exploring the impacts of physics and resolution on aqua...

41
1 Exploring the impacts of physics and resolution on aqua-planet 1 simulations from a non-hydrostatic global variable-resolution modeling 2 framework 3 1 Chun Zhao, 1 L. Ruby Leung, 2 Sang-Hun Park, 1 Samson Hagos, 1 Jian Lu, 1 Koichi 4 Sakaguchi, 1,3 Jinho Yoon, 1 Bryce E. Harrop, 2 William Skamarock, 2 Michael G. Duda 5 6 7 8 1 Atmospheric Sciences and Global Change Division, Pacific Northwest National 9 Laboratory, Richland, WA, USA 10 2 National Center for Atmospheric Research, Boulder, CO, USA 11 3 School of Earth Sciences and Environmental Engineering, Gwanju Institute of Science 12 and Technology, Gwangju, South Korea 13 14 15 16 Manuscript for submission to J. Adv. Model. Earth Syst. 17 18 *Corresponding author: Chun Zhao, phone: (509) 371-6372; email: [email protected] 19 20 Key points: 21 1. Aqua-planet simulations from hydrostatic (H) and non-hydrostatic (NH) MPAS are 22 investigated across various horizontal resolutions. 23 2. Simulated precipitation, column integrated moisture, and large-scale circulation is less 24 sensitive to resolution in NH-MPAS than H-MPAS. 25 3. Variable resolution simulation with NH-MPAS exhibits more zonally symmetric 26 features in the refined region compared to that with H-MPAS. 27 28 29

Upload: others

Post on 01-Apr-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

1

Exploring the impacts of physics and resolution on aqua-planet 1

simulations from a non-hydrostatic global variable-resolution modeling 2

framework 3 1Chun Zhao, 1L. Ruby Leung, 2Sang-Hun Park, 1Samson Hagos, 1Jian Lu, 1Koichi 4

Sakaguchi, 1,3Jinho Yoon, 1Bryce E. Harrop, 2William Skamarock, 2Michael G. Duda 5

6

7

8 1Atmospheric Sciences and Global Change Division, Pacific Northwest National 9

Laboratory, Richland, WA, USA 10 2National Center for Atmospheric Research, Boulder, CO, USA 11

3School of Earth Sciences and Environmental Engineering, Gwanju Institute of Science 12

and Technology, Gwangju, South Korea 13

14

15

16

Manuscript for submission to J. Adv. Model. Earth Syst. 17

18

*Corresponding author: Chun Zhao, phone: (509) 371-6372; email: [email protected] 19

20

Key points: 21

1. Aqua-planet simulations from hydrostatic (H) and non-hydrostatic (NH) MPAS are 22 investigated across various horizontal resolutions. 23 2. Simulated precipitation, column integrated moisture, and large-scale circulation is less 24 sensitive to resolution in NH-MPAS than H-MPAS. 25 3. Variable resolution simulation with NH-MPAS exhibits more zonally symmetric 26 features in the refined region compared to that with H-MPAS. 27 28

29

2

Abstract. 30

The non-hydrostatic Model for Prediction Across Scales (NH-MPAS) provides a global 31

framework to achieve high resolution using regional mesh refinement. Previous studies 32

using the hydrostatic version of MPAS (H-MPAS) with the physics parameterizations of 33

Community Atmosphere Model version 4 (CAM4) found notable resolution dependent 34

behaviors. This study revisits the resolution sensitivity using the NH-MPAS with both 35

CAM4 and CAM5 physics. A series of aqua-planet simulations at global quasi-uniform 36

resolutions and global variable resolution simulations with a regional mesh refinement 37

over the tropics are analyzed, with a primary focus on the distinct characteristics of NH-38

MPAS in simulating precipitation, clouds, and large-scale circulation features compared 39

to H-MPAS-CAM4. The resolution sensitivity of total precipitation and column 40

integrated moisture in NH-MPAS is smaller than that in H-MPAS-CAM4. This 41

contributes importantly to the reduced resolution sensitivity of large-scale circulation 42

features such as the inter-tropical convergence zone and Hadley circulation in NH-MPAS 43

compared to H-MPAS. In addition, NH-MPAS shows almost no resolution sensitivity in 44

the simulated westerly jet, in contrast to the obvious poleward shift in H-MPAS with 45

increasing resolution, which is partly explained by differences in the hyperdiffusion 46

coefficients used in the two models that influence wave activity. With the reduced 47

resolution sensitivity, simulations in the refined region of the NH-MPAS global variable 48

resolution configuration exhibit zonally symmetric features that are more comparable to 49

the quasi-uniform high-resolution simulations than those from H-MPAS that displays 50

zonal asymmetry in simulations inside the refined region. Overall, NH-MPAS with 51

CAM5 physics shows less resolution sensitivity compared to CAM4. 52

53

3

1. Introduction 54

Reliable climate prediction at regional scales remains one of the greatest 55

challenges in climate change science (IPCC, 2013). Although not a panacea for climate 56

modeling (NRC, 2012), increasing model horizontal resolution has been shown to have 57

some positive impacts on simulating regional climate (e.g., Giorgi and Mearns, 1991; 58

Giorgi and Marinucci, 1996; Leung et al., 2003; Bacmeister et al. 2014). Dynamical 59

downscaling using nesting and stretched grid modeling has been used since the 1990s to 60

achieve regional-scale simulations (Wang et al. 2004). More recently, it has become 61

computationally feasible to perform global simulations at grid sizes of 25 km to 50 km. 62

For example, simulations at a horizontal resolution of ~25 km were used to study tropical 63

cycle changes in a warmer climate (Murakami et al., 2010, 2012; Roberts et al., 2015; 64

Wehner et al., 2014, 2015). However, limited by computing resources such simulations 65

are still not routinely performed or considered standard in model intercomparison 66

experiments. Through advances in grid generation techniques and numeric capability of 67

handling distorted or non-orthogonal grids, global variable-resolution modeling offers an 68

alternative approach for modeling regional climate at reduced computational cost 69

compared to global uniform high-resolution modeling (e.g., Medvigy et al., 2013; 70

Rauscher et al., 2013; Zarzycki et al., 2014; Sakaguchi et al. 2015). Such modeling 71

frameworks can also be used as an effective tool for assessing the resolution dependence 72

or scalability of physical parameterizations (Rauscher et al. 2013; Hagos et al. 2013). 73

The Model for Prediction Across Scales (MPAS) is a variable-resolution model 74

that uses unstructured spherical centroidal Voronoi tessellations (SCVTs) (Du et al., 1999; 75

Ringler et al., 2008). A single scalar density function is used to generate global variable 76

resolution meshes with local refinement (e.g., Ju et al., 2011) over single or multiple 77

high-resolution regions. The MPAS atmospheric dynamical core has been evaluated 78

using the standard shallow-water test case suite of Williamson et al. (1992) for both 79

quasi-uniform (Ringler et al., 2010) and variable resolution (Ringler et al., 2011) meshes. 80

The hydrostatic version of the MPAS dynamical core (H-MPAS) was coupled with the 81

physics parameterizations of Community Atmosphere Model (CAM) version 4 (CAM4, 82

Neale et al. 2010) (H-MPAS-CAM4, Rauscher et al., 2013) for climate simulations. H-83

MPAS-CAM4 has been evaluated in both idealized (Rauscher et al. 2013) and realistic 84

4

(Sakaguchi et al., 2015) configurations. A series of studies have investigated the 85

resolution sensitivity of H-MPAS-CAM4 in simulating the scaling behavior of clouds 86

and precipitation (O’Brien et al. 2013), the structure of the Inter-Tropical Convergence 87

Zone (ITCZ) (Landu et al., 2014), precipitation extremes (Yang et al., 2014), atmospheric 88

river frequency (Hagos et al., 2015), and the position and strength of the eddy-driven jet 89

(Lu et al., 2015). 90

The aforementioned studies have revealed notable resolution dependence in 91

quasi-uniform resolution aqua-planet simulations with H-MPAS-CAM4. For example, 92

Landu et al. (2014) found a transition from a single to a double ITCZ as the MPAS mesh 93

spacing is reduced from 240km to 30km. The frequency of atmospheric rivers in the 94

same set of simulations decreases dramatically with increasing model resolution that also 95

shifts the subtropical jet poleward, reducing the likelihood of extratropical cyclones 96

tapping tropical moisture (Hagos et al. 2015). Lu et al. (2015) investigated the jet 97

sensitivity to model resolutions using wave activity budgets and found a systematic 98

weakening and poleward shift of the eddy-driven jet associated with the smaller effective 99

diffusivity with increasing horizontal resolution. Rauscher et al. (2013) and Hagos et al. 100

(2013) analyzed the zonally asymmetric circulation features in variable resolution H-101

MPAS-CAM4 aqua-planet simulations with a regional refinement over the tropics. They 102

attributed the anomalous circulation to the enhanced precipitation downwind of the 103

refined domain, and the Rossby waves excited by the anomalous diabatic heating (Gill 104

1980). The diverse resolution dependence behaviors found in these studies are partly 105

attributable to the resolution dependent physics parameterizations, particularly related to 106

the moist processes, but the ability to better resolve the eddies as model resolution 107

increases also plays an important role in simulations of large-scale features such as the jet 108

stream. Arguably both factors as well as their interactions contribute importantly to the 109

resolution-dependent performance of H-MPAS-CAM4, with implications for its 110

effectiveness as a multi-resolution modeling framework. 111

Parallel to the development of H-MPAS and its coupling with CAM4, Skamarock 112

et al. (2012; 2014) developed a non-hydrostatic version of MPAS (NH-MPAS). Recently, 113

the NH-MPAS dynamical core has been coupled with both CAM4 and CAM5 physics 114

packages (Neale et al., 2010 and 2012) (NH-MPAS-CAM4 and NH-MPAS-CAM5) in 115

5

the Community Earth System Model (CESM) framework for climate research (Park et al., 116

