Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | . 2

Open Access Discussions Chemistry and Physics Atmospheric O) have been increasing since pre- 2 ), relative to a uniform distribution which 4 25388 25387 ) and nitrous oxide (N 4 ering latitudinal distribution, highlighting the need for a careful evaluation of ff ), methane (CH 2 This discussion paper is/has beenand under Physics review (ACP). for Please the refer journal to Atmospheric the Chemistry corresponding final paper in ACP if available. Divecha Centre for Climate Change & Centre for Atmospheric and Oceanic Sciences, Indian tended to counteract the(Keith, 2000). undesired consequences of anthropogenic climate change Atmospheric concentrations of theide greenhouse (CO gases (GHGs)industrial such periods as primarily carbon2007). diox- because Their of increase fossil hasing fuel the the planetary potential use radiation to budget. andseveral To cause moderate geoengineering land-use future long proposals climate change term change haveneering (IPCC, and climate been is its change impacts, made an by recently. intentional alter- By large-scale definition, manipulation of geoengi- the environment, particularly in- by the global meanbut radiative with forcing, di are possibleSRM for proposals. a fixed total amount of aerosols 1 Introduction tal mass of sulfatemitigates aerosols changes (12.6 in Mt global of meanwhen SO aerosol temperature, concentration global is mean maximum radiativeclimate at forcing and the is consequently poles an larger leading intensified toulated hydrological a cycle. when warmer Opposite aerosol global changes mean concentration are1 sim- is K in maximized global in mean thebution temperature pattern: tropics. and this We 3 range % obtain is inHence, a about precipitation our range 50 study changes % demonstrates of by of that varying the a the range climate distri- of change global from mean a climate doubling states, of determined CO Solar radiation management (SRM)tial geoengineering option to has counteract been climatesimulations change. proposed to We as understand perform the a a global set poten- distribution of hydrological of implications idealized geoengineering solar of varying insolation the reduction latitudinal in SRM methods. We find that for a fixed to- Abstract Atmos. Chem. Phys. Discuss., 13, 25387–25415,www.atmos-chem-phys-discuss.net/13/25387/2013/ 2013 doi:10.5194/acpd-13-25387-2013 © Author(s) 2013. CC Attribution 3.0 License. Institute of Science, Bangalore – 560 012, India Received: 15 July 2013 – Accepted: 14Correspondence September to: 2013 A. – Modak Published: ([email protected]) 1 OctoberPublished 2013 by Copernicus Publications on behalf of the European Geosciences Union. Sensitivity of simulated climate to latitudinal distribution of solar insolation reduction in SRM geoengineering methods A. Modak and G. Bala 5 20 25 15 10 Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | ects ff climate 2 from the atmo- 2 longitude and the ◦ and geoengineering 2 ), doubled CO 2 CO ering latitudinal distribution × CO ff ects of varying the latitudinal × ff ects of space-based SRM meth- ects of varying the latitudinal dis- ff ff latitude and 2.5 ◦ 25390 25389 is 390 ppm and is 780 ppm in 2 2 CO × in 1 2 erent levels of uniform SRM forcing on regional climate response. Ban-Weiss ) and ten geoengineering simulations each with di ff 2 CO × We performed two sets of simulations: (1) fixed-SST (sea surface temperature) sim- We caution that our simulations are highly idealized and they are not meant to rep- Past climate modeling studies have modeled the e Proposed geoengineering methods are classified into two main groups: Solar Ra- of sulfate aerosol concentrationsmospheric but CO with fixed total amount. The concentration of at- and fraction. Themodel horizontal has resolution 26 is vertical 2 levels and the top of theulations model to (TOM) estimate is at thechange 3 hPa. radiative at forcing the which topthe is of SOM the measured simulations atmosphere as to (Hansen theSST study et and net al., the SOM, radiative 1997). we climate flux performed (2)(2 change. twelve The For cases: other both a set set control include (1 of simulations, fixed- We used the atmospheric general circulationCenter model, for CAM3.1 Atmospheric developed Research at (NCAR) the (Collins National model et CLM3.0 al., and 2004). to It a isto coupled slab to ocean represent the model land the (SOM) with interactionssystem. a with thermodynamic The sea the model ice ocean can model and be also sea configured ice with components prescribed of sea the surface climate temperature ically altered. We believe thatprevious our work study (Ban-Weiss should and be Caldeira,nal 2010), considered distribution because as of not complementary aerosols only toof but we a aerosols we vary which also the facilitates provide latitudi- atribution a of clear constraint aerosols. insight by on fixing the the e total amount 2 Model and experiments distribution. resent realistic latitudinal distribution ofthey aerosols are in designed geoengineering to scenarios. elucidatethe the Rather, latitudinal fundamental distribution properties of of aerosols and the hence climate solar system insolation when reduction is systemat- with constant total amount of sulfate aerosols but with systematically varying latitudinal lacking. In this study, we perform multiple idealized SRM geoengineering simulations Weiss and Caldeira, 2010;solar MacMartin radiation et in both al., space 2012) andcontrol time determine climate so while an the other geoengineered optimal studies world (Irvine reduction isfect et more of in al., similar di 2010; to Ricke the et al.,and 2010) Caldeira analyze the (2010) ef- varymitigate either both the the zonally averaged amountget. changes However, and in a surface simple latitudinal temperature and distribution ordistribution clear the of understanding of water of aerosols aerosols bud- the and to e in hence the solar tropics insolation reduction or (e.g. high more latitudes) concentration on the hydrological cycle and surface temperature is endorn et al., 2009; Jonesgeoengineering et al., would 2010). lead It to hasmean been a shown temperature weakening (Bala of et change al., the iset 2008) global that mitigated al., water SRM exactly. 