GEOPHYSICAL RESEARCH LETTERS, VOL. ???, XXXX, DOI:10.1029/, 1 2 What can moist thermodynamics tell us about circulation shifts in response to uniform warming? 1 2 Tiffany A. Shaw , Aiko Voigt Corresponding author: T. A. Shaw, Department of the Geophysical Sciences, 5734 S. Ellis Ave., Chicago, IL, 60637 USA. ([email protected]) 1 Department of the Geophysical Sciences, The University of Chicago, Chicago, IL, USA. 2 Lamont-Doherty Earth Observatory, Columbia University, New York, NY, USA D R A F T April 27, 2016, 6:44am D R A F T X-2 3 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION Aquaplanet simulations exhibit a robust expansion of the Hadley cell and 4 poleward jet shift in response to uniform warming of sea-surface tempera- 5 ture. Here moist thermodynamic and dynamic frameworks are combined to 6 make predictions of circulation responses to warming. We show Clausius-Clapeyron 7 (CC) scaling of specific humidity with warming predicts an expansion of the 8 Hadley circulation according to convective quasi-equilibrium dynamics. A 9 poleward jet shift follows from the control-climate relationship between the 10 Hadley cell edge and jet stream position. CC scaling of specific humidity with 11 warming also predicts decreased diffusivity and a poleward shift of the lat- 12 itude of maximum latent and dry-static energy transport according to mixing- 13 length theory. Finally, atmospheric cloud-radiative changes shift the latitude 14 of maximum energy transport poleward in most models. Our results show 15 moist thermodynamics can predict meridional shifts of the circulation when 16 combined with dynamical frameworks, however additional feedbacks are im- 17 portant for the simulated response. D R A F T April 27, 2016, 6:44am D R A F T SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION X-3 1. Introduction 18 Confidence in the projected thermodynamic and dynamic response of the climate sys- 19 tem to anthropogenic climate change requires a physical understanding of the underlying 20 mechanisms. It is well understood that the thermodynamic response to increased CO2 21 involves warming, increased near-surface specific humidity over the ocean following the 22 Clausius-Clapeyron (CC) relation, warming of the tropical-upper troposphere and a rais- 23 ing of the tropopause [e.g. Held , 1993; Allen and Ingram, 2002; Held and Soden, 2006; 24 Vallis et al., 2015]. Held and Soden [2006] used CC scaling to predict the “wet-get-wetter, 25 dry-get-drier” response to warming. The dynamic response to increased CO2 , e.g. the 26 response of the Hadley cell and jet stream, is less well understood [Shepherd , 2014; Vallis 27 et al., 2015]. However, the majority of coupled-climate models, atmosphere only mod- 28 els and aquaplanet models participating in the Coupled Model Intercomparison Project 29 Phase 5 (CMIP5) project a robust (in terms of model agreement) poleward expansion 30 of the circulation, including a poleward shift of the edge of the Hadley cell and zonal- 31 mean jet stream position, in response to anthropogenic climate change [Held , 1993; Vallis 32 et al., 2015; Medeiros et al., 2015]. While models may not agree on the magnitude of 33 the response, the meridional direction is robust and motivates the search for a physical 34 explanation. 35 Given our understanding of the moist thermodynamic response to global warming, we 36 ask the question: What can moist thermodynamics tell us about the circulation response 37 to uniform warming? Previous authors have linked moist thermodynamics, in particular 38 warming of the tropical upper troposphere and increased tropopause height, to Hadley D R A F T April 27, 2016, 6:44am D R A F T X-4 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 39 circulation expansion a posteriori [Held , 2000; Lu et al., 2007; Vallis et al., 2015]. Building 40 on those results here we put forward physical arguments linking moist thermodynamic 41 changes to dynamic responses using different theoretical frameworks. We focus on two dif- 42 ferent moist thermodynamic responses to warming. The first is the increase of near-surface 43 specific humidity with warming following CC scaling. We investigate the implication of 44 this increase for near-surface moist entropy and moist-static energy changes, which are 45 linked to dynamical responses following convective quasi-equilibrium dynamics [Emanuel 46 et al., 1994; Emanuel , 1995] and mixing-length theory [Held , 2000; Kushner and Held , 47 1998]. In both cases we show increased specific humidity following CC scaling can be used 48 to predict meridional shifts of the circulation in response to uniform warming. 49 The second response to warming we consider is changes in atmospheric cloud-radiative 50 effects (ACREs). ACRE changes can result from both thermodynamic as well as dynamic 51 changes [Stevens and Bony, 2013; Voigt and Shaw , 2015]. An important cloud response 52 to warming is the rise of tropical high clouds that is expected from the fixed-anvil temper- 53 ature hypothesis [Hartmann and Larson, 2002; Kuang and Hartmann, 2007; Zelinka and 54 Hartmann, 2010]. Here we consider ACRE changes as a thermodynamic response in the 55 sense that they change the energy balance of the atmosphere and demand a change in the 56 meridional energy transport [Hwang et al., 2011]. The ACRE-induced energy transport 57 can impact the latitude of maximum energy transport. The latitude of maximum energy 58 transport occurs in the extratropics where storm tracks dominate the energy transport. 