2016). This offers an opportunity to revisit the resolution dependence of MPAS to 117

document whether and how the resolution sensitivity may differ in the two versions of 118

MPAS dynamical cores and to evaluate the impacts of physics parameterizations, as the 119

CAM4 and CAM5 physics packages feature important differences in cloud microphysics 120

and turbulence parameterizations. Since previous studies have extensively explored the 121

performance of H-MPAS-CAM4, this study will primarily highlight the distinct 122

characteristics of NH-MPAS with CAM4 and CAM5 physics from H-MPAS-CAM4 in 123

aqua-planet simulations. Analysis will focus on the physics and resolution impacts on 124

simulated precipitation, clouds, and large-scale circulation. 125

The rest of the manuscript is organized as follows. Section 2 describes briefly the 126

MPAS dynamical core, the CAM physics parameterizations, and the model setup for 127

aqua-planet configuration. The series of aqua-planet experiments using NH-MPAS with 128

CAM4 and CAM5 physics are analyzed in section 3, with comparison to the previous 129

results from H-MPAS-CAM4. Comparisons are first discussed in the context of global 130

quasi-uniform resolution simulations at a range of horizontal resolutions, followed by 131

analysis of global variable resolution simulations. The findings are then summarized in 132

section 4. 133

134

2. Models and experiments 135

2.1 MPAS-Atmosphere (MPAS-A) dynamical cores 136

The MPAS atmospheric dynamical core uses the horizontal C-grid scheme 137

(Thuburn et al., 2009; Ringler et al., 2010) while the vertical and time discretization 138

follows that of the Weather Research and Forecasting (WRF) model (Skamarock et al., 139

2008). The MPAS-A hydrostatic and non-hydrostatic solvers are described in Park et al. 140

(2013) and Skamarock et al. (2012), respectively. The most significant difference 141

between the hydrostatic and non-hydrostatic atmospheric solvers is that H-MPAS uses a 142

pressure coordinate while NH-MPAS uses a height coordinate. Thus, the vertical levels 143

are necessarily different in the H-MPAS and NH-MPAS configurations in our 144

experiments. More details about the hydrostatic hybrid-sigma coordinate and non-145

hydrostatic hybrid-height coordinate can be found in Park et al. (2013). 146

6

MPAS-A can be configured to use any refined meshes where the line segment 147

connecting cell centers is orthogonal to the edge shared by the two cells. Figure 1 shows 148

the quasi-uniform and variable resolution meshes used in this study, with the meshes 149

coarsened to display the structure of the individual mesh cells. The quasi-uniform mesh 150

has essentially the same mesh spacing globally, while the variable resolution mesh has 151

finer mesh spacing in the refined region with a transition zone between the fine and 152

coarse resolution meshes. More details about the mesh generation can be found in Ringler 153

et al. (2011). 154

2.2 CAM physics parameterizations 155

The Community Atmosphere Model (CAM) is the atmospheric component of the 156

Community Earth System Model (CESM). Implemented in the CESM framework, H-157

MPAS and NH-MPAS are coupled with both CAM4 and CAM5 physics packages in the 158

same way as other dynamical cores in CESM (Rauscher et al. 2013; Park et al., 2016). By 159

default, CAM4 and CAM5 use 26 and 30 vertical levels, respectively. Both physics 160

packages are tuned with respective to their own distributions of vertical levels, which 161

should be considered intrinsic to each CAM physics package. More details about the 162

physics-dynamics coupling and vertical level distributions in CESM with H-MPAS and 163

NH-MPAS are described in Park et al. (2016). Large differences exist in physics and 164

chemistry parameterizations between CAM4 and CAM5. The CAM4 and CAM5 physics 165

were described fully by Neale et al. (2010) and Neale et al. (2012), respectively. Here, 166

some relevant details are briefly summarized. 167

CAM4 uses the Rasch-Kristjansson single moment bulk microphysics 168

parameterization that prognoses cloud condensate (with liquid and ice as separate classes) 169

and includes formulations of autoconversion, collection/accretion, and sedimentation 170

(Rasch and Kristjansson 1998; Neale et al. 2010). CAM5 uses the Morrison-Gettelman 171

two-moment bulk microphysics parameterization (Morrison and Gettelman 2008; 172

Gettelman et al. 2008; Neale et al. 2012) that prognoses cloud condensate (with liquid 173

and ice as separate classes) and cloud droplet number. It includes formulations of droplet 174

activation, condensation-deposition, evaporation-sublimation, autoconversion, accretion, 175

collection, sedimentation, homogenous and heterogeneous freezing, and ice melting. The 176

Zhang (Zhang et al., 2003) and Park (Neale et al., 2012) macrophysics parameterizations 177

7

are used in CAM4 and CAM5, respectively. The Zhang macrophysics parameterization 178

calculates the stratiform, stratus, and convective cloud fraction separately, and diagnoses 179

the total cloud fraction for a given grid cell as the convective cloud fraction plus the 180

larger of the stratiform and stratus cloud fractions. The Park macrophysics 181

parameterization does not calculate a separate stratus cloud fraction, and total fraction is 182

calculated as the sum of the stratiform and convective cloud fractions. Both CAM4 and 183

CAM5 use the modified Zhang and McFarlane (1995) deep convective parameterization 184

(Neale et al., 2012). However, CAM5 adopts a shallow convective scheme that uses a 185

realistic plume dilution equation and closure to accurately simulate the spatial 186

distribution of shallow convective activity (Park and Bretherton, 2009). In addition, 187

CAM4 uses a non-local transport scheme for the planetary boundary layer turbulence, 188

while CAM5 uses a TKE moisture turbulence scheme (Bretherton and Park, 2009; Neale 189

et al., 2012). For chemistry, CAM4 uses the bulk aerosol scheme, while CAM5 uses a 190

modal aerosol scheme (Liu et al., 2012; Neale et al., 2012). In addition to online 191

simulation of aerosols, both physics packages can be run with prescribed aerosol 192

concentrations, which are used in our aqua-planet experiments. 193

2.3 Numerical experiments 194

Following Rauscher et al. (2013), four different quasi-uniform resolution (UR) 195

meshes with mesh spacing of approximately 30, 60, 120, and 240 km and a variable 196

resolution (VR) mesh, similar to that shown in Fig. 1, have been configured. The VR 197

mesh has a circular high-resolution region centered at the equator. The refined mesh has a 198

radius of 30 degrees and a mesh spacing of approximately 30 km, with a transition zone 199

of 20 degrees within which the mesh spacing gradually increases to approximately 120 200

km. The numbers of grid cells of these meshes are listed in Table 1. The mesh spacing in 201

these experiments resides in the hydrostatic regime. The same parameter values for the 202

common physics schemes in CAM4 and CAM5 are used at multiple resolutions to 203

explore the simulation sensitivity only to mesh resolution. Hence the physics time step 204

(10 min), dynamics time step (100 sec), and moist-physics parameters are the same 205

across all experiments that use the same physics. Following the hierarchical framework 206

of Leung et al. (2013) and demonstrated by Rauscher et al. (2013) and Hagos et al. 207

(2013), the UR experiments are used to quantify the impacts of horizontal resolution, 208

8

while the VR experiments are mainly used to explore the potential of regional refinement 209

for regional climate simulation. 210

The aquaplanet experiments are configured following Neale and Hoskins (2000) 211

with SST distribution prescribed as follows: 212

𝑇! 𝜆,𝜙 = 27 1− 𝑠𝑖𝑛!3𝜙2

, 𝑖𝑓 −𝜋3< 𝜙 <

𝜋3

0, 𝑂𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒

Each simulation is run for 3 years, with the first 6 months discarded for spinup, leaving 213

2.5 years for analysis. Some sensitivity experiments at coarser mesh spacing were 214

extended to 5 years and the analysis results were comparable to those of the shorter 215

simulations. The H-MPAS-CAM4 used in this study is the same model used in Rauscher 216

et al. (2013), except for a bug fix in the interface between H-MPAS and the CAM4 217

physics that eliminated a small moisture imbalance between the global precipitation and 218

evaporation found in the older version; its regional impact is visible only in the tropics. 219

With this bug fix, our simulations show some differences from previously reported H-220

MPAS-CAM4 simulations (Rauscher et al. 2013), but the general model behaviors are 221

largely comparable.. 222

223

3. Results 224

3.1 Global quasi-uniform resolution simulations 225

3.1.1 Precipitation 226

Figure 2 shows the zonal mean precipitation rates from the UR NH-MPAS 227

simulations with CAM4 and CAM5 and the UR H-MPAS-CAM4 simulations at 228

approximately 30, 60, 120, and 240 km mesh spacing. Model outputs from simulations at 229

various resolutions are all regridded to the 2°×2° finite volume grid, but similar results 230

are obtained using model outputs at the original resolution of each experiment. Since 231

precipitation and its sensitivity to resolution beyond 60°N/S are subtle, only results 232

between 60°N/S are shown in Figure 2. All simulations produce the highest precipitation 233

at the equator (the ITCZ), with a secondary peak in the midlatitudes around 40°N/S 234

corresponding to the storm tracks and a minimum in the subtropics. In Rauscher et al. 235