2008; cycle Further, Ricke whenchanges it et the larger has al., global in 2010) been that some shown the (Robock regions level than of others. compensation Therefore, will vary some with recent residual studies (Ban- Wigley, 2006) and placing space(Early, based 1989). sun CDR shields methods in proposesphere between to the and accelerate Sun the thus and removal they of the2009). deal CO Earth with the root cause of global warmingods (The by Royal reducing Society, theCaldeira, solar 2007; constant (Govindasamy Caldeira and andof Caldeira, Wood, stratospheric 2000; 2008; aerosol Matthews methods Lunt and (Robock et et al., al., 2008) 2008; or Rasch modeled et the al., 2008a, e b; Heck- diation Management (SRM)(Shepherd methods et and al., Carbon 2009).planet dioxide In is Removal reduced the (CDR) by first artificially methods solation approach, enhancing compensates the the the planetary amount radiative forcing ofods due so solar are to that absorption the rising injecting reduced GHGs. by sulfate in- Some the aerosols proposed meth- in the stratosphere (Budkyo, 1982; Crutzen, 2006; 5 5 25 15 20 10 25 15 20 10 Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | - 2 cos b ) to be respec- 2 1 and − represents ) CO a 2 ϕ − × -forcing heats -forcing (Allen 2 2 (2 2 : the plant stomata open less and precipitation sensitivity (% 2 1 0.05 µm and geometric standard physiological forcing refers to the − ) ≈ 2 2 − ) in agreement with previous studies 1 − is dictated by the uniform distribution case Q 2 % K 25392 25391 ≈ is integrated over the sphere, the result is 12.6 Mt ects are not modeled and aerosol loadings for other Q ff ect is essentially equivalent to making latitudinal changes are varied to obtain various distributions of concentrations ff ) while latitudinal distributions are varied. Since aerosols are b 4 physiological forcing (Betts et al., 2007; Cao et al., 2010) which 2 and ering fast response in precipitation for solar and CO ) (1) ff a ϕ cos( radiative forcing is countered by the reduction in solar radiation, the CO 2.0 (as used in a geoengineering scenario in a previous study, Rasch et b while the global mean surface temperature rise is about 2.1 K and the pre- 2 ≈ is the concentration of the additional mass of sulfate aerosols, + 2 − a Q = ) . The fixed-SST simulations lasted for 30 yr and the last 20 yr are used to calculate 4 ϕ ) are the uniform and non-uniform components of the distributions and In agreement with past studies (e.g. Lunt et al., 2008; Bala et al., 2008), we find that The slight warming in the geoengineering case where forcing is close to zero (Fig. 1c) Besides a simulation with uniform aerosol concentration, our geoengineering simula- In each of the geoengineering simulations (Table 1, Fig. 1a) aerosol mass is added ( ϕ and Ingram, 2002; Balathe et troposphere, increases al., the 2008, vertical stability 2009; and Andrews thus et leads to al., precipitation 2009). suppres- CO widely and thus decreasespiration the and canopy causes transpiration surfacewhere which CO warming. in Therefore, turn in reducesphysiological forcing evapotran- the leads zero to a radiative slight forcing warming. case in the geoengineering scenario withprecipitation uniform though distribution the of temperature aerosol change therecurs is is because completely a of mitigated decline di (Fig. in 1b). This oc- tively, values that are similar to Bala et al. (2009). is because of the CO is not counteracted by adirect decrease physiological in response solar of flux. plants CO to elevated CO 3.1 Global mean temperature and precipitationWe response find that the3.5 W radiative m forcing forcipitation doubling increase the is atmospheric aboutusing CO 4.3 the % same (i.e. modeland (Rasch 1d et indicate a al., climate 2008b;change sensitivity Bala in of precipitation 0.53 et K for (Wm al., unit 2009). change in The radiative slopes forcing) of in 1.53 Fig. % (Wm 1c in all the cases. Ourwhich choice had of near-zero 12.6 global Mt mean for temperature change relative to the control case. 3 Results (Table 1, Fig. 1a). However, when ( the latitude. Both tions can be grouped into twoaerosol categories: concentrations (1) at Three “Tropics” the simulations equator withtrations maximum and (2) at Six “Polar” the cases withconcentration poles. maximum are concen- The developed using latitudinal the expression: distribution ofQ the stratospheric sulfate aerosol where be log-normally distributed withdeviation dry median radius al., 2008b). The aerosol indirectspecies e like sea-salt, soil dust,simulations. black and organic carbon are unchanged in each of the to the model background concentration atWeiss the and TOM Caldeira, as 2010). was As done in inis Ban-Weiss a and recent prescribed Caldeira study (2010), (Ban- and this additionalWeiss hence sulfate and Caldeira it (2010), we is introduceis not the constant constraint transported (12.6 that Mt the around. total SO prescribed amount However, at of in aerosol TOM, contrast theto to e the Ban- solar constant. Sulfate aerosol particle size is prescribed and is assumed to ulations. The background sulfate aerosol amountSO in this version of thethe model radiative is forcing. 