Annual-mean values of regional-mean warming and land-sea contrast (°C) for 867

each GCM (blue dots) with the ensemble mean (red dot). The five GCMs that are also 868

dynamically downscaled are highlighted in green. 869

870

FIG. 11: . Annual-mean warming patterns (°C) generated by applying the statistical 871

model to all 32 GCMs. Warming patterns are shown for the mid-century period (2041-872

2060) relative to the baseline period(1981-2000), under the RCP8.5 scenario. 873

874

FIG.12: . Ensemble-mean annual-mean warming and upper and lower bounds (°C), based 875

on a 95% confidence interval, for 32 statistically downscaled GCMs run with the RCP8.5 876

scenario. 877

878

FIG.13: Ensemble-mean monthly-mean warming (°C) computed by averaging the 879

monthly statistically downscaled warming patterns over 32 CMIP5 GCMs. 880

881

882

41

TABLE 1. Details of the WCRP CMIP5 global climate models used in this study. Check 882

marks indicate which scenarios are used. Five models were dynamically downscaled 883

(bold). All available models are statistically downscaled using the hybrid method. 884

885 MODEL COUNTRY INSTITUTE RCP2.6 RCP8.5

ACCESS1.0 Australia Commonwealth Scientific and

Industrial Research Organization

ACCESS1.3 Australia Commonwealth Scientific and

Industrial Research Organization

BCC-CSM1.1 China Beijing Climate Center, China

Meteorological Administration

BNU-ESM China College of Global Change and Earth

System Science, Beijing Normal

University

Can-ESM2 Canada Canadian Centre for Climate

Modelling and Analysis

CCSM4 USA National Center for Atmospheric

Research

CESM1(BGC) USA National Science Foundation,

Department of Energy, National

Center for Atmospheric Research

CESM1(CAM5) USA National Science Foundation,

Department of Energy, National

Center for Atmospheric Research

CESM1(WACCM) USA National Science Foundation,

Department of Energy, National

42

Center for Atmospheric Research

CMCC-CM Italy Centro Euro-Mediterraneo per I

Cambiamenti Climatici

CNRM-CM5 France Centre National de Recherches

Meteorologiques

CSIRO-Mk3.6.0 Australia Commonwealth Scientific and

Industrial Research Organization

EC-EARTH Europe EC-Earth Consortium

FGOALS-s2 China LASG, Institute of Atmospheric

Physics, Chinese Academy of

Sciences

FIO-ESM China The First Institute of Oceanography

GFDL-CM3 USA NOAA Geophysical Fluid

Dynamics Laboratory

GFDL-ESM2M USA NOAA Geophysical Fluid Dynamics

Laboratory

GFDL-ESM2G USA NOAA Geophysical Fluid Dynamics

Laboratory

GISS-E2-H USA NASA Goddard Institute for Space

Studies

GISS-E2-R USA NASA Goddard Institute for Space

Studies

HadGEM2-AO UK Met Office Hadley Centre

HadGEM2-CC UK Met Office Hadley Centre

HadGEM2-ES UK Met Office Hadley Centre

INMCM4 Russia Institute for Numerical Mathematics

43

IPSL-CM5A-LR France Institut Pierre Simon Laplace

IPSL-CM5A-MR France Institut Pierre Simon Laplace

MIROC-ESM Japan AORI (U. Tokyo), NIES,

JAMESTEC

MIROC-ESM-

CHEM

Japan AORI (U. Tokyo), NIES,

JAMESTEC

MIROC5 Japan AORI (U. Tokyo), NIES,

JAMESTEC

MPI-ESM-LR Germany Max Planck Institute for

Meteorology

MRI-CGCM3 Japan Meteorological Research Institute

NorESM1-M Norway Norwegian Climate Center

886 887 888

889

890

891

892

893

894

895

896

897

898

899

44

TABLE 2. Comparison of the spatial correlation and root mean squared error (RMSE) for 900

the raw GCM, linear interpolated and the statistically downscaled warming patterns, 901

relative to the dynamically downscaled warming. By virtue of their orthogonality, errors 902

in regional mean and spatial pattern are shown separately. 903

904

Annual Monthly Method

Spatial

Correlation

Spatial

RMSE

(°C)

Regional-

Mean

RMSE (°C)

Spatial

Correlation

Spatial

RMSE

(°C)

Regional-

Mean

RMSE

(°C)

Raw GCM 0.64 0.26 0.21 0.49 0.35 0.26

Linear

Interpolation

0.79 0.20 0.22 0.60 0.29 0.26

Statistical

Model

0.95 0.14 0.23 0.67 0.24 0.27

Note: Bolded values indicate improvements over linear interpolation. 905

906

907

908

909

910

911

912

45

(a)

−130 −125 −120 −115

30

35

40

(b)