(2013) and Landu et al. (2014), H-MPAS-CAM4 produced a single ITCZ at 120 and 240 236

km but a double ITCZ at 30 km and 60 km. Rauscher et al. (2013) noted a dependence of 237

9

the ITCZ structure on the model hyperdiffusion coefficient, as a sensitivity experiment 238

with a smaller hyperdiffusion coefficient yielded a single ITCZ even at high resolution. 239

Landu et al. (2014) attributed the double ITCZ feature in H-MPAS-CAM4 at high 240

resolution to a feedback between convection and large-scale circulation. However, our 241

simulations with H-MPAS and NH-MPAS all produce a single ITCZ regardless of model 242

resolution. As discussed in Section 2.3, the H-MPAS-CAM4 used in this study includes a 243

bug fix that balances the global precipitation and evaporation more precisely. This results 244

in increased atmospheric moisture in the simulations, as evidenced by the higher 245

precipitation rate in our simulations compared to that reported previously at comparable 246

resolutions (e.g., Landu et al. 2014; Yang et al. 2014). Consistent with the mechanisms 247

discussed in Landu et al. (2014), increased moisture increases the latent heat released by 248

convection in the tropics. This induces larger upper tropospheric warming near the 249

equator that increases poleward heat transport. The latter warms the upper troposphere 250

and increases atmospheric stability poleward, limiting CAPE and the ITCZ to a narrower 251

region centering about the equator. Hence, with the increased moisture resulting from the 252

bug fix, H-MPAS-CAM4 produce a single ITCZ even at high resolution compared to the 253

double ITCZ in the H-MPAS-CAM4 high resolution simulations reported by Laudu et al. 254

(2014). NH-MPAS simulates even higher moisture content than H-MPAS, so a single 255

ITCZ is produced at all resolutions. 256

To better quantify the sensitivity of precipitation to resolutions and physics, 257

Figure 3 shows the mean tropical (10°S-10°N) total, parameterized convective, and 258

resolved large-scale precipitation simulated at different resolutions. As stronger updrafts 259

are resolved with reduced mesh spacing, large-scale precipitation rates increase with 260

increasing resolutions (Williamson, 2008a), while convective precipitation rates decrease 261

as moisture is reduced by large-scale condensation. Hence, the ratio of convective to 262

large-scale precipitation decreases with increasing resolutions. For NH-MPAS, this ratio 263

is more sensitive to resolution with the CAM5 physics compared to the CAM4 physics. 264

For the same physics (CAM4), H-MPAS generally simulates lower convective 265

precipitation rates and higher large-scale precipitation rates than NH-MPAS at all 266

resolutions, and the differences are larger at higher resolution. 267

10

Because of the opposite sensitivity of the simulated convective and large-scale 268

precipitation rates to resolution in NH-MPAS-CAM5, the total precipitation increases 269

only by a small amount of 4% (9.1 to 9.5 mm day-1) with resolutions from 240 km to 30 270

km, and is almost insensitive (9.4 to 9.5 mm day-1) to resolutions from 120 km to 30 km. 271

The NH-MPAS-CAM4 simulated total precipitation is lower and more sensitive to 272

resolution (11%, 7.8 to 8.7 mm day-1 from 240 km to 30 km and 5%, 8.3 to 8.7 mm day-1 273

from 120 km to 30 km) than NH-MPAS-CAM5. With CAM4 physics, the resolution 274

sensitivity of total precipitation for H-MPAS and NH-MPAS are comparable, although 275

differences between the sensitivity of convective and large-scale precipitation to 276

resolutions are large between H-MPAS and NH-MPAS. The sensitivity of the H-MPAS-277

CAM4 simulated total precipitation to resolution in this study is smaller than that 278

reported by Rauscher et al. (2013), which is related to the moisture balance issue noted 279

earlier. Overall, the total precipitation simulated in all experiments in this study is less 280

sensitive to resolutions compared to the results in Rauscher et al. (2013). 281

In addition to the mean, the probability density functions (PDFs) of hourly 282

tropical (10°S-10°N) precipitation for lower intensity events (0-150 mm day-1, 1 mm day-283 1 bins) and higher intensity events (0-1000 mm day-1, 10 mm day-1 bins) are shown in 284

Figure 4. The results from simulations at various resolutions are all regridded to the 2°×2° 285

grid. For lower intensity events (Figure 4a), it is evident that the coarser resolution 286

simulations produce lower frequencies of more intense precipitation and higher 287

frequencies of less intense precipitation compared to simulations at higher resolutions, 288

with a transition occurring at around 45 mm day-1. At the same resolution, CAM5 289

simulates higher frequencies of more intense precipitation than CAM4 in NH-MPAS and 290

H-MPAS-CAM4 simulates slightly higher frequencies of more intense precipitation than 291

NH-MPAS-CAM4. Although, the mean precipitation has weak sensitivity to resolution 292

from 120 km to 30 km (Fig. 3), the impacts of resolution on the precipitation event 293

frequencies are evident. The frequency distributions from simulations at 60 km and 30 294

km are similar for NH-MPAS with CAM5 and CAM4. 295

For higher intensity events (Figure 4b), resolution has significant impacts on 296

extreme events even from 60 km to 30 km. The NH-MPAS-CAM5 simulations produce 297

higher frequencies of extreme precipitation than NH-MPAS-CAM4. The results from H-298

11

MPAS-CAM4 and NH-MPAS-CAM4 are similar. Overall, resolution has significant 299

impacts on precipitation intensity, especially for the low probability – high intensity 300

events, with no sign of convergence even at 30 km resolution, and simulations with 301

CAM5 physics produce more intense precipitation than CAM4. Figure 5 shows the PDFs 302

of 6-hourly tropical (10°S-10°N) upward vertical velocity averaged below 700 hPa. In 303

general, the PDFs of extreme precipitation are consistent with those of upward vertical 304

velocity. The figure shows that NH-MPAS-CAM5 produces higher frequencies of 305

stronger updrafts than NH-MPAS-CAM4. The distributions of vertical velocity from H-306

MPAS-CAM4 and NH-MPAS-CAM4 are similar with slightly more frequent extreme 307

updrafts in the former. The differences in vertical coordinate between NH-MPAS and H-308

MPAS and vertical resolution between CAM4 and CAM5 may contribute partly to their 309

differences in PDFs of vertical velocity. 310

311

3.1.2 Large-scale circulation 312

3.1.2.1 Westerly jet 313

Figure 6 shows the climatologically and zonally averaged zonal winds from the 314

UR NH-MPAS simulations with CAM4 and CAM5 and the UR H-MPAS-CAM4 315

simulations at 240 km and the difference between 240 km and other resolutions. The 316

general circulation pattern is similar to that reported in Rauscher et al. (2013). All 317

simulations produce low-level tropical easterlies equatorward of 23°N/S and a single core 318

for the westerly jet at ~200 hPa and ~30°N/S. From the H-MPAS-CAM4 simulations, 319

Rauscher et al. (2013) found that the near-surface westerlies display increasing speed 320

with increasing resolution while the upper level westerlies display decreasing speed with 321

increasing resolution. In this study, the maximum speed of the near-surface (900 hPa) 322

zonally averaged zonal wind varies very little with resolutions from 240 km to 30 km, 323

ranging between 13.1 and 13.5 m s-1 from NH-MPAS-CAM5 and between 13.9 and 14.3 324

m s-1 from NH-MPAS-CAM4, respectively. Similar small sensitivity of the near surface 325

maximum westerly speed is also found with H-MPAS-CAM4, ranging between 14.0 and 326

14.4 m s-1. Larger differences between 240 km and other resolutions in H-MPAS than 327

NH-MPAS are mainly due to the poleward shift of the westerlies in H-MPAS that will be 328

further discussed later. The more robust sensitivity across all the experiments with NH-329

12

MPAS or H-MPAS is that of the intensity of the upper level (250hPa) westerly jet to 330

resolution: the jet speed decreases from 56.1 to 54.3 m s-1 and from 52.6 to 50.0 m s-1 in 331

NH-MPAS-CAM5 and NH-MPAS-CAM4, respectively, as the resolution increases from 332

240 km to 30 km, while it weakens more significantly from 56.7 to 50.3 m s-1 in H-333

MPAS-CAM4 from 240 km to 60 km. The greater sensitivity found in H-MPAS-CAM4 334

is more comparable to that in Rauscher et al. (2013). 335

The most recurrent sensitivity about the westerly jet might be its poleward shift 336

with enhanced resolution as reported in many previous studies (Pope and Stratton, 2002; 337

Williamson, 2008a; Wan et al. 2013; Lu et al. 2015). The subtropical low level westerly 338

jet position was found to have a large influence on the frequency of atmospheric river 339

events, with lower atmosphere river frequency corresponding to a more poleward 340

placement of the low level westerly jet in H-MPAS-CAM4 (Hagos et al. 2015). Figure 7 341

shows the maximum near-surface (900 hPa) zonal mean zonal wind and its corresponding 342

locations (latitudes) from the NH-MPAS simulations with CAM4 and CAM5 physics and 343

the H-MPAS-CAM4 simulations at four different resolutions. Consistent with Figure 6, 344