1.38 The Mt SOMfor climate simulations change lasted analysis for since 60 allstate yr SOM in and simulations approximately the reach 25 a last yr. near-equilibrium 30 yr climate are used simulations. The concentrations of other greenhouse gases are kept constant in all sim- 5 5 20 25 15 10 25 15 20 10 Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | N) in 1991 (Tren- ◦ erent global mean ff and river discharge into ff -forcing is not mitigated by the spec- 2 erent latitudinal distributions along with ff 25394 25393 ective over the poles relative to the tropics (Fig. 1c).This ff . This indicates that a range of climate states are possible for a 2 concentrations (geoengineering simulations) lead to di 2 sets the changes in global mean precipitation but it has a residual warming ff C. Our results are broadly in agreement with other modeling studies: in an Arc- ◦ The variation of global mean surface temperature and precipitation with global mean We find that the prescribed aerosols with di As the polar maximum of the aerosol concentration increases the global mean tem- In contrast, as magnitude of the tropical maximum concentration increases both Our geoengineering simulations with varying aerosol distributions indicate a linear is further confirmed in Fig.planetary 2 which albedo shows compared that to the the Polarplanetary cases Tropics cases. have albedo The smaller changes increase radiative in forcinghave drive associated lower the with albedo temperature changes changeswetter relative while and opposite to thus is the true the uniform for Polar Tropics case cases. cases and henceradiative forcing are (Fig. warmer 1c and over and the 1d) poles shows increasesthe that (Polar1 global to as Polar6) the mean the maximum temperature residual aerosol and forcing concentration precipitation increases increase. and hence Similarly, as the maximum and the changes inprecipitation global changes mean (Fig. temperature 1b), (Fig.and we 1c) precipitation find and changes a between (Fig. linear temperature 1d).Theresidual relationship and Polar between radiative geoengineering radiative scenarios forcing forcing have whilebecause positive solar the forcing is Tropics less scenarios e have negative residual forcing taneously over either land(% or change in ocean. precipitation Weland per also and unit ocean notice change (Fig. that in 1b).(ocean) the as temperature) Here, the we is hydrological ratio have almost sensitivity of defined(ocean) same change the averaged in over surface hydrological land both temperature. sensitivity (ocean) over averaged precipitation land to change in land doubled CO forcings (Fig. 1c and d). Since there are linear relationships between radiative forcing mean precipitation. The reduction insistent precipitation with in observed our decline “Tropics” inthe simulations ocean precipitation are following over the con- land, tropical volcanic runo berth eruption and Mount Dai, Pinatubo 2007). (15 Interestingly, we findchanges that in in global none of mean the surface geoengineering temperature scenarios and precipitation can be mitigated simul- ics1) where the temperature change is nearly zero shows a reduction in the global global mean temperature and precipitation decreases. One of the Tropics cases (Trop- perature increases with concomitant increase inthe global mean linear precipitation relationship as in implied by Fig.almost 1b. o One of theof polar 0.4 maximum SRM simulationstic (Polar3) geoengineering study (Caldeira andover Wood, , 2008) residual with global reducedare solar mean found. constant warming only and enhancements of global precipitation case to residual warming of3 0.7 % K (residual for drying the of Polar6case) case) 2 in % in precipitation for global changes Tropics3 mean which case temperaturefrom are to and about doubling residual 50 of increase % of or CO moreconstant 1 % of amount for the of the changes aerosols. that Polar6 result engineering simulation can be inferredprecipitation from and the linear temperature relationship changes between and changes the in fast response component. relationship between the global mean surface temperaturechange change (Fig. and the 1b). precipitation Thethat regression none lines of do thesimultaneously. not Though distribution pass can the through mitigate totalsimulation the global amount is origin mean fixed, of which we temperature aerosols obtain implies and in a precipitation range each of of 1 K the (residual geoengineering cooling of 0.3 K for the Tropics3 slightly causing much smallerworld change the precipitation in suppression precipitation. caused by Therefore,ified CO in amount of a solar geoengineered forcingprecipitation which is mitigates simulated temperature change. innot This all pass suppression through geoengineering of the simulations origin (the in regression Fig. line 1b). Therefore does the precipitation change in any geo- sion (Cao et al., 2012). In contrast, solar forcing tends to heat the atmosphere only 5 5 25 15 20 10 25 15 20 10 Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | erent nature of the ff . This is because of the di 25396 25395 2 CO × ) show, similar to changes in global annual mean values, case is significant