59 We investigate the connection between moist thermodynamic and dynamic responses 60 to uniform warming using prescribed sea surface temperature (SST) aquaplanet models. D R A F T April 27, 2016, 6:44am D R A F T SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION X-5 61 Aquaplanet models are an important part of the model hierarchy and provide an idealized 62 setting (comprehensive moist physics with simplified lower boundary condition) for un- 63 derstanding the circulation response to global warming [Blackburn et al., 2013; Medeiros 64 et al., 2015]. CMIP5 aquaplanet model simulations exhibit a robust poleward expansion 65 of the tropics and poleward shift of the eddy-driven jet stream in response to uniform 66 warming [Medeiros et al., 2015] consistent with the response in coupled-climate models 67 with realistic boundary conditions [Vallis et al., 2015]. 2. Data 68 We make use of the aquaControl and aqua4K CMIP5 prescribed SST aquaplanet model 69 experiments [Taylor et al., 2012]. Most models prescribed the aquaControl SST follow- 70 ing the QOBS distribution of Neale and Hoskins [2001], however the FGOALS-g2 and 71 MIROC5 models used the CTRL distribution [Medeiros et al., 2015]. Aqua4K mimics 72 global warming by uniformly increasing the aquaControl SST by 4K. We use monthly 73 and zonally-averaged zonal wind (ua), temperature (ta), specific humidity (hus), radia- 74 tion (rlut, rlutcs, rlus, rlds, rldscs, rsut, rsutcs, rsus, rsds, rsdscs) and surface flux (hfls, 75 hfss) data from ten CMIP5 models (CCSM4, CNRM-CM5, FGOALS-g2, FGOALS-s2, 76 HadGEM2-A, IPSL-CM5A-LR, MIROC5, MPI-ESM-LR, MPI-ESM-MR, MRI-CGCM3). 77 The data are interpolated onto a common latitude grid with 0.1◦ resolution. Since the 78 aquaplanet climate is equatorially symmetric, data from the Northern and Southern hemi- 79 spheres are averaged together. We quantify inter-model spread as ± 1 standard deviation. 80 Maximum values are obtained by a quadratic fit using the maximum grid point and two 81 adjacent points. D R A F T April 27, 2016, 6:44am D R A F T X-6 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 3. Results 82 We begin by documenting shifts of the Hadley cell edge and eddy-driven jet stream 83 position in response to uniform SST warming by 4K. We define the edge of the Hadley 84 cell as the latitude where the zonal-mean zonal wind transitions from tropical easterlies 85 to extratropical westerlies at 925 hPa, i.e., the latitude of zero zonal wind. The eddy- 86 driven jet stream position is defined as the latitude of maximum zonal-mean zonal wind 87 at 925 hPa. The Hadley cell edge and jet stream position in aquaControl are significantly 88 correlated (r = 0.96, p = 0.00) and a linear regression accounts for 92% of the model 89 spread (Fig. 1a, solid black line). A similar relationship occurs for aqua4K (r = 0.95, 90 p = 0.00, r2 = 0.90, Fig. 1a, dashed black line). 91 All ten CMIP5 aquaplanet models exhibit a poleward shift of the Hadley cell edge (Fig. 92 1b, y-axis) and jet stream position (Fig. 1c, y-axis) in response to uniform warming. Our 93 results are consistent with alternative metrics of the Hadley cell edge, e.g. the latitude 94 where the mean-meridional streamfunction vertically integrated between 700 and 300 hPa 95 equals zero [see Fig. 5 of Medeiros et al., 2015] and jet stream position, e.g. the latitude 96 where zonal-mean zonal wind averaged between 850 and 700 hPa is maximum [see Fig. 6 97 of Medeiros et al., 2015]. 98 In the following subsections we provide examples of how moist thermodynamic changes 99 can be used to infer circulation shifts in response to uniform warming. We focus on two 100 thermodynamic changes in response to uniform warming: 1) increased specific humidity 101 following CC scaling and 2) ACRE changes. D R A F T April 27, 2016, 6:44am D R A F T SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION X-7 3.1. Circulation changes predicted from CC scaling 102 103 Following Held and Soden [2006] we assume changes in near-surface specific humidity q exhibit CC scaling in response to warming, i.e., δq ≈ αδT q (1) 104 where α = L/RT 2 ≈ 0.07K−1 , L is the latent heat of vaporization, R is the dry gas 105 constant, T is temperature, and δT = 4K. The change in specific humidity affects two 106 important thermodynamic variables: near-surface or sub-cloud (925 hPa) moist entropy, 107 sb = cp ln θe ≈ cp ln(θ + Lq/cp ), and moist-static energy, ms = cp T + Lq + Φ, where cp is 108 the specific heat at constant pressure, θe and θ are the near-surface equivalent potential 109 temperature and potential temperature, respectively, and Φ is geopotential. 110 111 Following CC scaling, the near-surface moist entropy and moist-static energy responses to uniform warming are δsb ≈ cp δT (cp + αLq) δθe ≈ θe θe (2) δms ≈ cp δT + Lδq ≈ δT (cp + αLq). 112 (3) These responses are largest in the tropics implying an increased meridional gradient, i.e., 1 ∂δsb 1 ∂ q L 1 ∂q L q ∂T ≈ αδT L ≈ αδT ≈ α2 δT a ∂φ a ∂φ θe θe a ∂φ θe a ∂φ 1 ∂δms 1 ∂q q ∂T ≈ αδT L ≈ α2 δT L . a ∂φ a ∂φ a ∂φ (4) (5) 113 where a is the radius of the Earth and we have assumed constant relative humidity with 114 latitude in the last step, which is a reasonable approximation for aquaControl (Supple- 115 mentary Fig. 1a). We neglect −Lq∂θe /∂φ/θe2 in (4) because specific humidity dominates 116 the equivalent potential temperature gradient in low latitudes (Supplementary Fig. 1b). D R A F T April 27, 2016, 6:44am D R A F T X-8 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 117 These thermodynamic responses to uniform warming can be connected to dynamical re- 118 sponses via two frameworks: 1) quasi-equilibrium dynamics and 2) mixing-length theory. 