NH-MPAS with both CAM5 and CAM4 show only small sensitivity in their near surface 345

jet position to resolution, while H-MPAS-CAM4 simulates systematic poleward shifts of 346

the jet position from 33°N to 37.75°N as the mesh size decreases from 240 km to 30 km, 347

with the most significant shift occurring between 240 km and 120 km. This poleward 348

shift of about 5° between 240 km and 30 km with H-MPAS-CAM4 is similar to that 349

reported by Lu et al. (2015) for the same model. The sources for the distinct sensitivity of 350

the NH-MPAS and H-MPAS models deserve further analysis. 351

To examine the above sensitivity, a finite-amplitude wave activity budget at 250 352

hPa based on the vorticity equation from Lu et al. (2015) is adopted here. This 353

diagnostics links the upper tropospheric eddy vorticity forcing for the eddy-driven jet to 354

the dissipation and the wave activity source. Readers are referred to Nakamura and Zhu 355

(2010) for the concept of finite-amplitude wave activity and to Lu et al. (2015) for the 356

theoretical framework for the wave activity budget. For brevity, only the eddy vorticity 357

flux, which directly maintains the eddy-driven component of the westerly jet, the wave 358

activity source, and the effective diffusivity are presented for NH-MPAS-CAM5 (Figure 359

8, upper panels) and NH-MPAS-CAM4 (Fig. 8 middle panels), in contrast with the 360

13

corresponding results for H-MPAS-CAM4 (Fig. 8, lower panels). First, one can see that 361

the peaks of the eddy vorticity flux correspond well with the near-surface jet positions 362

shown in Figure 7, indicating that the meridional structure of the latter is maintained by 363

that of the former. As such the eddy vorticity flux exhibits the same sensitivity to model 364

resolution as the jet position. The distinct sensitivities between NH-MPAS and H-MPAS 365

appear to arise from two factors: the different behaviors of the wave activity source, and 366

the different sensitivities of the mid-latitude effective diffusivity, which is a measure of 367

the efficiency of the dissipation of wave activity. 368

To investigate the role of each factor mentioned above, the middle column of 369

Figure 8 shows that in NH-MPAS (for both CAM4 and CAM5 physics), the peak of the 370

wave activity source moves slightly equatorward with increasing resolution, while the 371

opposite is true for H-MPAS-CAM4. Additionally, the NH-MPAS simulations show 372

much weaker sensitivity in the mid-latitude effective diffusivity (right column) to 373

resolution than the H-MPAS simulations. As discussed in Lu et al (2015), the smaller 374

mid-latitude effective diffusivity in higher resolution allows more wave activity to 375

survive the mid-latitude dissipation, giving rise to greater meridional propagation of wave 376

activity across the jet and larger eddy vorticity flux (as one can see from comparing the 377

bottom left and bottom right panels in Figure 8). This sensitivity of the effective 378

diffusivity is less acute in NH-MPAS as the resolution increases and the associated 379

mechanism as counterbalance to the weakening and equatorward shift of the wave source 380

is also much weakened. On the other hand, the wave source in H-MPAS-CAM4 tends to 381

shift poleward and the effective diffusivity decreases more drastically with resolution. As 382

a consequence, H-MPAS has a much greater sensitivity in the position of the eddy-driven 383

jet than the NH-MPAS. The greater sensitivity of the mid-latitude effective diffusivity in 384

H-MPAS than NH-MPAS may be explained by the differences in the hyperdiffusion 385

coefficients used in the corresponding dynamical cores (Rauscher et al., 2013; Park et al., 386

2016). Table 1 shows the hyperdiffusion coefficients among the experiments. The H-387

MPAS experiments used overall larger coefficients than the NH-MPAS experiments for 388

the same mesh spacing, and the corresponding ratio of the coefficients between H-MPAS 389

and NH-MPAS decreases with increasing resolution. Thus, from the lowest to the highest 390

resolution the hyperdiffusivity decreases with a much greater proportion in H-MPAS than 391

14

NH-MPAS, which may have contributed to the greater sensitivity of the mid-latitude 392

effective diffusivity in the former (Fig. 8). A sensitivity experiment with H-MPAS-393

CAM4 at 240 km with a smaller hyperdiffusion coefficient shows smaller difference in 394

the mid-latitude effective diffusivity between 240 km and 120 km (not shown), further 395

supporting the influence of the hyperdiffusion coefficient on the effective diffusivity and 396

wave activity simulated by H-MPAS-CAM4. 397

398

3.1.2.2 Hadley circulation 399

Previous studies using H-MPAS-CAM4 show some sensitivity of the Hadley 400

circulation strength to model resolution. Using the maximum mean meridional 401

streamfunction as a measure of the Hadley circulation strength (Oort and Yienger 1996; 402

Mitas and Clement 2005), H-MPAS-CAM4 simulates a stronger Hadley circulation with 403

indices ranging from 1.86 to 2.52 (×1011 kg s-1) with increasing resolution from 240 km 404

to 30 km, which is mainly due to the low value (1.86×1011 kg s-1) at 240 km. NH-MPAS 405

with CAM5 and CAM4 physics simulates smaller sensitivity of the Hadley circulation 406

strength to resolution with indices ranging between 2.63-2.70 (×1011 kg s-1) and 2.24-2.43 407

(×1011 kg s-1), respectively. CAM5 simulates a slightly stronger Hadley circulation than 408

CAM4. To better understand this resolution impact on the Hadley circulation, we 409

compute the Normalized Gross Moist Stability (NGMS; see Raymond et al. 2007, 2009). 410

NGMS relates the relative vertically integrated divergence of moist static energy (h, 411

formally moist entropy, but they are approximately equivalent) to the vertically integrated 412

divergence of water vapor (q). At steady state, this ratio is equivalent to the fluxes of 413

energy and moisture through the top of atmosphere and surface. 414

NGMS = −∇⋅h!v

Lv ∇⋅q!v = FS − RP − E

415

In the equation above, !v is the wind vector, vL is the latent heat, FS is the surface flux of 416

energy (sensible plus latent), R is the integrated radiative sink of energy from the 417

atmosphere, P is surface precipitation, E is surface evaporation, and • denotes a 418

pressure integral from the surface to the top of the atmosphere. Note that precipitation 419

and evaporation on the right hand side of the equation above are in units of W m-2 so that 420

15

NGMS remains dimensionless. While NGMS is commonly computed using the 421

circulation to make predictions for precipitation, here we compute it using the TOA and 422

surface fluxes to predict the circulation changes. If we neglect the role of eddies and 423

assume that the structure of h and q are relatively constant between simulations (as one 424

might expect for a fixed moist adiabatic temperature profile with fixed relative humidity), 425

then we expect a stronger circulation for a weaker NGMS owing to the fact that the 426

divergence of humidity term grows much faster than the divergence of energy with 427

circulation intensification. 428

We compute NGMS over the tropical rain band where P is larger than E (P-E > 0; 429

the upwelling region of the Hadley circulation). The change in NGMS from 240 km to 30 430

km resolution is -0.025, -0.009, -0.066 for NH-MPAS-CAM5, NH-MPAS-CAM4, and 431

H-MPAS-CAM4, respectively. From above, the changes in maximum streamfunction are 432

0.07, 0.19, and 0.66 (all x1011 kg/s) for NH-MPAS-CAM5, NH-MPAS-CAM4, and H-433

MPAS-CAM4, respectively. Both the difference in NGMS and the difference in 434

maximum streamfunction are more than doubled in the H-MPAS-CAM4 experiment 435

compared to NH-MPAS with CAM5 and CAM4 (see Figure 9). NGMS is influenced by 436

either changes in the water cycle (P-E) or in the energy fluxes (FS-R). We find here that 437

the energy fluxes are relatively constant with changing resolution, such that increases in 438

P-E (which are dominated by changes in precipitation; not shown) are the dominant 439

control on the decreases in NGMS, and hence the increases in circulation strength. Figure 440

9 shows that increases in P-E with resolution mimic the increases in streamfunction 441

strength and decreases in NGMS, while changes in FS-R do not. For NH-MPAS-CAM5, 442

NH-MPAS-CAM4, and H-MPAS-CAM4, the changes in P-E are 40, 24, and 106 W m-2, 443

respectively, while the changes in energy fluxes are only -2.1, 2.0, and -1.2 W m-2, 444

respectively. The different resolution impacts between NH-MPAS and H-MPAS are 445

primarily associated with the differences in relative humidity and resolved vertical 446

velocity profiles (not shown) that influence the cloud microphysical processes and 447

precipitation. Since the energy fluxes at the top of the atmosphere and surface are 448

insensitive to the resolution changes, the NGMS framework shows that Hadley 449

circulation strength increases in the models are coupled to the increases in precipitation. 450

451

16

3.2 Global variable resolution simulations 452

3.2.1 Precipitation 453

Figure 10 shows the spatial distribution of the difference in average total, 454

parameterized convective, and resolved large-scale precipitation between the UR MPAS 455

simulations at 30 km and 120 km mesh spacing and between the VR (30->120 km) and 456