over the whole globe but not significant 2 erence (RMSD) of the geoengineering simulations with re- 2 ff CO forcing is uniform over the whole globe. In Polar3 case, the CO × 2 × case, both temperature and precipitation changes are large and signif- 2 case. 2 CO × CO , Uniform, Polar3, Tropics1, Polar6 and Tropics3 cases. We notice that the ra- × 2 forcing and solar forcing: solar forcing is larger in the tropics and smaller in the CO 2 In the 2 Figure 7 shows the spatial pattern of radiative forcing in selected simulations: The root mean square di × to the 2 regions. However, temperature isin largely precipitation mitigated relative in toshows Tropics1 the but the uniform there extreme distribution cases; is case.warms the decrease The the case last planet with four while largest panelsthe planet polar the of with weighting Fig. case large (Polar6) reduction 8 withperatures in significantly largest precipitation. mostly The tropical occur seasonal variations weighting in inweaker (Tropics3) the response residual overcools in tem- high the latitudes summer (Fig. with(Fig. 9) 7; following stronger the right response seasonal panels). cycle in The oflarge magnitude the radiative forcing of in winter seasonal the and variations tropics in precipitation in response the is geoengineering cases but with reduced intensity compared The uniform geoengineering casereduced (Uniform) precipitation shows relative mitigation to in 1 CO the temperature with poles whereas the CO change in precipitation is largely mitigated but there is significant warming over large diative forcing in the 2 in most regions inlocations the in geoengineering cases. case The oftropical radiative regions Polar forcing and cases. is positive In positive in polar in the regions. most Tropics cases, theicant over forcing the whole is globe negative (Fig.than 8). in in The the the temperature tropics, increase in overtet agreement poles with al., is previous much 2008; studies larger Matthews (Caldeira and and Wood, Caldeira, 2008; 2007; Lun- Robocket al., 2008; Rasch et al., 2008b). the monotonic increase inchanges precipitation in with zonal poleward meanchanges weighting precipitation in is zonal minus mean clearly evaporation precipitation (water seen. (Figs. budget) 4c, The are 5e and similar 5f). 2 to tion (Fig. 6) shows that the position of ITCZ remains the same in all the cases and (Figs. 4b, 5c and 5d).We findto large be changes seen in precipitation as in shifts the in tropics the which is intertropical likely convergence zone (ITCZ) but closer examina- and the control case (1 a monotonic increase at eachtice latitude a with increased similar polar monotonic weightingtemperature increase (Fig. (Fig. in 4a). 5a We zonal-mean and no- landtion 5b). show and Further, residual zonal-mean we high ocean find latitudewarming surface warming. that in In almost the the all Tropics high cases, geoengineeringchange we latitudes simula- find in and smaller zonal-mean cooler residual precipitation tropics.trol between Similar case the to show geoengineering temperature a changes, cases monotonic the and increase the at con- each latitude with increased polar weighting the least distance from theobjective origin is in to Fig. minimize 3, RMSD suggesting in that both it temperature has and3.2 the precipitation least simultaneously. RMSD if Precipitation the and temperature response inThe Tropics and change Poles in zonal-mean surface temperature between the geoengineering cases tropical maximum is increased (Fig.the 3). control We scenario normalize so RMSDbe the by compared RMSD standard to between deviation the a in ratio regional geoengineered of and variations spatial a in control RMSD theof world control and zonal can simulation. standard mean Figure deviation 3a RMSD ofRMSD shows the and the the control standard spread simulation deviation isprecipitation. while is more, the In shown 0.40 ratio case in to of0.95 Fig. 1.4 RMSD for 3b. for in surface In temperature surface zonal case and temperature means, of 0.27 and the spatial to spread 0.25 0.38 is to for relatively precipitation. 0.40 less: The for uniform 0.30 case to has variation is noticed. spect to the controlscenario case, shows normalized that by the RMSDtion the in spatial of temperature standard aerosols increases deviation as at in the the the maximum poles concentra- control increases and the RMSD in precipitation increases as aerosol concentration over the equator increases (Tropics1 to Tropics3), an opposite 5 5 25 15 20 10 25 15 20 10 Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | cient ffi erent climate ff sets the warming leads ects on di ff ff erent latitudinal distribution it is ff , 2002. cient to infer the global mean climate sets the precipitation results in a rela- ffi ff 25398 25397 10.1038/nature01092 ects of deep ocean circulation are not modeled here. However, ff Financial support for A. Modak was provided by the Divecha Centre for In this study, we have not considered the consequences of sulfate aerosol chemistry In summary, for a fixed total mass of aerosols, we find that the global mean climate Our findings should be viewed in the light of the limitations and uncertainties involved In agreement with earlier studies (e.g. Bala et al., 2008), we find that both tempera- cycle, Nature, 419, 224–232, doi: to infer the global mean climate change. to the uniformbecause distribution the case global (whichcompared mean mitigates to residual global the radiative meanmaximum forcing uniform temperature in is case. change) the positive The tropics. inmean