−120 −119 −118 −11732.5

33

33.5

34

34.5

35

35.5

36

0 500 1000 1500 2000 2500 3000 913

914

FIG. 1: (a) Model setup, with three nested WRF domains at resolutions of 18, 6, and 2 915

km. Topography (in meters) is shown in color. (b) The innermost domain of the regional 916

simulation, with 2 km resolution. Topography is shown in meters. Black dots indicate the 917

locations of the 24 stations used for surface air temperature validation. 918

919

920

921

922

923

924

925

926

927

928

929

46

5 10 15 20 25 305

10

15

20

25

30

r=0.72,Fallr=0.94,Winterr=0.82,Springr=0.93,Summer

point measurement (°C)

WRF

sim

ulat

ion

(°C)

(a) spatial variability

00.10.20.30.40.50.60.70.80.9

1

Baske

rsfiel

dBea

umon

tBurb

ank

Camari

lloEdw

ards

Fulle

rton

John

Way

neLa

ncas

ter

Long

Beach

Los A

lamito

s

Los A

ngele

sMoja

ve

Mounta

in Wilso

nOxn

ardPalm

dale

Point M

ugu

Riversi

de

San C

lemen

te

Santa

Barbara

Santa

Monica

Van N

uys

Buoy 4

6045

Buoy 4

6025

Sandb

urg

Corre

latio

n Co

effic

ient

(b) temporal variability

930

FIG. 2: Validation of WRF dynamical downscaling with against a network of 24 stations, 931

for the period 1995-2001. (a) Simulated versus observed seasonal average climatological 932

temperatures, for each of station. (b) Correlations at each station between the time series 933

of simulated and observed monthly temperature anomalies. Anomalies are relative to the 934

monthly climatology. 935

936

937

938

939

940

941

942

943

944

945

946

47

Jan Feb Mar Apr

May Jun Jul Aug

Sep Oct Nov Dec

Surface Air Warming (oC)0 1 2 3 4 5

947

FIG. 3: Monthly-mean surface air temperature change (°C) for the mid-century period 948

(2041-2060) relative to the baseline period (1981-2000) averaged over the five 949

dynamically downscaled GCMs. 950

951

952

953

954

955

956

48

EOF 1 EOF 2 EOF 3

957

FIG. 4: The spatial patterns associated with the three largest EOFs, in descending order of 958

size. EOF analysis was performed on the monthly warming patterns from the five 959

dynamically downscaled models, with the monthly domain averages removed. EOF 1 960

accounts for 74% of the variance. This EOF is referred to as the Coastal-Inland Pattern 961

(CIP) because of its negative loadings over the coastal land areas and positive loadings 962

inland. EOFs 2 and 3 account for 13% and 5% of the variance, respectively. 963

964

965

966

967

968

969

970

971

972

973

49

Coastal Inland Pattern (EOF 1)

−6 −4 −2 0 2 4 6(10−3 g kg−1)

Baseline Specific Humidity

11 10 9 8 7

974

FIG 5: Coastal-Inland Pattern (left) and surface specific humidity climatology (g kg-1) of 975

the baseline period (right). The two spatial patterns are highly correlated (r = -0.97). 976

977

978

979

980

981

982

983

984

985

986

987

988

989

50

−130 −125 −120 −115

28

30

32

34

36

38

40

42

44 (c)

Correlation Correlation Correlation

−130 −125 −120 −115

28

30

32

34

36

38

40

42

44 (a)

0.7 0.75 0.8 0.85 0.9 0.95 1

−130 −125 −120 −115

28

30

32

34

36

38

40

42

44 (b)

0 0.2 0.4 0.6 0.8 1 −0.5 0 0.5

990

FIG. 6: Correlations between GCM warming (interpolated to an 18-km grid) and the 991

dynamically downscaled (a) regional mean warming and (b) land-sea contrast in the 992

warming. The sampled regional mean warming and inland warming are calculated as 993

averages over the warming in the black boxes in panels (a) and (b), respectively. Panel 994

(c) shows partial correlations between the interpolated GCM warming and the 995

dynamically downscaled land-sea contrast with the effect of the sampled inland warming 996

removed. The ocean warming is calculated as the average over the black box in (c). 997

998

999

1000

1001

1002

1003

1004

1005

1006

1007

51

0 1 2 3 40

0.5

1

1.5

2

2.5

3

3.5

4

4.5

GCM Sampled (oC)

Dyna

mica

lly D

owns

cale

d (o C)

Regional Mean Warming

r = 0.95

CCSM4GFDL−CM3MIROC−ESM−CHEMCNRM−CM5MPI−ESM−LR

−1 0 1 2 3−1

−0.5

0

0.5

1

1.5

2

2.5

3

GCM Sampled (oC)

Dyna

mica

lly D

owns

cale

d (o C)