UR (120 km) simulations with CAM4 and CAM5 physics. All the simulation results are 457

remapped to a one-degree grid. Since precipitation and its sensitivity to resolutions 458

beyond 60°N/S are subtle, only results between 60°N/S are shown. The values of tropical 459

precipitation averaged over the refined region (10°S-10°N, 30°W-30°E) from all 460

simulations are summarized in Table 2. As discussed above, all simulations at UR 30 km 461

produce more large-scale and less convective tropical precipitation than those at UR 120 462

km, with negligible (0.08 mm day-1) and small (0.43 mm day-1) difference in zonal mean 463

tropical total precipitation in NH-MPAS-CAM5 and NH-MPAS-CAM4, respectively. In 464

general, the VR simulations produce comparable resolution impacts on convective and 465

large-scale precipitation in the refined region (black circle) as displayed in the UR 120 466

km and 30 km simulations. The average convective (large-scale) precipitation in the 467

refined region of VR is reduced (enhanced) by -1.48 mm day-1 and -0.31 mm day-1 (1.49 468

mm day-1 and 0.63 mm day-1) in NH-MPAS CAM5 and CAM4, respectively, compared 469

to the corresponding UR 120 km simulations. These changes are comparable to those 470

between UR 30 km and UR 120 km. The sensitivity of convective and large-scale 471

precipitation to model resolution using CAM5 is larger than that using CAM4 as 472

discussed before. H-MPAS-CAM4 also simulates an evident reduction (enhancement) of 473

convective (large-scale) precipitation in the refined region from UR 120 km to VR 30-474

>120km. The averaged total precipitation in the refined region is insensitive (changing 475

within 5%) to the resolution change from 120 km to 30 km in all simulations. 476

Similar to the findings from Rauscher et al. (2013) and Hagos et al. (2013), H-477

MPAS-CAM4 simulates a zonally asymmetric distribution of resolution impacts on total 478

precipitation in the refined region. That is, despite the smaller changes in total 479

precipitation with resolution in the refined region compared to previous studies, there is 480

still an evident increase of precipitation on the western (downwind) side and decrease on 481

the eastern (upwind) side of the refined region. This asymmetric distribution is mainly 482

17

due to the asymmetric enhancement and reduction of the large-scale and convective 483

precipitation, respectively, associated with the advection of atmospheric state from the 484

upstream low-resolution region into the refined region. In comparison, NH-MPAS CAM5 485

and CAM4 simulate a zonally symmetric distribution of resolution impact, a small 486

increase (~3%) of total precipitation over both sides of the refined region. This results in 487

smaller diabatic heating and streamfunction anomalies compared to those found in 488

Rauscher et al. (2013) and Hagos et al. (2013) (not shown). The zonally asymmetric 489

precipitation sensitivity to resolutions and the associated circulation anomalies have been 490

attributed to the response of the moist physics to local mesh refinement, which is 491

confirmed by the Held-Suarez (Held and Suarez, 1994) test, in which model physics are 492

replaced by prescribed forcing and dissipation (Rauscher et al., 2013). 493

Of interest in the NH-MPAS-CAM4 VR simulation is that the same moist physics 494

(CAM4) coupled to different dynamical cores (NH-MPAS versus H-MPAS) produces a 495

zonally symmetric resolution impact inside the refined region. It is also noteworthy that 496

with a spectral element hydrostatic dynamical core coupled to the CAM4 and CAM5 497

physics and a pressure based vertical coordinate, Zarzycki et al. (2014) also found 498

zonally symmetric distributions of precipitation changes with resolutions in the refined 499

region. In their study as well as in ours, the moist physics appears to respond to model 500

resolution changes more rapidly across the transition zone. One reason for this may be 501

due to the lower sensitivity of NH-MPAS simulated integrated water vapor (IWV) to 502

resolution. The resolution impacts on IWV averaged over the refined region from NH-503

MPAS CAM5 and CAM4 (-0.50 and -1.09 mm, respectively) are smaller than that (-1.47 504

mm) from H-MPAS-CAM4. Averaged near the equator over the refined region (5°S-5°N, 505

30°W-30°E), a resolution change from 120 km to 30 km increases the near-surface 506

(below 700 hPa) vertical velocity slightly by about -0.002 Pa s-1 in the NH-MPAS CAM5 507

and CAM4 simulations, but the impact on the H-MPAS-CAM4 simulated vertical 508

velocity is even smaller. These suggest that the smaller sensitivity of IWV rather than 509

vertical velocity with resolution in the NH-MPAS simulations contribute to the zonally 510

symmetric distribution of precipitation changes in the refined region, which is in stark 511

contrast to the asymmetric distribution simulated in H-MPAS. 512

3.2.2 Clouds 513

18

The monotonic decrease of cloud fraction with increasing horizontal grid 514

resolution is a well-known issue with the CAM physics (e.g., Williamson et al., 2008a,b; 515

Rauscher et al., 2013; O’Brien et al., 2013; Zarzycki et al., 2014). The increased updraft 516

velocity with increasing resolution (Fig. 5) and mass continuity imply that masses are 517

transported upward through narrow columns, therefore reducing the fractional area 518

covered by clouds. O’Brien et al. (2013) and Zarzycki et al. (2014) found smaller 519

sensitivity of the simulated cloud fraction to resolutions in CAM5 compared to CAM4. 520

Figure 11 shows the spatial distributions of cloud fraction from the MPAS UR 30 km and 521

120 km simulations and the VR simulations. NH-MPAS with CAM4 and CAM5 both 522

display smaller cloud fraction in UR 30 km compared to UR 120 km. The cloud fraction 523

in the refined region of the VR simulations is also smaller than that of the low resolution 524

region and comparable to that of UR 30 km. Consistent with previous studies (O’Brien et 525

al., 2013; Zarzycki et al., 2014), CAM5 produces larger cloud fraction than CAM4, and 526

simulates less sensitivity of cloud fraction to resolution. Averaged over the refined region 527

at 30 km, NH-MPAS-CAM5 simulates cloud fraction (0.62) higher than that (0.28) from 528

NH-MPAS-CAM4. In the meantime, the 15% reduction of cloud fraction (0.73 to 0.62) 529

from 120 km to 30 km averaged over the refined region in NH-MPAS-CAM5 is less than 530

the 35% reduction (0.43 to 0.28) in NH-MPAS-CAM4. H-MPAS-CAM4 simulates a 531

similar reduction of cloud fraction as NH-MPAS-CAM4 but with a larger cloud fraction 532

value of 0.35. Similar to the precipitation response, H-MPAS-CAM4 simulates a zonally 533

asymmetric distribution of cloud fraction reduction between the eastern and western sides 534

of the refined region. As with precipitation, such behavior is not apparent in the NH-535

MPAS simulations with either CAM4 or CAM5 physics. 536

537

4. Summary and discussion 538

This study analyzed a series of aqua-planet simulations conducted using NH-539

MPAS with the CAM4 and CAM5 physics to document the distinct simulated features of 540

precipitation, clouds, and large-scale circulation compared to behaviors reported in 541

previous studies for H-MPAS-CAM4. Similar to previous studies (e.g., Rauscher et al., 542

2013), the quasi-uniform resolution experiments show large reductions in the ratio of 543

convective to large-scale precipitation with increasing model resolution related to the 544

19

sensitivity of large-scale condensation from the cloud microphysics to the resolved 545

vertical velocities. This ratio has larger sensitivity in NH-MPAS-CAM5 than NH-MPAS-546

CAM4. Compared to H-MPAS-CAM4, NH-MPAS-CAM4 generally simulates more 547

convective precipitation and less large-scale precipitation at all resolutions but their 548

sensitivities to resolution are comparable. Despite the apparent sensitivity of each 549

precipitation component to resolution, sensitivity of the total precipitation to resolution is 550

very small, particularly for NH-MPAS-CAM5. 551

As model resolution increases, all models produced higher frequencies of more 552

intense precipitation with no sign of convergence even when the simulations are all 553

interpolated to the same coarser resolution grid. CAM5 has a propensity to produce more 554

intense precipitation than CAM4. This is consistent with the changes of extreme upward 555

vertical velocity to resolution. Although the differences are small, NH-MPAS-CAM4 556

consistently produces slightly lower frequencies of intense precipitation than H-MPAS-557

CAM4, which may be related to the treatment of vertical velocity that is diagnosed in H-558

MPAS and predicted in NH-MPAS. Thus in H-MPAS, the positive feedback between 559

latent heating from large-scale condensation produced by cloud microphysics and the 560

vertical velocity may be stronger because the adjustment of the vertical velocity to the 561

latent heating is instantaneous based on diagnostic relationships. Note that the same 562

physics and dynamics time steps as well as moist physics parameters are used in all 563

experiments so the different model sensitivities can be attributed to differences in the 564

dynamical cores or physics parameterizations rather than time steps and convective time 565

scales, as discussed in Williamson et al. (2008a). 566

The NH-MPAS and H-MPAS simulations display significantly different 567

sensitivity of large-scale circulation features to model resolution. The most notable 568

difference is the simulated structure of the ITCZ: instead of a transition from single to 569

double ITCZ in previous simulations as model resolution increases, our simulations show 570

an unambiguous single ITCZ at all resolutions examined. With a well-balanced global 571

moisture budget, NH-MPAS with CAM4 or CAM5 physics also produced a single ITCZ 572

at all model resolutions ranging from 240 km to 30 km. These findings support the 573

mechanisms identified by Landu et al. (2014), where the feedback between convection 574

and large-scale circulation is differentiated by the influence of atmospheric moisture 575

20

content on CAPE (which is lower in the previous H-MPAS-CAM4 simulations) and the 576

parameterized convection in the simulations. 577

From analysis of the westerly jet and Hadley circulation, NH-MPAS also displays 578

much lower sensitivity to model resolution than H-MPAS. Lu et al. (2015) found that the 579

low level westerly jet shifts poleward with increasing resolution in H-MPAS-CAM4, 580

which also significantly affects the simulated frequency of atmospheric river events 581