opposite Further, radiative these our forcing, is not study cases true the clearly when details when indicates of that aerosol latitudinal knowledge distribution concentration of of is aerosols, global is su Irvine et al., 2010; Ricke et al., 2010) focused onon regional the ozone disparities. layer (Tilmes etcomponents al., and 2009). the Our e modelwe lacks a believe dynamic our ocean and resultswould sea be on ice unchanged temperature when and additional components precipitation and is feedbacks are so included. is fundamental warmer that and wetter they when aerosol concentration is maximum over the poles relative jected aerosol precursors intowhile other to studies the (Heckendorn et stratosphereHommel al., with and 2009; Graf, fixed Pierce 2011; et particle English al., etcluding size 2010; al., the Niemeier distribution 2012) microphysics et have of al., demonstrated particle 2010; theprimarily growth. importance on Further, of we global in- have mean focused climate our while investigation several other studies (e.g Robock et al., 2008; variables (Rasch et al., 2008b). Some modeling studies (Robock et al., 2008) have in- time and is shown to be important in precisely estimating the e to a drier climatetively while warmer the world distribution (note which that o tion). Bala For et a al. fixed (2008)possible used total to a amount achieve uniform a of solar range aerosols of insolationstates. global but reduc- mean with radiative forcing di and thus a range of climate in this study. Our simulations are(to highly reduce idealized the as solar we insolation) have instead prescribedscribed of sulfate a aerosol injecting and fixed transporting particle them. size We have distribution pre- but particle size distribution would evolve with global mean forcing (Fig. 11) for this set of simulationsture too. and precipitation changes cannoting be simulations mitigated considered simultaneously in insolar this all insolation geoengineer- study reduction). (that The is, latitudinal even distribution with which o non-uniform distribution of the model (Fig. 1b).tensity We as also the observe maximum a aerosolincreases similar concentration are monotonic over of increase the the in poles orderintensity precipitation increases of (99th in- 10 (Fig. percentile) % between 10). for the The low extremeconfirm cases intensity that (Tropics3 (5th and global Polar6). percentile) In mean and orderchange radiative 2–3 to % forcing we for is performed large su four additionalaerosols geoengineering varied simulations (10 Mt, with totalfind 11 Mt, that amount 13 the of Mt global and mean temperature 14 Mt) and precipitation for changes the follow uniform the changes distribution in case. We In this study, we findthe that global when mean the radiative forcing latitudinalture changes distribution as which of indicated leads sulfate by to the aerosols changesin climate is in global sensitivity altered surface mean of tempera- the precipitation model are (Fig. simulated 1c). as Consequent dictated changes by the hydrological sensitivity of 4 Discussion and conclusions Allen, M. R. and Ingram, W. J.: Constraints on future changes in climate and the hydrologic Acknowledgements. Climate Change, Indian Institute ofsearch Centre, Science. Indian We Institute thank of the Science Supercomputer for Education providing and the Re- computational resources. References 5 5 25 15 20 10 25 15 20 10 Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | , - 2 , 2006. s in geo- , 2009. ff , 2007. 10.5194/acp- , 2010. 10.1007/s00382-009- , 2008. STR National Center for + 10.1073/pnas.0711648105 , 2011. ect, JBIS-J. Brit. Interpla., 42, ff , 2009. 10.1007/s10584-006-9101-y 10.1073/pnas.0700419104 , 2010. 10.1088/1748-9326/4/4/045108 ce HadGEM2 climate model and comparison 10.1029/2010GL044447 ffi 10.1002/asl.285 10.1098/rsta.2008.0132 set greenhouse e 25400 25399 ff , 2012. , 2012. 10.1175/2008JCLI2759.1 , 2007. , 2010. , 2008. 10.1088/1748-9326/5/3/034009 due to plant responses to increasing carbon dioxide, Nature, 448, 1037– ff 10.1038/nclimate1722 , 2012. y P. B., and Taylor, K. E.: Impact of geoengineering schemes on the global hy- ff 10.1038/nature06045 , 2009. 10.1029/2008GL033674 10.1088/1748-9326/7/3/034015 10.1073/pnas.0913000107 neering, P. Natl. Acad. Sci. USA, 104, 9949–9954, doi: 245–284, 2000. evaluation of thedoi: climatic impacts of geoengineering, Geophys.engineering through Res. optimal choice Lett., of3, non-uniform 35, radiative 365–368, forcing, L12710, doi: Nature Climate Change, Atmospheric Research Boulder CO 226 pp., 2004. a policy dilemma?, Climatic Change, 77, 211–220, doi: 567–569, 1989. spheric SO2 injection: results fromwith the the Met Goddard O Institute fordoi:10.5194/acp-10-5999-2010, Space 2010. Studies ModelE, Atmos. Chem. Phys., 10, 5999–6006, and solar irradiancedoi: on the time scale of daysBriegleb, to B., Bitz, weeks, C., Environ. Lin,atmosphere S.-J., Res. model Zhang, Lett., M., (CAM and 7, 3.0) Dai, 034015, Y.: NCAR Description Tech. of Rep. the NCAR NCAR/TN-464 community aerosols, Atmos. Sci. Lett., 12, 168–175, doi: impacts, Geophys. Res. Lett., 37, L18702, doi: 1 to the FourthSolomon, Assessment S., Report Qin, of D.,Miller, the H. Manning, Intergovernmental L.: Panel M., (Cambridge on & Chen, Climate New Z., York: Change Cambridge Marquis, ed University Press) M., 996 Averyt, pp., 2007. K. B., Tignor, M., and doi: Philos. T. Roy. Soc. A, 366, 4039–4056, doi: physiological forcing to future climate change, P. Natl. Acad. Sci. USA, 107, 9513–9518, 102, 6831–6864, 1997. son, L. W., and Peter,ture T.: and The ozone, impact Environ. of Res. geoengineering Lett., aerosols 4, 045108, on doi: stratospheric tempera- stratospheric sulfur geoengineering, Atmos. Chem.12-4775-2012 Phys., 12, 4775–4793, doi: induced climate change, Geophys. Res. Lett., 27, 2141–2144, 2000. drological cycle, P. Natl.2008. Acad. Sci. USA, 105, 7664–7669, doi: Lett., 5, 034009, doi: L., Huntingford, C., Jones,continental C. runo D., Sexton, D.1041, M. doi: H., and Webb, M. J.: Projected increase in 0583-y tions for the global hydrological cycle, Clim. Dynam. 