119 In both cases we make predictions of circulation shifts in response to uniform warming 120 and compare them to the simulated response (aqua4K-aquaControl). 121 Quasi-equilibrium connects thermally-direct circulations to the region of supercritical 122 near-surface moist entropy and constrains the vertical structure of the atmosphere to be 123 moist adiabatic [Emanuel et al., 1994]. The moist adiabatic assumption is reasonable for 124 Earth’s tropics and subtropics [Korty and Schneider , 2007] and quasi-equilibrium is a 125 reasonable approximation over the tropical oceans [Brown and Bretherton, 2007] and in 126 Monsoon regions [Boos and Emanuel , 2009; Nie et al., 2010]. 127 According to quasi-equilibrium dynamics, the balanced component of the surface zonal 128 wind us is related to the degree of supercriticality of the near-surface moist entropy sb , 129 i.e., us = (Ts − Tt ) ∂ (sb − scrit ) fa ∂φ (6) 130 where scrit is the critical value of moist entropy, Ts and Tt are the temperatures at the 131 surface and tropopause, respectively, and f is the Coriolis parameter [see equation(18) 132 in Emanuel , 1995]. The tropopause is defined following the World Meteorological Orga- 133 nization’s definition. The critical value of moist entropy is connected to the onset of a 134 thermally direct circulation. Here we estimate the critical value following equation (11) 135 in Emanuel [1995], i.e., Ω2 a2 (cos2 φm − cos2 φ)2 = cp ln(θe,crit ) = cp ln θem exp − 2cp (Ts − Tt ) cos2 φ ( scrit D R A F T April 27, 2016, 6:44am !) (7) D R A F T X-9 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 136 where θem is the maximum equivalent potential temperature located at latitude φm and 137 Ω is the rotation rate. According to quasi-equilibrium dynamics the multi-model mean 138 edge of the Hadley cell is located where sb − scrit is minimum, i.e., at 22.5◦ , as compared 139 to its actual location of 24.6◦ in aquaControl. 140 The surface zonal wind response to uniform warming can be decomposed as δus = δ(Ts − Tt ) ∂ (Ts − Tt ) ∂ (Ts − Tt ) ∂ (sb − scrit ) + δsb − δscrit fa ∂φ fa ∂φ fa ∂φ (8) 141 [see (6)]. We focus on the predicted zonal-wind response due to changes in moist entropy 142 following CC scaling [see (4)], which dominate in the tropics and subtropics (Supplemen- 143 tary Fig. 1c), i.e., δus ≈ (Ts − Tt ) L ∂q (Ts − Tt ) 2 Lq ∂T (Ts − Tt ) ∂ δsb ≈ αδT ≈ α δT < 0. fa ∂φ fa θe ∂φ fa θe ∂φ (9) 144 This predicted decrease in zonal wind enhances tropical easterlies and shifts the Hadley 145 cell edge poleward. 146 The predicted shift of the Hadley cell edge in response to uniform warming following 147 CC scaling, quasi-equilibrium dynamics and using only information from aquaControl, is 148 poleward when the predicted zonal-wind response is added to the aquaControl zonal wind 149 (3.2◦ ±0.6◦ , Fig. 1b). The predicted poleward shift overestimates the simulated (aqua4K- 150 aquaControl) zonal wind response and Hadley cell edge shift (1.6◦ ±0.3◦ , Fig. 1b). The 151 predicted response is not correlated with the simulated response across the models. 152 Given the predicted Hadley cell edge for the warmed climate, we predict the jet stream 153 position for the warmed climate using the linear relationship between Hadley cell edge and 154 jet position in aquaControl (Fig. 1a, top left). Consistent with the linear relationship, 155 a poleward shift of the edge of the Hadley cell following CC scaling leads to a poleward D R A F T April 27, 2016, 6:44am D R A F T X - 10 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 156 shift of jet stream position (3.8◦ ±0.7◦ , Fig. 1c). Once again the predicted poleward shift 157 generally overestimates the aqua4K-aquaControl response (2.8◦ ±0.7◦ , Fig. 1c) and the 158 predicted response is not correlated with the simulated response across the models. 159 Changes in specific humidity following CC scaling also affect the near-surface moist- 160 static energy gradient, which is connected to the kinematic diffusivity D and vertically- 161 integrated meridional transport of moist-static energy hvmi via mixing-length theory, 162 i.e., hvmi ≈ −(D/a)∂ms /∂φ where h·i represents a vertical integral between the surface 163 and 10 hPa weighted by the acceleration due to gravity. The diffusivity in aquaControl, 164 i.e., D ≡ −(1/a)hvmi/(∂ms /∂φ), involves midlatitude (38.7◦ ) and high-latitude (65.5◦ ) 165 maxima (Fig. 2a, vertical black lines). The diffusivity maxima are related to latent and 166 dry-static energy transport, respectively (Supplementary Fig. 2a). 167 168 Following mixing-length theory, changes in moist-static energy transport can be decomposed as δhvmi ≈ − δD ∂ms D ∂δms − . a ∂φ a ∂φ (10) 169 In contrast to previous studies that examined heat transport changes assuming fixed 170 diffusivity, i.e., δD ≈ 0, and hence circulation [e.g. Frierson et al., 2007; Kang et al., 171 2009], here we are interested in diffusivity changes as a proxy for the circulation response 172 to uniform warming. Assuming δhvmi ≈ 0, i.e., compensation occurs between dry-static 173 and latent energy transport [see Supplementary Fig. 2b, Frierson et al., 2007; Manabe et 174 al., 1965, 1975; Held and Soden, 2006; Hwang et al., 2011], and following CC scaling [see 175 (5)] we obtain δD ≈ − ∂δms /∂φ ∂q/∂φ ∂T /∂φ D ≈ −αδT L D ≈ −α2 δT Lq D < 0. ∂ms /∂φ ∂ms /∂φ ∂ms /∂φ D R A F T April 27, 2016, 6:44am (11) D R A F T X - 11 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 176 177 178 Thus, CC scaling predicts diffusivity will decrease in response to uniform warming. The predicted latent and dry-static energy transport responses to uniform warming following CC scaling and mixing-length theory are: L ∂q L ∂T δhLvqi ≈ − (δD + αδT D) ≈ − (δD + αδT D)αq , a ∂φ a ∂φ δD ∂T δD ∂s ≈ −cp , δhvsi ≈ − a ∂φ a ∂φ (12) (13) 179 which satisfy δhvmi ≈ 0 assuming (11). We note that a diffusive latent energy transport 180 prediction is not accurate for the deep tropics, however we are interested in changes 181 in the latitude of maximum transport, which occur in midlatitudes where the diffusive 182 approximation is valid. We label the terms in (12)-(13) proportional to δD as dynamic 183 transport responses and those proportional to δT as thermodynamic transport responses. 