(Hagos et al., 2015). The H-MPAS-CAM4 simulations in this study show similar results, 582

but the jet position simulated by NH-MPAS CAM5 and CAM4 is almost insensitive to 583

horizontal resolutions. The difference is mainly attributable to the poleward shift of the 584

peaks of eddy vorticity flux with increasing resolution in H-MPAS-CAM4, which is 585

forced by a poleward shift of the wave activity source as well as a large reduction in 586

effective diffusivity with increasing resolution, thus more wave activity survives the mid-587

latitude dissipation as the resolution increases. On the contrary, the NH-MPAS 588

simulations show an equatorward shift of the wave activity source and smaller reductions 589

in effective diffusivity with increasing resolution. The different resolution sensitivity 590

between NH-MPAS and H-MPAS is partly related to the hyperdiffusivity used in the two 591

dynamical cores. The sensitivity of the Hadley circulation strength to resolution 592

simulated by NH-MPAS is similar between CAM4 and CAM5, and is weaker than H-593

MPAS with CAM4. The diagnosis of NGMS shows that the increase in the hydrologic 594

cycle with increasing resolutions, apart from energy fluxes, dominates the increase in 595

Hadley circulation strength. The NGMS diagnostic confirms that the lower sensitivity of 596

the Hadley circulation to resolution in NH-MPAS compared to H-MPAS can be 597

attributed to the lower sensitivity of the tropical total precipitation in the former. 598

Lastly, we analyzed the VR simulations for the three models with different 599

dynamical cores and physics parameterizations. The averaged total precipitation over the 600

refined region is much less sensitive to the resolution change from 120 km to 30 km in all 601

experiments compared to previous studies. Therefore, the magnitude of anomalous 602

rotational flow at 200 hPa that is driven by the anomalous heating over the refined region 603

is also much weaker than previous studies (e.g., Rauscher et al., 2013; Hagos et al., 2013). 604

Despite the reduced sensitivity of precipitation to resolution, the simulated cloud fraction 605

still significantly decreases with increasing resolution in the UR simulations and over the 606

21

refined region in the VR simulations. Consistent with previous studies (O’Brien et al., 607

2013; Zarzycki et al., 2014), CAM5 produces larger cloud fraction and is less sensitive 608

than CAM4 to resolutions. H-MPAS-CAM4 in this study still simulates an asymmetric 609

distribution of resolution impacts on total precipitation over the refined region, as noted 610

in Rauscher et al. (2013) and Hagos et al. (2013). In contrast, NH-MPAS simulates a 611

symmetric distribution of resolution impact over the refined region, which is consistent 612

with what Zarzycki et al. (2014) found using a spectral element dynamical core and 613

CAM4 and CAM5 physics. We found lower sensitivity of IWV in NH-MPAS 614

transitioning from the low to the high resolution region than in H-MPAS. 615

Correspondingly, NH-MPAS-CAM4 also simulates a symmetric distribution of cloud 616

fraction in the refined region, in contrast to H-MPAS that simulates an asymmetric 617

distribution. 618

Overall, this study reveals a lower sensitivity in NH-MPAS to horizontal 619

resolution compared to H-MPAS in many climatic features. This provides benefits to the 620

VR configuration for producing more zonally symmetric features that are more 621

comparable to the UR simulations in an aqua-planet framework. The differences between 622

NH-MPAS and H-MPAS are related to differences in the dynamical cores, including the 623

non-hydrostatic versus hydrostatic formulations, differences in vertical coordinates and 624

vertical resolutions, differences in hyperdiffusion coefficients, and other features. The 625

differences between the CAM4 and CAM5 simulations may be partly a result of different 626

cloud microphysics and boundary layer schemes. Our analyses show that important 627

differences in humidity and precipitation can result from differences in the formulation of 628

the dynamical cores alone, which have notable effects on the tropical large-scale 629

circulation features including the ITCZ and Hadley circulation. Given the strong coupling 630

between dynamics and physics, more in-depth analyses and numerical experiments are 631

needed to disentangle the physical mechanisms governing any specific process presented 632

in this manuscript. The reduced resolution sensitivity of NH-MPAS demonstrated in this 633

study should motivate future research to further investigate the model behaviors and 634

explore its use in very high resolution and convection-permitting simulations using 635

regional refinement. 636

637

22

Acknowledgements 638

This research was supported by the Office of Science of the U.S. Department of 639

Energy (DOE) as part of the Regional & Global Climate Modeling (RGCM) program. 640

Dr. J.-H. Yoon was partially supported by funding from the Korean Polar Research 641

Institute. This study used computing resources from the National Energy Research 642

Scientific Computing Center, which is the DOE Office of Science User 643

Facility supported by the Office of Science of the U.S. Department of Energy under 644

Contract No. DE-AC02-05CH11231 and Contract DE-AC02-06CH11357, respectively. 645

Pacific Northwest National Laboratory is operated by Battelle Memorial Institute for the 646

DOE under contract DE-AC05-76RL01830. 647

648

23

Reference 649

Bacmeister, J. T., M. F. Wehner, R. B. Neale, A. Gettelman, C. Hannay, P. H. Lauritzen, 650

J. M. Caron, and J. E. Truesdale: Exploratory high-resolution climate simulations 651

using the Community Atmosphere Model (CAM). J. Climate, 27, 3073–3099, 652

doi:10.1175/JCLI-D-13-00387.1, 2014. 653

Du, Q., V. Faber, and M. Gunzburger, 1999: Centroidal Voronoi tessellations: 654

Applications and algorithms. SIAM Rev., 41 (4), 637–676. 655

Gettelman, A., H. Morrison, and S. J. Ghan: A new two-moment bulk stratiform cloud 656

microphysics scheme in the NCAR Community Atmosphere Model (CAM3), Part 657

II: Single-column and global results, J. Clim., 21 (15), 3660–3679, 2008. 658

Gill, A. E., 1980: Some simple solutions for heat-induced tropical circulation. Quart. J. 659

Roy. Meteor. Soc., 106, 447–462. 660

Giorgi, F., and L. O. Mearns (1991), Approaches to the simulation of regional climate 661

change: A review, Rev. Geophys., 29(2), 191–216, doi:10.1029/90RG02636. 662

Giorgi, F., and M. R. Marinucci, 1996: An investigation of the sensitivity of simulated 663

precipitation to the model resolution and its implications for climate studies. Mon. 664

Wea. Rev., 124, 148-166, doi:10.1175/1520-665

0493(1996)124<0148:AIOTSO>2.0.CO;2. 666

Hagos, S., R. Leung, S. A. Rauscher, and T. Ringler: Error characteristics of two grid 667

refinement approaches in aquaplanet simulations: MPAS-A and WRF. Mon. Wea. 668

Rev., 141, 3022–3036, doi:10.1175/MWR-D-12-00338.1, 2013. 669

Hagos, S., L. R. Leung, Q. Yang, C. Zhao, and J. Lu, 2015: Resolution and dynamical 670

core dependence of atmospheric river frequency in global model simulations. J. 671

Climate, 28, 2764-2776, doi:10.1175/JCLI-D-14-00567.1. 672

Held, I. M., and M. J. Suarez, 1994: A proposal for the intercomparison of the dynamical 673

cores of atmospheric general circulation models. Bull. Amer. Meteor. Soc., 75, 674

1825–1830. 675

IPCC (2013), Climate Change 2013: The Physical Science Basis, the contribution of 676

Working Group I to the Fifth Assessment Report of the Intergovernmental Panel 677

on Climate Change, T. F. Stocker, D. Qin, G.-K. Plattner, M. Tignor, S. K. Allen, 678

24

J. Boschung, A. Nauels, Y. Xia, V. Bex, and P.M. Midgley (Eds.), Cambridge 679

University Press, Cambridge, UK and New York, NY, USA. 680

Ju, L. T., T. Ringler, and M. Gunzburger, 2011: Voroni tessellations and their application 681

to climate and global modeling. Numerical Techniques for Global Atmospheric 682

Models, P. H. Lauritzen et al., Eds., Lecture Notes in Computational Science and 683

Engineering, Vol. 80, Springer, 313–342. 684

Landu, K., L. R. Leung, S. Hagos, V. Vinoj, S. Rauscher, T. Ringler, and M. Taylor, 685

2014: The dependence of ITCZ structure on model resolution and dynamical core 686

in aquaplanet simulations, J. Climate, 27, 2375–2385, doi:10.1175/JCLI-D-13-687

00269.1. 688

Leung, L. R., Mearns, L. O., Giorgi, F., and Wilby, R. L.: Regional Climate Research, 689

Bull. Amer. Meteorol. Soc. 84, 89–95, 2003. 690

Leung, L. R., T. Ringler, W. D. Collins, and M. Ashfaq: A hierarchical evaluation of 691

regional climate simulations. Eos, Trans. Amer. Geophys. Union, 94, 297–298, 692

2013. 693

Liu, X., Easter, R. C., Ghan, S. J., Zaveri, R., Rasch, P., Shi, X., Lamarque, J.-F., 694

Gettelman, A., Morrison, H., Vitt, F., Conley, A., Park, S., Neale, R., Hannay, C., 695

Ekman, A. M. L., Hess, P., Mahowald, N., Collins, W., Iacono, M. J., Bretherton, 696

C. S., Flanner, M. G., and Mitchell, D.: Toward a minimal representation of 697

aerosols in climate models: description and evaluation in the Community 698

Atmosphere Model CAM5, Geosci. Model Dev., 5, 709-739, doi:10.5194/gmd-5-699

709-2012, 2012. 700

Lu, J., G. Chen, L. R. Leung, D. A. Burrows, Q. Yang, K. Sakaguchi, and S. Hagos: 701