35, 423–434, doi: J. Climate., 22, 2557–2570, doi: Matthews, H. D. and Caldeira, K.: Transient climate-carbon simulations of planetary geoengi- MacMartin, D. G., Keith, D. W., Kravitz, B., and Caldeira, K.: Management of trade-o Lunt, D. J., Ridgwell, A., Valdes, P. J., and Seale, A.: ‘Sunshade world’: a fully coupled GCM Crutzen, P. J.: Albedo enhancement by stratospheric sulfur injections: AEarly, contribution J. to resolve T.: Space-based solar shield to o Keith, D. W.: Geoengineering the climate: history and prospect, Annu. Rev. Energ. Env., 25, Jones, A., Haywood, J., Boucher, O., Kravitz, B., and Robock, A.: Geoengineering by strato- Collins, W. D., Rasch, P.J., Boville, B. A., Hack, J. J., McCaa, J. R., Williamson, D. L., Kiehl, J. T., Hommel, R. and Graf, H. F.: ModellingIrvine, P. the J., Ridgwell, size A., distribution and Lunt, of D. J.: geoengineered AssessingIPCC stratospheric the 2007 Climate regional Change disparities 2007: in The geoengineering Physical Science Basis Contribution of Working Group Cao, L., Bala, G., and Caldeira, K.: Climate response to changes in atmospheric carbon dioxide Cao, L., Bala, G., Caldeira, K., Nemani, R., and Ban-Weiss, G. A.: Importance of carbon dioxide Caldeira, K. and Wood, L.: Global and Arctic : Numerical model studies, Heckendorn, P.,Weisenstein, D., Fueglistaler, S., Luo, B. P.,Rozanov, E., Schraner, M., Thoma- Hansen, J., Sato, M., and Ruedy, R.: Radiative forcing and climate response, J. Geophys. Res., Budkyo, M. I.: The Earth’s Climate Past and Future, Academic, 1982. English, J. M., Toon, O. B., and Mills, M. J.: Microphysical simulationsGovindasamy, B. of and sulfur Caldeira, burdens K.: from Geoengineering Earth’s radiation balance to mitigate CO Betts, R. A., Boucher, O., Matthew, C., Cox, P. M., Falloon, P. D., Gedney, N., Hemming, D. Bala, G., Du Ban-Weiss, G. A. and Calderia, K.: Geoengineering as an Optimization Problem, Environ. Res. Bala, G., Caldeira K., and Nemani, R.: Fast versus slow response in climate change: implica- Andrews, T., Forster, P.M., and Gregory, J. M.: A surface energy perspective on climate change, 5 5 30 25 30 20 25 15 20 10 15 10 Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | , 2010. cient forma- ffi , 2008a. 3.00 12.60 2.00 12.60 1.00 12.60 , 2010. − − − 10.1002/asl.304 10.1098/rsta.2008.0131 , 2006. 10.1029/2010GL043975 Total Mass Total Mass from Total Mass 25401 25402 component (Mt) component (Mt) ) from uniform non-uniform (Mt) 2 − ects of Mount Pinatubo volcanic eruption on the hy- 5.34 13.60 ff 16.02 15.60 10.67 14.60 − − − , 2007. , 2008b. , 2009. , 2008. 10.1126/science.1131728 ) (mg m a b erent geoengineering experiments. Total additional mass is 12.6 Mt 2 ff − Description of di 10.1029/2007GL030524 10.1029/2008JD011420 10.1029/2008JD010050 10.1029/2007GL032179 in all the geoengineering simulations but the distribution varies. Tropics3 30.58 Tropics2 28.62 Name of the UniformPolar1Polar2Polar3 24.70Polar4Polar5 23.52Polar6 21.56Tropics1 19.60 – 3.19 17.64 8.55 15.68 13.89 13.72 26.66 19.22 24.56 12.60 29.90 12.00 11.00 10.00 9.00 8.00 7.00 0.60 – 1.60 2.60 12.60 12.60 12.60 3.60 12.60 4.60 12.60 5.60 12.60 12.60 Experiments (mg m 4 ence, 314, 452–454, doi: drological cycledoi: as an analog of geoengineering, Geophys. Res. Lett., 34, L15702, engineered aerosols on thedoi: troposphere and stratosphere, J. Geophys. Res., 114, D12305, engineering withdoi: tropical and Arctic SO2Rayner, injections, S., Redgwell, J. C., and Watson, Geophys.and A.: uncertainty, Geoengineering The Res., the Royal climate: Academy, 113, 2009. science, governance D16101, management, Nat. Geosci., 3, 537–541, 2010. tion of stratospheric aerosolaircraft, for Geophys. climate Res. engineering Lett., by 37, L18805, emission doi: of condensible vaporand from Garcia, R. R.:aerosols, Philos. An T. overview Roy. Soc. of A, 366, geoengineering 4007–4037, of doi: climatestratospheric using sulfate stratospheric aerosols: sulphate Thedoi: role of particle size, Geophys. Res. Lett., 35, L02809, aerosol on the emission strategy, Atmos. Sci. Lett., 12, 189–194, doi: Table 1. SO Wigley, T. M. L.: A combined mitigation/geoengineering approach to climate stabilization, Sci- Tilmes, S., Garcia, R. R., Kinnison, D. E., Gettelman,Trenberth, A., K. and Rasch, E. P. J.: and Impact Dai, of geo- A.: E Shepherd, J., Caldeira, K., Haigh, J., Keith, D., Launder, B., Mace, G., MacKerron, G., Pyle, J., Robock, A., Oman, L., and Stenchikov, G. L.: Regional climate responses to geo- Ricke, K. L., Morgan, M. G., and Allen, M. R.: Regional climate response to solar-radiation Pierce, J. R., Weisenstein, D. K., Heckendorn, P., Peter, T., and Keith,Rasch, D. W.: P. J., E Tilmes, S., Turco, R. P., Robock, A., Oman, L.,Rasch, P. Chen, J., Crutzen, C. P. J., C., and Stenchikov, Coleman, G. D. B.: L., Exploring the geoengineering of climate using Niemeier, U., Schmidt, H., and Timmreck, C.: The dependency of geoengineered sulfate 5 25 20 15 10 Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper |

20 represent the averages P ∆ TS and ∆ 25404 25403 Radiative forcing (RF) vs. surface temperature change. Polar

(c) the horizontal and vertical bars represent the standard error of the Radiative forcing vs. % precipitation change. Precipitation increases (d) (d) and Surface temperature change (K) vs. precipitation change (%) relative to the (c) , (b) (b)

Latitudinal profiles of sulfate aerosol concentration in the SRM geoengineering exper- case from slab ocean simulations (global mean values – squares, land mean values 2

Figure 1 Figure

1xCO Fig. 1. (a) iments. Polar1–6 have maximum concentrationat over the the equator. poles and Tropics1–3 have maximum – stars, ocean mean valuesform – case triangles). and There a is concomitantNone warming of increase in all the in Polar regression precipitation. cases lines Oppositeigated relative pass is to simultaneously. through the the In origin; case uni- the temperature forover and case Tropics the cases. precipitation of respective cannot domain. land becases mit- and have larger ocean, forcing relativetrue to for the Tropics cases. uniform case and hence are warmer while opposite is with residual RF (i.e. withweighting. increase In in polar weighting) whilerespective decreases variables with which increase are in calculatedcase tropical from of the radiative forcing last it 30 yr is calculated of from 60 yr the SOM last simulations 20 yr while of in 30 yr fixed-SST simulations. 492 493 491 Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper |

21

22 22 25406 25405

Figure 3 Figure

Figure 3 Figure

498 497 498 497

Change in planetary albedo in fixed-SST vs. surface temperature change in slab ocean

Figure 2 Figure

Fig. 2. geoengineering simulations. Thetemperature radiative forcing changes. associated Polar with casesand hence albedo have are changes lower warmer drive albedovertical and wetter the changes bars while represent relative oppositeerrors is to the are true standard the calculated for from error uniform Tropics cases. theare of case last The calculated 30 the horizontal from yr and the respective of last 60 variables; yr 20 yr SOM temperature of simulations standard 30 while yr albedo fixed-SST standard simulations. errors 495 496 494 Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | (b) PmE: polar ∆ ) and precipita-

P 23 ∆ Zonal mean (c) TS), precipitation ( ∆ TS increases monotonically with increase in ∆ axis represents the best precipitation mitigating 25408 25407 x axis represents the best surface temperature mitigating

Zonal mean y (a) erence (RMSD) of surface temperature and precipitation between ff PmE). ∆ 4

Figure : polar maximum causes enhanced precipitation.