184 Note there is no predicted thermodynamic dry-static energy transport response to uniform 185 warming (there is no temperature gradient change). We cannot make predictions assuming 186 compensation and fixed diffusivity in response to uniform warming because δhvmi = δD = 187 0 → ∂δms /∂φ = 0 [see (10)]. 188 The predicted diffusivity decrease in response to uniform warming following CC scal- 189 ing is largest in high latitudes (Fig. 2a, red line). However, the largest fractional de- 190 crease occurs in the subtropics and high latitudes (Fig. 2b, red line). The predicted 191 subtropical decrease leads to a poleward shift of the latitude of maximum midlatitude 192 diffusivity (0.4◦ ±0.2◦ , Fig. 2c). However, the predicted shift underestimates the aqua4K- 193 aquaControl response (3.5◦ ±0.8◦ , Fig. 2d) and the two responses are not significantly cor- 194 related. The underestimate occurs because the aqua4K-aquaControl diffusivity increases 195 in midlatitudes (Fig. 2a,b, black lines) enhancing the poleward shift. The predicted diffu- D R A F T April 27, 2016, 6:44am D R A F T X - 12 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 196 sivity decrease in high latitudes leads to an equatorward shift of maximum high-latitude 197 diffusivity (-0.4◦ ±0.3◦ , Fig. 2c). This prediction is not accurate: the simulated high- 198 latitude diffusivity shifts poleward in most models (1.8◦ ±1.7◦ , Fig. 2d). The discrepancy 199 occurs because the aqua4K-aquaControl diffusivity does not change in high-latitudes (Fig. 200 2a,b, black line) instead the largest decrease occurs further south, which leads to the pole- 201 ward shift. We discuss the discrepancy between the predicted and simulated high-latitude 202 diffusivity response below. 203 The predicted dynamic latent-energy transport decreases in response to uniform warm- 204 ing (Fig. 3a, solid red line) and shifts the latitude of maximum latent energy transport 205 poleward in most models (0.2◦ ±0.2◦ , Fig. 3c), which agrees with the aqua4K-aquaControl 206 response (0.8◦ ±0.5◦ , Fig. 3c). However, this is partly offset by increased thermodynamic 207 transport (Fig. 3a, dashed red line). The predicted dynamic dry-static energy transport 208 decreases in response to uniform warming (Fig. 3b, solid red line) and shifts the latitude 209 of maximum dry-static energy transport poleward (0.6◦ ±0.2◦ , Fig. 3d). However, the 210 aqua4K-aquaControl dynamic dry-static energy transport response increases in midlati- 211 tudes (Fig. 3b, black solid line) and shifts the latitude of maximum transport equatorward 212 (-2.2◦ ±0.7◦ , Fig. 3d). The aqua4K-aquaControl thermodynamic dry-static energy trans- 213 port response (Fig. 3b, dashed black line), which is not considered in the prediction [see 214 (13)], leads to a poleward shift of the latitude of maximum dry-static energy transport. 215 The discrepancy between predicted and simulated diffusivity (Fig. 2a,b red versus black 216 lines) and energy transport (Fig. 3a,b, red versus black lines) responses suggests a missing 217 component of the mixing-length theory prediction. The difference between the simulated D R A F T April 27, 2016, 6:44am D R A F T SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION X - 13 218 and predicted diffusivity response (Supplementary Fig. 3a, black line) is due to changes 219 in moist-static energy transport (assumed to be zero in the prediction) and deviations of 220 the moist-static energy gradient response from CC scaling in response to uniform warm- 221 ing. The simulated increase of moist-static energy transport in response to warming leads 222 to increased diffusivity, particularly in high latitudes (Supplementary Fig. 3a, red line). 223 Deviations of the moist-static energy gradient response from CC scaling leads to increased 224 mid-latitude diffusivity (Supplementary Fig. 3a, blue line), which dominates the discrep- 225 ancy between the simulated and predicted dynamic latent and dry static energy transport 226 responses (Supplementary Fig. 3b, solid lines). The discrepancy between the simulated 227 and predicted thermodynamic latent and dry static energy transport responses is domi- 228 nated by deviations of the moist-static energy gradient response from CC scaling. The 229 change in dry static energy gradient (dominated by temperature) and the change in la- 230 tent energy gradient both produce meridional dipoles of energy transport (Supplementary 231 Fig. 3b, dashed lines). Meridional dipoles of latent energy transport occur in response to 232 internal eddy-zonal flow feedbacks [Boer et al., 2001; Lorenz and Hartmann, 2001, 2003; 233 Thompson and Woodworth, 2014], thus these feedbacks might be a missing component of 234 the predictions. 