Toward the dynamical convergence on the jet stream in aquaplanet AGCMs, J. 702

Clim., 28(17), 6763–6782, doi:10.1175/JCLI-D-14-00761.1, 2015. 703

Medvigy, D., R. L. Walko, M. J. Otte, and R. Avissar, 2013: Simulated changes in 704

northwest U.S. climate in response to Amazon deforestation. J. Climate, 26, 705

9115–9136, doi:10.1175/JCLI-D-12-00775.1. 706

Mitas, C. M., and A. Clement (2005), Has the Hadley cell been strengthening in recent 707

decades? Geophys. Res. Lett., 32, L03809, doi:10.1029/2004GL021765. 708

25

Morrison, H., and A. Gettelman: A new two-moment bulk stratiform cloud microphysics 709

scheme in the NCAR Community Atmosphere Model (CAM3), Part I: 710

Description and numerical tests, J. Clim., 21 (15), 3642–3659, 2008. 711

Murakami H, Sugi M: Effect of model resolution on tropical cyclone climate projections. 712

Scientific Online Letters on the Atmosphere 6:73-76, doi: 10.2151/sola.2010-019, 713

2010. 714

Murakami H, Wang Y, Yoshimura H, Mizuta R, Sugi M, Shindo E, Adachi Y, Yukimoto 715

S, Hosaka M, Kusunoki S, Ose T, Kitoh A: Future Changes in Tropical Cyclone 716

Activity Projected by the New High-Resolution MRI-AGCM*. J Clim 25:3237-717

3260, doi: 10.1175/JCLI-D-11-00415.1, 2012. 718

Nakamura, N. and D. Zhu: Finite-amplitude wave activity and diffusive flux of potential 719

vorticity in eddy–mean flow interaction. J. Atmos. Sci., 67, 2701–2716, 720

doi:10.1175/2010JAS3432.1, 2010. 721

Neale, R. B., and B. J. Hoskins, 2000: A standard test for AGCMs including their 722

physical parameterizations: I: The proposal. Atmos. Sci. Lett., 1, 101–107, 723

doi:10.1006/asle.2000.0022. 724

Neale, R. B., and Coauthors, 2010: Description of the NCAR Community Atmosphere 725

Model (CAM 4.0). NCAR Tech. Note NCAR/ TN-4851STR 726

(http://www.cesm.ucar.edu/models/ccsm4.0/cam/docs/description/cam4_desc.pdf, 727

last access: May 26, 2016). 728

Neale, R. B., and Coauthors, 2012: Description of the NCAR Community Atmosphere 729

Model (CAM 5.0). NCAR Tech. Note NCAR/ TN-4851STR 730

(http://www.cesm.ucar.edu/models/cesm1.0/cam/docs/description/cam5_desc.pdf, 731

last access: May 26, 2016). 732

NRC (2012), National Research Council Board, A National Strategy for Advancing 733

Climate Modeling, The National Academies Press, Washington, DC (2012) 734

(http://www.nap.edu/catalog/13430/a-national-strategy-for-advancing-climate-735

modeling, last access: May 25, 2016). 736

O’Brien, T. A., F. Li, W. D. Collins, S. A. Rauscher, T. D. Ringler, M. Taylor, S. M. 737

Hagos, and L. R. Leung, 2013: Observed scaling in clouds and precipitation and 738

26

scale incognizance in regional to global atmospheric models. J. Climate, 26, 739

9313-9333, doi:10.1175/JCLI-D-13-00005.1. 740

Oort, A. H., and J. J. Yienger, 1996: Observed interannual variability in the Hadley 741

circulation and its connection to ENSO. J. Climate, 9, 2751–2767. 742

Park, S., and C. S. Bretherton: The university of washington shallow convection and 743

moist turbulence schemes and their impact on climate simulations with the 744

community atmosphere model, J. Climate, 22, 3449–3469, 2009 745

Park, S. H., W. C. Skamarock, J. B. Klemp, L. D. Fowler, and M. G. Duda, 2013: 746

Evaluation of global atmospheric solvers using extensions of the Jablonowski and 747

Williamson baroclinic wave test case. Mon. Wea. Rev., 141, 3116–3129, 748

doi:10.1175/MWR-D-12-00096.1. 749

Park, S., W. Skamarock, B. Medeiros, P. Lauritzen: Vertical resolution sensitivity on a 750

non-hydrostatic dynamical core with CAM physics for Aquaplanet experiments, 751

to be submitted to Journal of Advances in Modeling Earth Systems, 2016. 752

Pope, V. D., and R. A. Stratton, 2002: The processes governing horizontal resolution 753

sensitivity in a climate model. Climate Dyn., 19, 211–236, doi:10.1007/s00382-754

001-0222-8. 755

Rasch, P. J., and J. E. Kristj´ansson: A comparison of the CCM3 model climate using 756

diagnosed and predicted condensate parameterizations, J. Climate, 11, 1587–757

1614, 1998. 758

Rauscher, S. A., T. D. Ringler, W. C. Skamarock, and A. a. Mirin, 2013: Exploring a 759

global multiresolution modeling approach using aquaplanet simulations. J. Clim., 760

26, 2432–2452, doi:10.1175/JCLI-D-12-00154.1. 761

Rauscher, S. A., and T. D. Ringler, 2014: Impact of variable resolution meshes on mid-762

latitude baroclinic eddies using CAM-MPAS-A. Mon. Weather Rev., 763

140715124709009, doi:10.1175/MWR-D-13-00366.1. 764

Raymond, D. J. Sessions, S. L., and Fuchs, Ž., 2007, A theory for the spinup of tropical 765

depressions. Quarterly Journal of the Royal Meteorological Society, 1754, 1743—766

1754, doi: 10.1002/qj.125 767

27

Raymond, David J., Sessions, Sharon L., Sobel, Adam H., and Fuchs, Željka, 2009, The 768

mechanics of gross moist stability. Journal of Advances in Modeling Earth 769

Systems, 2, doi: 10.3894/JAMES.2009.1.9. 770

Ringler, T., L. Ju, and M. Gunzburger, 2008: A multiresolution method for climate 771

system modeling: Application of spherical centroidal Voronoi tessellations. 772

Ocean Dyn., 58, 475–498, doi:10.1007/s10236-008-0157-2. 773

Ringler, T., J. Thuburn, J. B. Klemp, and W. C. Skamarock, 2010: A unified approach to 774

energy conservation and potential vorticity dynamics for arbitrarily structured C-775

grids. J. Comput. Phys., 229, 3065–3090, doi:10.1016/j.jcp.2009.12.007. 776

Ringler, T., D. Jacobsen, M. Gunzberger, L. Ju, M. Duda, and W. Skamarock, 2011: 777

Exploring a multiresolution modeling approach within the shallow-water 778

equations. Mon. Wea. Rev., 139, 3348–3368. 779

Roberts MJ, Vidale PL, Mizielinski MS, Demory ME, Schiemann R, Strachan J, Hodges 780

K, Bell R, Camp J: Tropical cyclones in the upscale ensemble of high resolution 781

global climate models. J Climate, 28:574-596, doi: 10.1175/JCLI-D-14-00131.1, 782

2015. 783

Skamarock, W. C., J. B. Klemp, J. Dudhia, D. O. Gill, D. M. Barker, X.-Y. Huang, W. 784

Wang, and J. G. Powers, 2008: A description of the Advanced Research WRF 785

version 3. NCAR Tech. Note NCAR/TN-4751STR, 113 pp. 786

Skamarock, W. C., J. B. Klemp, L. D. Fowler, M. G. Duda, S.-H. Park, and T. D. 787

Ringler, 2012: A multiscale nonhydrostatic atmospheric model using centroidal 788

Voronoi tesselations and C-grid staggering. Mon. Wea. Rev., 140, 3090–3105, 789

doi:10.1175/MWR-D-11-00215.1. 790

Skamarock, W. C., S.-H. Park, J. B. Klemp, and C. Snyder: Atmospheric kinetic energy 791

spectra from global high-resolution nonhydrostatic simulations. J. Atmos. Sci., 792

71, 4369–4381, doi:10.1175/JAS-D-14-0114.1, 2014. 793

Sakaguchi K., L. R. Leung, C. Zhao, Q. Yang, J. Lu, S. M. Hagos, S. Rauscher, L. Dong, 794

T. Ringler, and P. H. Lauritzen: Exploring a Multi-resolution Approach Using 795

AMIP Simulations. Journal of Climate, 28(14), 5549-5574, doi:10.1175/JCLI-D-796

14-00729.1, 2015. 797

28

Thuburn, J., T. D. Ringler, W. C. Skamarock, and J. B. Klemp: Numerical representation 798

of geostrophic modes on arbitrarily structured C-grids. J. Comput. Phys., 228, 799

8321–8335, 2009. 800

Wan, H., M. A. Giorgetta, and L. Bonaventura: Ensemble Held–Suarez test with a 801

spectral transform model: Variability, sensitivity, and convergence. Mon. Wea. 802

Rev., 136, 1075–1092, doi:10.1175/2007MWR2044.1, 2008. 803

Wang, Y., L. R. Leung, J. L. McGregor, D. K. Lee, W. C. Wang, Y. Ding, and F. 804

Kimura: Regional Climate Modeling: Progress, Challenges, and Prospects. 805

Journal of the Meteorological Society of Japan, 82(6):1599-1628, 2004. 806

Wehner MF, Reed KA, Li F, Prabhat, Bacmeister JT, Chen CT, Paciorek C, Gleckler PJ, 807