Land−Sea Contrast

r = 0.79

1008

FIG. 7: Scatter plots of dynamically downscaled regional mean warming versus GCM-1009

sampled regional mean warming (left), and dynamically downscaled land-sea contrast 1010

versus the GCM-sampled land-sea contrast (right). For each GCM (colors), the twelve 1011

monthly-mean warming values are shown. Approximations used by the statistical model 1012

are shown as black dashed lines. 1013

1014

1015

1016

1017

1018

1019

1020

1021

1022

1023

1024

52

Dynamical Statistical Linear Interp. Raw GCM

CCSM

4G

FDL−

CM3

MIR

OC−

ESM−C

HEM

CNRM

−CM

5M

PI−E

SM−L

R

Surface Air Warming (oC)

0 1 2 3 4 5

1025

FIG.8: Annual-mean warming projections (°C) for five GCMs produced by four different 1026

methods: nearest GCM grid box (first column), linear interpolation of GCM (second 1027

column), statistical downscaling with the hybrid technique (third column), and dynamical 1028

downscaling (fourth column). Projections are for mid-century (2041-2060) relative to the 1029

baseline period (1981-2000) under the RCP8.5 scenario. The statistically generated 1030

pattern for a particular GCM was creating using the hybrid technique applied to the other 1031

four GCMs. 1032

53

oC

−1

−0.5

0

0.5

1

oC

0

1

2

3

4

5(a) Raw GCM (b) Linear Interp. (c) Statistical (d) Dynamical

(e) Raw GCM (f) Linear Interp. (g) Statistical

1033

FIG. 9: Annual-mean warming (°C) averaged over five GCMs downscaled using four 1034

different methods: (a) nearest GCM grid box, (b) linear interpolation of GCM, (c) 1035

statistical downscaling with hybrid technique, (d) dynamical downscaling. Bias of first 1036

three methods relative to dynamical downscaling (°C) shown in (e)-(g). 1037

1038

1039

1040

1041

1042

1043

1044

1045

1046

1047

1048

54

0 0.5 1 1.5 2 2.5 3 3.5 40

0.5

1

1.5

GFDL−ESM2M

GFDL−ESM2G

GFDL−CM3

CNRM−CM5

CanESM2

GISS−E2−R

HadGEM2−CC

HadGEM2−ES

IPSL−CM5A−LR

IPSL−CM5A−MR

MIROC−ESM−CHEM

MIROC−ESMMIROC5

MPI−ESM−LRMRI−CGCM3

NorESM1−M

INM−CM4

ACCESS1.0

ACCESS1.3

BNU−ESM

CESM1(BGC)

CESM1(CAM5)

CESM1(WACCM)

CMCC−CM

CSIRO−Mk3.6.0

EC−EARTH

FGOALS−s2

FIO−ESM

GISS−E2−H HadGEM2−AO

BCC−CSM1.1

CCSM4

Ensemble Mean

Regional Mean Warming (oC)

Land−S

ea C

ontra

st (o C)

1049

FIG.10: Annual-mean values of regional-mean warming and land-sea contrast (°C) for 1050

each GCM (blue dots) with the ensemble mean (red dot). The five GCMs that are also 1051

dynamically downscaled are highlighted in green. 1052

1053

1054

1055

1056

1057

1058

1059

1060

55

ACCESS1.0 ACCESS1.3 BCC−CSM1.1 BNU−ESM CanESM2 CCSM4 CESM1(BGC) CESM1(CAM5)

CESM1(WACCM) CMCC−CM CNRM−CM5 CSIRO−Mk3.6.0 EC−EARTH FGOALS−s2 FIO−ESM GFDL−CM3

GFDL−ESM2G GFDL−ESM2M GISS−E2−H GISS−E2−R HadGEM2−AO HadGEM2−CC HadGEM2−ES INM−CM4

IPSL−CM5A−LR IPSL−CM5A−MR MIROC−ESM MIROC−ESM−CHEM MIROC5 MPI−ESM−LR MRI−CGCM3 NorESM1−M

Surface Air Warming (oC)

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

1061

FIG. 11: Annual-mean warming patterns (°C) generated by applying the statistical model 1062

to all 32 GCMs. Warming patterns are shown for the mid-century period (2041-2060) 1063

relative to the baseline period(1981-2000), under the RCP8.5 scenario. 1064

1065

1066

1067

1068

1069

1070

1071

1072

1073

1074

1075

56

Lower Bound Ensemble Mean

Surface Air Warming (oC)

Upper Bound

0 1 2 3 4 5

1076

FIG.12: Ensemble-mean annual-mean warming and upper and lower bounds (°C), based 1077

on a 95% confidence interval, for 32 statistically downscaled GCMs run with the RCP8.5 1078

scenario. 1079

1080

1081

1082

1083

1084

1085

1086

1087

1088

1089

1090

1091

1092

57

Jan Feb Mar Apr

May Jun Jul Aug

Sep Oct Nov Dec

Surface Air Warming (oC)

0 1 2 3 4 5

1093

FIG.13: Ensemble-mean monthly-mean warming (°C) computed by averaging the 1094

monthly statistically downscaled warming patterns over 32 CMIP5 GCMs. 1095