3.2. Circulation changes inferred from ACRE response 235 The second response to warming we consider is ACRE changes. ACRE is defined as 236 the difference of CRE at the top of the atmosphere and surface, including both longwave 237 sf c toa and shortwave components, i.e., ACRE = (F toa − Fclr ) − (F sf c − Fclr ) where F is the 238 total radiative flux, Fclr is the clear-sky radiative flux, sf c refers to the surface, and toa D R A F T April 27, 2016, 6:44am D R A F T X - 14 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 239 refers to the top of the atmosphere. The radiative fluxes are defined as positive in the 240 downward direction. Here we assess how ACRE changes in response to uniform warming 241 affect moist-static energy transport, in particular its impact on the latitude of maximum 242 transport. In contrast to the previous subsection where we made predictions following 243 CC scaling solely based on information from aquaControl, the predictions made here use 244 information from both aquaControl and aqua4K, i.e., the ACRE response to uniform 245 warming is taken as given. 246 The ACRE response to uniform warming generally increases in the tropics and robustly 247 decreases in high latitudes (Fig. 4a). (We removed the global average but that does not 248 affect our conclusions.) Longwave radiation dominates the ACRE response (Supplemen- 249 tary Fig. 4a). We note that ACRE changes are an indirect measure of the cloud-radiative 250 response. We confirmed the ACRE changes for the MPI-ESM-LR and IPSL-CM5A-LR 251 models agree with the cloud-radiative response diagnosed from forward partial radiative 252 perturbation calculations following Weatherald and Manabe [1988] (see Supplementary 253 Fig. 4b). 254 255 The ACRE changes can be converted into a meridional moist-static energy transport response following Hwang and Frierson [2011], i.e., δFACRE = "Z φ −π/2 Z 2π 2 a cos φ(δACRE)dλdφ + 0 Z π/2 Z 2π φ # 2 a cos φ(δACRE)dλdφ /2. (14) 0 256 The ACRE changes increase poleward energy transport in both hemispheres (Fig. 4b). 257 When the ACRE-induced energy transport response is added to moist-static energy trans- 258 port in aquaControl, there is a poleward shift of the latitude of maximum energy transport 259 in all but one model (0.63◦ ± 0.40◦ , Fig. 4c). D R A F T April 27, 2016, 6:44am D R A F T SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION X - 15 260 The ACRE-induced poleward shift of the latitude of maximum energy transport is 261 consistent with the poleward shift for aqua4K-aquaControl (0.53◦ ±0.31◦ , Fig. 4c). How- 262 ever, the predicted transport shift is smaller than the predicted jet shift (Fig. 1c) most 263 likely because ACRE changes are small in midlatitudes. Previous work has shown height- 264 dependent midlatitude cloud-radiative changes in response to uniform warming, which do 265 not project onto vertically-integrated ACRE, shift the jet poleward. 4. Conclusion and Discussion 266 A complete understanding of the response to anthropogenic climate change must involve 267 an understanding of the linkages between thermodynamic and dynamic responses. Here, 268 using physical arguments and aquaplanet simulations, we linked moist thermodynamic 269 and dynamic responses to uniform warming. We focused on CMIP5 aquaplanet simu- 270 lations because they exhibit robust poleward circulation shifts in responses to uniform 271 warming [Medeiros et al., 2015] and therefore provide an idealized setting for understand- 272 ing the circulation response in coupled-climate model simulations with realistic boundary 273 conditions. Our results build upon previous work that connected thermodynamic and 274 dynamic responses to global warming a posteriori [e.g. Lu et al., 2007; Vallis et al., 2015] 275 and used dry dynamical core simulations to investigate circulation responses [e.g. Butler 276 et al., 2010; Mbengue and Schneider , 2013; Lu et al., 2014]. 277 We focused on two moist thermodynamic responses to uniform warming and what they 278 tell us about meridional shifts of the circulation. The first is increased specific humidity 279 with warming following CC scaling. According to quasi-equilibrium dynamics, the in- 280 creased meridional gradient of sub-cloud moist entropy in response to uniform warming D R A F T April 27, 2016, 6:44am D R A F T X - 16 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 281 following CC scaling amplifies tropical easterlies and shifts the edge of the Hadley cell, 282 defined here as the latitude of zero near-surface zonal-mean zonal wind, poleward. The 283 control-climate relationship between the Hadley cell edge and jet position predicts the jet 284 will shift poleward following CC scaling. While quasi-equilibrium dynamics is an approx- 285 imate framework, we have shown it predicts a robust direction for the circulation shift 286 in response to uniform warming (e.g., it predicts a poleward not an equatorward shift) 287 and provides an explanation for the circulation shift (related to increased moist-entropy 288 gradients following CC scaling). 289 According to mixing-length theory, the increased meridional gradient of moist-static 290 energy in response to uniform warming following CC scaling decreases diffusivity assum- 291 ing moist-static energy transport changes are small. The diffusivity decrease is largest is 292 the subtropics and high latitudes and shifts the latitude of maximum midlatitude diffu- 293 sivity (related to moisture transport) poleward and latitude of maximum high latitude 294 diffusivity (related to dry-static energy transport) equatorward. The diffusivity changes 295 can be connected to dynamic energy transport responses that lead to poleward shifts of 296 the latitude of maximum latent and dry-static energy transport in response to uniform 297 warming. 