Sperber KR, Collins WD, Gettelman A, Jablonowski C: The effect of horizontal 808

resolution on simulation quality in the Community Atmospheric Model, CAM5.1. 809

J Adv Model Earth Syst 6(4):980-997, doi:10.1002/2013MS000276, 2014. 810

Wehner MF, Prabhat, Reed KA, Stone D, Collins WD, Bacmeister JT: Resolution 811

dependence of future tropical cyclone projections of cam5.1 in the US CLIVAR 812

Hurricane Working Group idealized group idealized configurations. J Clim 813

28:3905-3925, doi: 10.1175/JCLI-D-14-00311.1, 2015. 814

Williamson, D. L., J. Drake, J. Hack, R. Jakob, and P. Swarztrauber: A standard test set 815

for numerical approximations to the shallow water equations in spherical 816

geometry. J. Comput. Phys., 102, 211–224, 1992. 817

Williamson, D. L.: Convergence of aqua-planet simulations with increasing resolution in 818

the Community Atmospheric Model, version 3. Tellus, 60A, 848–862, 2008a. 819

Williamson, D. L.: Equivalent finite volume and Eulerian spectral transform horizontal 820

resolutions established from aqua-planet simulations. Tellus, 60A, 839–847, 821

2008b. 822

Yang, Q., L. R. Leung, S. A. Rauscher, T. D. Ringler, and M. A. Taylor, 2014: 823

Atmospheric moisture budget and spatial resolution dependence of precipitation 824

extremes in aquaplanet simulations, J. Climate, 27, 3565–3581, 825

doi:10.1175/JCLI-D-13-00468.1. 826

29

Zarzycki, C. M., M. N. Levy, C. Jablonowski, J. R. Overfelt, M. A. Taylor, and P. A. 827

Ullrich, 2014: Aquaplanet experiments using CAM’s variable-resolution 828

dynamical core. J. Climate, 27, 5481–5503, doi:10.1175/JCLI-D-14-00004.1. 829

Zhang, G. J., and N. A. McFarlane: Sensitivity of climate simulations to the 830

parameterization of cumulus convection in the Canadian Climate Centre general 831

circulation model, Atmosphere-Ocean, 33, 407–446, 1995. 832

Zhang, M., W. Lin, C. S. Bretherton, J. J. Hack, and P. J. Rasch: A modified formulation 833

of fractional stratiform condensation rate in the NCAR community atmospheric 834

model CAM2, J. Geophys. Res., 108 (D1), 2003. 835

836

837

838

30

839

840

841

842

843 Table 1. The numbers of grid cells and hyperdiffusion coefficients (m4/s) in the H-MPAS 844 and NH-MPAS experiments. The last column shows the ratio of hyperdiffusion 845 coefficients between the H-MPAS and NH-MPAS experiments. 846 847 Resolution

# of grid cells

H-MPAS NH-MPAS Ratio (H/NH MPAS)

240 km 10242 5×1015 6.91×1014 7.24

120 km 40962 5×1014 8.64×1013 5.79

60 km 163842 5×1013 1.08×1013 4.63

30 km 655362 5×1012 1.35×1012 3.70

30->120 km

102402 5×1012 to 5×1014 scaled

1.35×1012 to 8.64×1014

scaled 3.70 to 5.79 scaled

848 849 850 Table 2. Tropical precipitation averaged over the refined region (10°S-10°N, 30°W-851 30°E) for NH-MPAS and H-MPAS at 30 km, 120 km, and 30->120 km. 852 853

Model Total (mm day-1) Convective (mm day-1) Large-scale (mm day-1) 30 km

120km 30->120k

m

30km 120km

30->120km

30km

120km

30->120km

NH-MPAS-CAM5 9.40 9.48 9.51 4.46 5.93 4.45 4.94 3.56 5.05

NH-MPAS-CAM4 8.69 8.26 8.58 4.96 5.12 4.81 3.73 3.14 3.77

H-MPAS-CAM4 8.92 8.98 8.64 3.35 4.55 3.15 5.57 4.43 5.48

854 855 856 857 858 859 860

31

861

862

863

864

(a) (b) 865

866

867

868 869

Figure 1. The (a) quasi-uniform mesh and (b) variable-resolution mesh used in the UR 870

and VR experiments. Both meshes are plotted at resolutions significantly lower than used 871

in the experiments to show the mesh cells. (c) Mesh cell size as a function of distance 872

from the refinement center (0°N, 0°E) in VR grid. 873

874

875

876

32

877

878

879

880

881

882

883 Figure 2. Zonal mean precipitation rates from the NH-MPAS simulations with CAM4 884

and CAM5 and the H-MPAS-CAM4 simulations at four different global quasi-uniform 885

resolutions of approximately 30, 60, 120, and 240 km. The results from simulations at 886

various resolutions are all regridded to the 2°×2° finite volume grid. Note that the result 887

of NH-MPAS-CAM5 at 120 km mostly overlaps with the 60 km result. 888

889

890

891

892

893

894

895

896

897

898

33

899

900

901

902

903

904 Figure 3. Mean tropical (10°S-10°N) total, parameterized convective, and resolved large-905

scale precipitation simulated at different resolutions by H-MPAS-CAM4 and NH-MPAS 906

with CAM4 and CAM5 physics. 907

908

909

910

911

912

913

914

915

916

917

918

919

920

34

921

922

923

924

925

926 Figure 4. Probability density functions (PDFs) of hourly tropical (10°S-10°N) 927

precipitation for (a) lower intensity events (0-150 mm day-1, 1 mm day-1 bins) and (b) 928

higher intensity events (0-1000 mm day-1, 10 mm day-1 bins). 929

930

931

932

933

934

935

936

937

938

939

940

941

942

943

944

945

35

946

947

948

949

950 Figure 5. Probability density functions (PDFs) of 6-hourly tropical (10°S-10°N) upward 951

vertical velocity averaged below 700 hPa. 952

953

954

955

956

957

958

959

960

961

962

963

964

965

966

36

967

968

969 Figure 6. Vertical cross section of zonally averaged zonal winds from the UR NH-MPAS 970

simulations with CAM5 and CAM4 physics and the H-MPAS-CAM4 simulations at 240 971

km, and the difference between 240 km and three different resolutions of approximately 972

30, 60, 120 km. 973

974

975

976

977

978

979

980

981

37

982

983

984

985 Figure 7. Maximum zonal monthly mean westerly wind speed averaged between 950 and 986

800 hPa and the corresponding latitude from the UR NH-MPAS simulations with CAM4 987

and CAM5 physics and the UR H-MPAS-CAM4 simulations at resolutions of 30, 60, 988

120, and 240 km. 989

990

991

992

993

994

995

996

997

998

999

1000

1001

38

1002

1003

1004

1005 Figure 8. Eddy vorticity flux, wave activity source, and effective diffusivity from the 1006

wave activity budget analysis of Lu et al. (2015) for the NH-MPAS and H-MPAS 1007

simulations at various global quasi-uniform resolutions. 1008

1009

1010

1011

1012

1013

1014

1015

1016

1017

1018

1019

20 40 60 80-5

0

5

10

NH-CAM

5

× 10 -5eddy vorticity flux (ms- 2 )

240km120km60km30km

20 40 60 80-5

0

5

10

15

20 × 10 -5wave activity source (ms- 2 )

20 40 60 8002468

1012 × 10 6

effective diffusivity (m2s- 1 )

20 40 60 80-5

0

5

10

NH-CAM

4

× 10 -5

20 40 60 80-5

0

5

10

15

20 × 10 -5

20 40 60 8002468

1012 × 10 6

20 40 60 80latitude

-5

0

5

10

H-CA

M4

× 10 -5

20 40 60 80latitude

-5

0

5

10

15

20 × 10 -5

20 40 60 80latitude

02468

1012 × 10 6

39

1020

1021

1022

1023

1024

1025 Figure 9. Differences between experiments with 30 km horizontal resolution and 240 km 1026

horizontal resolution for: (a) maximum streamfunction; (b) NGMS; and (c) FS-R (left 1027

side) and P-E (right side). The changes in NGMS, FS-R, and P-E are all averaged over the 1028

tropical rain band (defined here as the region where P-E>0). Note that we have flipped 1029

the signs for ΔNGMS since increasing resolution decreases NGMS in all experiments. 1030

1031

1032

1033

1034

1035

1036

1037

1038

1039

1040

1041

1042

1043

40

1044

1045 Figure 10. Spatial distribution of the difference in averaged total, parameterized 1046

convective, and resolved large-scale precipitation between the MPAS simulations at UR 1047

30 km (UR30) and 120 km (UR120) resolutions and between the VR 30->120 km (VR) 1048

and UR120 resolutions with CAM4 and CAM5 physics. The results are shown between 1049

60°N/S. The black solid line circle denotes the refined region with a radius of 30°. The 1050

area between the black solid circle and black dashed circle denotes the transition zone 1051

between the refined region and the boundary of coarse resolution with a radius of 60°. 1052

1053

1054

1055

1056

41

1057

1058

1059

1060 Figure 11. Spatial distributions of cloud fraction from the MPAS UR simulations at 30 1061

km and 120 km and the VR simulations at 30->120 km in different experiments. The 1062

results are shown between 90°N/S. The black solid circle denotes the refined region with 1063

a radius of 30°. The black dashed circle denotes the boundary of the coarse resolution 1064

region with a radius of 60°. 1065

1066

1067