298 Discrepancies exist between the simulated and mixing-length theory predicted re- 299 sponses, in particular, simulated diffusivity actually increases in midlatitudes and does 300 not change significantly in high latitudes in response to uniform warming. The increased 301 simulated diffusivity is due to increased moist-static energy transport (e.g., due to ACRE 302 changes), which is not accounted for in the prediction, and to deviations of the moist-static D R A F T April 27, 2016, 6:44am D R A F T SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION X - 17 303 energy gradient response from CC scaling in response to uniform warming. The differ- 304 ences between the simulated and predicted thermodynamic transport responses involve 305 meridional dipoles that reflect changes in temperature and moisture gradients. Meridional 306 dipoles of energy transport have been shown to occur in response to eddy-zonal flow feed- 307 backs [Boer et al., 2001; Lorenz and Hartmann, 2001, 2003; Thompson and Woodworth, 308 2014]. These feedbacks, which are not included in the prediction, might be important for 309 determining the full magnitude of the circulation response to warming. 310 The second moist thermodynamic response to uniform warming examined here is 311 changes in ACRE and its impact on moist-static energy transport. ACRE mostly in- 312 creases in the tropics and robustly decreases in high latitudes in response to uniform 313 warming. The inferred meridional moist-static energy transport response is poleward in 314 most models and maximizes poleward of the latitude of maximum energy transport in the 315 control climate. Consequently, the latitude of maximum energy transport shifts poleward 316 in all but one CMIP5 aquaplanet model. Our results support previous work that has 317 shown the cloud radiative response to warming induces a poleward jet shift [Voigt and 318 Shaw , 2015; Ceppi and Hartmann, 2016] and contributes to model spread in moist-static 319 energy transport responses [Hwang et al., 2011]. 320 The moist thermodynamic-dynamic links examined here suggest circulation shifts in 321 response to warming may be fundamentally related to moisture gradients in the current 322 climate and ultimately to temperature gradients assuming fixed relative humidity [e.g. 323 eqns. (9), (11), (12), (13)]. They occur even when the warming perturbation is spatially 324 uniform. The aquaplanet responses to uniform warming are correlated with meridional D R A F T April 27, 2016, 6:44am D R A F T X - 18 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 325 temperature gradients in the control climate (Supplementary Fig. 5), highlighting a po- 326 tential emergent constraint on the circulation response to uniform warming. Additional 327 research is needed to assess the connection across an ensemble of coupled-climate models 328 with realistic boundary conditions to determine if a robust constraint emerges. 329 The circulation shifts predicted from the moist thermodynamic changes and dynamical 330 frameworks considered here are approximate. While they correctly predict the direction 331 of multi-model-mean meridional shift of the circulation, they could not account for the 332 magnitude of the simulated responses. Furthermore, the predicted shifts were not sig- 333 nificantly correlated with the simulated shifts in most cases suggesting thermodynamic 334 changes alone are not sufficient to predict model spread. An assessment across a larger 335 range of models is needed. The results suggest that thermodynamic [Voigt and Shaw , 336 2015; Ceppi and Hartmann, 2016] and dynamic feedbacks [Lu et al., 2014] are likely im- 337 portant for determining the simulated circulation responses to warming. Nevertheless 338 our results show moist thermodynamics can be combined with physical arguments and 339 dynamical frameworks to make predictions of circulation shifts in response to uniform 340 warming. 341 Acknowledgments. AV and TAS are supported by the David and Lucile Packard 342 Foundation. TAS is also supported by NSF award AGS-1255208 and the Alfred P. Sloan 343 Foundation. We thank the reviewers whose comments helped to improve the manuscript. 344 The model data in this study are available from TAS upon request. We thank N. Hender- 345 son and H. Liu for help downloading the CMIP5 data. We acknowledge the World Cli- 346 mate Research Programme’s Working Group on Coupled Modelling, which is responsible D R A F T April 27, 2016, 6:44am D R A F T SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION X - 19 347 for CMIP, and we thank the climate modeling groups for producing and making available 348 their model output. For CMIP the U.S. Department of Energy’s Program for Climate 349 Model Diagnosis and Intercomparison provides coordinating support and led development 350 of software infrastructure in partnership with the Global Organization for Earth System 351 Science Portals. References 352 353 354 355 356 357 358 359 Allan, R. P., (2011), Comparing satellite data and models to estimate cloud radiative effects at the surface and in the atmosphere, Meterol. Appl., 18, 324-333. Allen, M. R. and W. J. Ingram, (2002), Constraints on future changes in climate and the hydrological cycle, Nature, 419, 224-232. Blackburn, M. and B. J. Hoskins, (2013), Context and Aims of the Aqua-Planet Experiment, J. Met. Soc. Japan, 91A, 1-15. Boer, G. J., S. Fourest, and B. Yu, (2001), The Signature of the Annular Modes in the Moisture Budget, J. Climate, 14, 3655-3665. 360 Boos, W. R. and K. A. Emanuel, (2009), Annual intensification of the Somali jet in a 361 quasi-equilibrium framework: Observational composites, Q. J. Roy. Met. Soc., 135, 362 319-335. 363 364 Brown, R. G. and C. Bretherton, (2007), A test of the strict quasi-equilibrium theory on long time and space scales, J. Atmos. Sci., 54, 624-638. 365 Butler, A. H., D. W. J. Thompson, and R. Heikes (2010), The steady-state atmospheric 366 circulation response to climate change-like thermal forcings in a simple general circula- 367 tion model, J. Climate, 23, 3474-3496. D R A F T April 27, 2016, 6:44am D R A F T X - 20 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 368 Ceppi, P. and D. L. Hartmann (2015), Connections between clouds, radiation, and midlat- 369 itude dynamics: a review, Current Climate Change Reports, 3474-349610.1007/s4061- 370 015-0010-x. 371 372 373 374 375 376 Ceppi, P. and D. L. Hartmann (2016), Clouds and the atmospheric circulation response to warming. J. Climate, 29, 783-799,10.1175/JCLI-D-15-0394.1. Emanuel, K. A., J. D. Neelin and C. S. Bretherton (1994), On large-scale circulations in convecting atmospheres, Q. J. Roy. Met. Soc., 120, 1111-1143. Emanuel, K. A. (1995), On thermally direct circulation in moist atmospheres, J. Atmos. Sci., 52, 1529-1534. 377 Frierson, D. M. W., I. M. Held, and P. Zurita-Gotor (2007), A gray-radiation aquaplanet 378 moist GCM. Part II: Energy transports in altered climates, J. Atmos. Sci., 64, 1680- 379 1693. 380 381 Hartmann, D. L. and K. Larson (2002), An important constraint on tropical cloud-climate feedback, Geophys. Res. Lett., 10.1029/2002GL015835. 382 Held, I. M. (2000), The general circulation of the atmosphere, Woods Hole Pro- 383 gram in Geophysical Fluid Dynamics, http://www.gfdl.noaa.gov/cms-filesystem- ac- 384 tion/userfiles/ih/lectures/woodshole.pdf. 385 386 387 388 Held, I. M. (1993), Large-scale dynamics and global warming, Bull. Amer. Met. Soc., 74, 228-241. Held, I. M. and B. J. Soden (2006), Robust Responses of the Hydrological Cycle to Global Warming, J. of Climate, 19, 5686-5699. D R A F T April 27, 2016, 6:44am D R A F T X - 21 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 389 390 391 392 Hwang, Y.-T., D. M. W. Frierson, Held, I. M. and B. J. Soden (2011), Corrigendum to Held and Soden (2006), J. of Climate, 24, 1559-1560. Hwang, Y.-T., and D. M. W. Frierson (2011), Increasing atmospheric poleward energy transport with global warming, Geophys. Res. Lett., 37, 10.1029/2010GL045440. 393 Kang, S. M., D. M. W. Frierson, and I. M. Held, (2009), The Tropical Response to 394 Extratropical Thermal Forcing in an Idealized GCM: The Importance of Radiative 395 Feedbacks and Convective Parameterization, J. Atmos. Sci., 66, 2827-2812. 396 397 398 399 400 401 402 403 404 405 406 407 Kuang, Z. and D. L. Hartmann, (2007), Testing the fixed anvil temperature hypothesis in a cloud-resolving model, J. Climate, 20, 2051-2057. Kushner, P. J. and I. M. Held (1998), A test using atmospheric data of a method for estimating oceanic eddy diffusivity, Geophys. Res. Lett., 22, 4213-4216. Korty, R. L. and T. Schneider, (2007), A Climatology of the Tropospheric Thermal Stratification Using Saturation Potential Vorticity, J. Climate, 20, 5977-5991. Lorenz, D. J. and D. L. Hartmann (2001), Eddy-Zonal Flow Feedback in the Southern Hemisphere, J. Atmos. Sci., 58, 3312-3327. Lorenz, D. J. and D. L. Hartmann (2003), Eddy-Zonal Flow Feedback in the Northern Hemisphere Winter, J. Climate, 16, 1212-1227. Lu, J., G. A. Vecchi and T. Reichler, (2007), Expansion of the Hadley cell under global warming, Geophys. Res. Lett., 34, 10.1029/2006GL028443. 408 Lu, J., L. Sun, Y. Wu and G. Chen, (2014), The role of subtropical irreversible PV 409 mixing in the zonal-mean circulation response to global warming-like thermal forcing, 410 J. Climate, 27, 2297-2316. D R A F T April 27, 2016, 6:44am D R A F T X - 22 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION Manabe, S., J. Smagorinsky and R. F. Strickler (1965), Simulated climatology of a general circulation model with a hydrologic cycle, Mon. Wea. Rev., 93, 769-798. Manabe, S., K. Bryan, and M. J. Spelman, 1975: A global ocean-atmosphere climate model. Part I: The atmospheric circulation, J. Phys. Oceanogr., 5, 3-19. Mbengue, C. O. and T. Schneider (2013), Storm track shifts under climate change: what can be learned from large-scale dry dynamics, J. Climate, 26, 9923-9930. Medeiros, B., B. Stevens, and S. Bony, (2015), Using aquaplanets to understand the robust responses of comprehensive climate models to forcing, Clim. Dyn., 7, 1957-1977. Neale, R. and B. J. Hoskins, (2001), A standard test for AGCMs including their physical parametrizations: I: The proposal, Atmos. Sci. Lett., 10.1006/asle.2000.0019. Nie, J., W. R. Boos, and Z. Kuang, (2001), Observational Evaluation of a Convective Quasi-Equilibrium View of Monsoons, J. Climate, 23, 4416-4428. Shepherd, T. G., (2014), Atmospheric circulation as a source of uncertainty in climate change projections, Nature Geosc., 7, 703-708. Stevens, B. and S. Bony, (2013), What are climate models missing?, Science, 340, 10.1126/science.1237554. Taylor, K. E., R. J. Stouffer, and G. A. Meehl, (2012), An overview of CMIP5 and the experiment design, B. Amer. Met. Soc., 10.1175/BAMS-D-11-00094.1. Thompson, D. W. J. and J. D. Woodworth (2014), Barotropic and Baroclinic Annular Variability in the Southern Hemisphere, J. Atmos. Sci., 71, 1480-1493. 431 Vallis, G. K., P. Zurita-Gotor, C. Cairns, and J. Kidston (2015), Response of the large- 432 scale structure of the atmosphere to global warming, Q. J. Roy. Met. Soc., 141, 1479- D R A F T April 27, 2016, 6:44am D R A F T X - 23 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION 433 434 435 436 437 438 439 1501. Voigt, A., and T. A. Shaw, 2015: Circulation response to warming shaped by radiative changes of clouds and water vapour. Nature Geosc.,10.1038/NGEO2345. Wetherald, R. T., and S. Manabe (1988), Cloud Feedback Processes in a General Circulation Model. J. Atmos. Sci., 45, 1397-1416. Zelinka, M. and D. L. Hartmann, (2010), Why is longwave cloud feedback positive?, J. Geophys. Res., 115, D16117, doi:10.1029/2010JD013817. D R A F T April 27, 2016, 6:44am D R A F T X - 24 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION (a) latitude jet (deg.) 48 (r,p) = (0.96,0.00) y = 1.16x + 9.00, r2 = 0.92 44 Multi−model mean CCSM4 CNRM−CM5 FGOALS−g2 FGOALS−s2 HadGEM2−A IPSL−CM5A−LR MIROC5 MPI−ESM−LR MPI−ESM−MR MRI−CGCM3 40 36 32 20 22 24 26 28 latitude u=0 (deg.) 30 32 (b) Simulated ∆latitude u=0 (deg.) 2.5 (r,p) = (0.27,0.42) Multi−model mean CCSM4 CNRM−CM5 FGOALS−g2 FGOALS−s2 HadGEM2−A IPSL−CM5A−LR MIROC5 MPI−ESM−LR MPI−ESM−MR MRI−CGCM3 2 1.5 1 1.5 2 2.5 3 3.5 Predicted ∆latitude u=0 (deg.) 4 4.5 (c) Simulated ∆latitude jet (deg.) 4 (r,p) = (0.44,0.18) Multi−model mean CCSM4 CNRM−CM5 FGOALS−g2 FGOALS−s2 HadGEM2−A IPSL−CM5A−LR MIROC5 MPI−ESM−LR MPI−ESM−MR MRI−CGCM3 3 2 1 1 Figure 1. 2 3 4 Predicted ∆latitude jet (deg.) 5 6 (a) Latitude of eddy-driven jet stream versus latitude of the Hadley cell edge in aquaControl (circles) and aqua4K (squares). The aquaControl correlation (r = 0.96, p = 0.00), and linear regression (r2 = 0.92, black line, regression coefficients top left) are statistically significant. Similar results are obtained for aqua4K (r = 0.95, p = 0.00, r2 = 0.89, dashed line). (b) Simulated shift (aqua4K-aquaControl) of the Hadley cell edge versus the predicted shift in response to uniform warming. The correlation and p-value are in the top-left corner. (c) Same as (b) but for the jet shift predicted using the linear relationship in Fig. 1a. D R A F T April 27, 2016, 6:44am D R A F T X - 25 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION (a) (b) 10 0 0 δD/D (%) δD (x109 kgs−1) 2 −2 −10 −4 −20 −6 10 20 30 40 latitude 50 60 (c) 20 30 40 latitude 50 60 70 (d) 5 (r,p) = (−0.05,0.88) 4 Multi−model mean CCSM4 CNRM−CM5 FGOALS−g2 FGOALS−s2 HadGEM2−A IPSL−CM5A−LR MIROC5 MPI−ESM−LR MPI−ESM−MR MRI−CGCM3 3 2 0 0.5 Predicted ∆latitude max. D (deg.) 1 Simulated ∆latitude max. D (deg.) Simulated ∆latitude max. D (deg.) 10 70 6 (r,p) = (0.33,0.31) 4 Multi−model mean CCSM4 CNRM−CM5 FGOALS−g2 FGOALS−s2 HadGEM2−A IPSL−CM5A−LR MIROC5 MPI−ESM−LR MPI−ESM−MR MRI−CGCM3 2 0 −2 −1.5 −1 −0.5 0 Predicted ∆latitude max. D (deg.) 0.5 Figure 2. (a) Predicted (red) and simulated (aqua4K-aquaControl, black) diffusivity response to uniform warming. (b) Same as in (a) but for percent change. The vertical black lines indicate locations of diffusivity maxima in aquaControl. (c) Simulated shift (aqua4K-aquaControl) of the latitude of maximum midlatitude diffusivity versus the predicted shift in response to uniform warming. The correlation and p-value are in the top-left corner. (d) Same as (c) but for the latitude of maximum high-latitude diffusivity. D R A F T April 27, 2016, 6:44am D R A F T X - 26 SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION (b) (a) 0.2 0.1 0.15 0.05 PW PW 0.1 0.05 0 0 −0.05 −0.05 −0.1 10 20 30 40 latitude 50 60 70 20 30 40 latitude 50 60 70 80 2 (r,p) = (−0.52,0.09) Multi−model mean CCSM4 CNRM−CM5 FGOALS−g2 FGOALS−s2 HadGEM2−A IPSL−CM5A−LR MIROC5 MPI−ESM−LR MPI−ESM−MR MRI−CGCM3 1 0 −1 −0.5 Figure 3. 0 Predicted ∆latitude max. vq (deg.) 0.5 Simulated ∆latitude max. vs (deg.) (d) (c) Simulated ∆latitude max. vq (deg.) −0.1 10 80 (r,p) = (−0.54,0.08) −1 Multi−model mean CCSM4 CNRM−CM5 FGOALS−g2 FGOALS−s2 HadGEM2−A IPSL−CM5A−LR MIROC5 MPI−ESM−LR MPI−ESM−MR MRI−CGCM3 −2 −3 −4 0 0.5 1 Predicted ∆latitude max. vs (deg.) 1.5 Predicted (red) and simulated (aqua4K-aquaControl, black) dynamic (solid) and thermodynamic (dashed) (a) latent and (b) dry-static energy transport responses to uniform warming following mixing-length theory. Note there is no predicted change of thermodynamic dry static energy transport in response to uniform warming (no dashed red line in Fig. 3b). The vertical black lines indicate the position of maximum transport in aquaControl. (c) Simulated shift (aqua4K-aquaControl) of the latitude of maximum latent energy transport [see (12)] versus the predicted shift in response to uniform warming. The correlation and p-value are in the top-left corner. (d) Same as (c) but for the shift of the latitude of maximum dry-static energy transport [see (13)]. D R A F T April 27, 2016, 6:44am D R A F T SHAW & VOIGT: MOIST THERMODYNAMICS AND CIRCULATION X - 27 (a) 30 20 Multi−model mean CCSM4 CNRM−CM5 FGOALS−g2 FGOALS−s2 HadGEM2−A IPSL−CM5A−LR MIROC5 MPI−ESM−LR MPI−ESM−MR MRI−CGCM3 Wm−2 10 0 −10 −20 −80 −60 −40 −20 0 latitude 20 40 60 80 (b) 0.15 0.1 Multi−model mean CCSM4 CNRM−CM5 FGOALS−g2 FGOALS−s2 HadGEM2−A IPSL−CM5A−LR MIROC5 MPI−ESM−LR MPI−ESM−MR MRI−CGCM3 PW 0.05 0 −0.05 −0.1 −0.15 −80 −60 −40 −20 0 latitude 20 40 60 80 Simulated ∆latitude max. vm (deg.) (c) 1.5 (r,p) = (0.39,0.23) 1 Multi−model mean CCSM4 CNRM−CM5 FGOALS−g2 FGOALS−s2 HadGEM2−A IPSL−CM5A−LR MIROC5 MPI−ESM−LR MPI−ESM−MR MRI−CGCM3 0.5 0 −0.5 −0.5 Figure 4. 0 0.5 1 1.5 Predicted ∆latitude max. vm (deg.) 2 (a) ACRE response to uniform warming (aqua4K-aquaControl). (b) Moist-static energy transport response due to ACRE changes [see (14)]. The vertical black line indicates the position of maximum moist-static energy transport in aquaControl. (c) Simulated shift (aqua4K-aquaControl) of the latitude of maximum energy transport from the ACRE-induced energy transport response versus the predicted shift. The correlation and p-value are in the top-left corner. D R A F T April 27, 2016, 6:44am D R A F T
© Copyright 2025 Paperzz