P 500 501 499 ∆ Zonal means of change in surface temperature ( Root mean square di Fig. 4. tion minus evaporation ( maximum concentrations over theZonal poles mean and decreases with increase in tropical maxima. scenario while the one closest to maximum causes enhanced precipitationthe minus last 30 evaporation. yr Results of shown 60 yr are simulations. averages of geoengineering and control simulationscenario normalized computed by for the spatial global standard domainThe (top deviation annual panel) in and means the for the control standard of zonal averages deviation. the (bottom panel). Simulation last nearest 30 yr to of the 60 yr control simulation are used to estimate the Fig. 3. scenario. Scenarios with maximumtemperature aerosol and concentrations conversely at simulations theprecipitation. with poles maximum have larger at RMSD the in equator have larger RMSD in Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper |

(c, 25 Zonal

24

(a, b) ) and precipita- P ∆ TS), precipitation ( Polar maximum causes enhanced pre- ∆ (e, f) 25410 25409

PmE) over ocean (left panels) and land (right panels). ∆

Figure 5 Figure TS increases monotonically with increase in the magnitude of maximum concentration Changes in zonal mean surface temperature ( Zonal mean precipitation over the globe. The position of intertropical convergence zone

Figure 6 Figure ∆

Polar maximum causes enhanced precipitation. 503 504 502 507 506 505 (ITCZ) remains the samecreases monotonically in over all the equator theshown as geoengineering are the global averages cases. of mean The radiative the forcing last zonal increases. 30 mean Results yr of precipitation 60 de- yr simulations. Fig. 6. Fig. 5. tion minus evaporation ( mean of aerosols over poles andd) decreases with increase in the magnitude of tropical maximum. cipitation minus evaporation. Results shown are averages of the last 30 yr of 60 yr simulations. Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | t

and extreme scenarios

27 2 26 CO × test. Both surface temperature and pre- t

25412 25411

, uniform, and some Polar and Tropics geoengineering scenarios. 2

Figure 8 Figure CO

Figure 7 Figure

× 512 513 511 509 510 508 ). Hatching indicates the region where the changes are significant at 1 % 2 , uniform, and some Polar and Tropics geoengineering scenarios relative to 2 CO CO × × Changes in annual-mean surface temperature (left panels) and precipitation (right pan- Spatial pattern of radiative forcing (left panels) and the seasonal cycle of radiative forcing test. Results shown aretemperature averages and of sea ice the fraction. last 20 yr of 30 yr simulations with fixed sea surface In the Unifrom andnegative Tropical cases, forcing in there the is low acycle latitudes residual is indicating clearly positive an visible forcing inexact in compensation. incases The the the the residual polar high residual seasonal regions latitudes seasonal inwhere and cycle the has the Uniform a changes case much while are in smaller significant the strength. at Polar3 Hatching 1 and indicates % Tropics1 the level. region Significance level was estimated by Student (right panels) in the 2 level. Significance level was estimated using Student the control (1 Fig. 8. els) in the 2 cipitation changes are large(Polar6 and and significant Tropics3). Although everywhere significant intion over the changes large are 2 regions, small both in temperaturebut the Uniform and not case. precipita- global Polar3 scenario mean mitigatesbut temperature global with mean reduced while precipitation precipitation. Tropics1 Results scenario shown are mitigates averages of global the mean last 30 temperature yr of 60 yr simulations. Fig. 7. Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper |

29

28 ). Seasonal variations in temperature re- 2 , uniform and some Polar and Tropics geo- 2 CO × test. In case of surface temperature, the change CO t × 25414 25413 Ocean. There is a monotonic increase in precipitation for all (c)

Figure 9 Figure Land,

515 516 514 (b)

Figure 10 Figure

518 517 Globe, (a) Percentile values (p5, p25, median, p75, p85, p90, p95 and p99) of precipitation inten- Seasonal cycle of the changes in the zonally averaged surface temperature (left pan- percentile values as themonthly maximum mean concentration precipitation of aresimulations aerosols used are over used to poles for calculate increases. the the statistics. Grid-level percentile values.The last 30 yr of 60 yr sity over Fig. 10. in seasonal cycle is significant30 for yr both of the 60 Polar yr cases. simulations. Results shown are averages of the last els) and precipitation (right panels) in the 2 Fig. 9. engineering scenarios relative tosponse the mostly control occur (1 in the highin latitudes the with summer. larger warming HatchingSignificance in indicates the level winter was the and estimated region weaker by warming where Student the changes are significant at 1 % level. Discussion Paper | Discussion Paper | Discussion Paper | Discussion Paper | Radiative forcing vs.

30 (b) 25415 the horizontal and vertical bars represent the standard (b) and

(a)

Figure 11 Figure

520 519 Radiative forcing (RF) vs. surface temperature change. % precipitation change for14 uniform Mt. distribution More scenarios aerosol withdrier 10 mass Mt, climate, leads 11 and Mt, to smaller 12.6 Mt,warmer negative aerosol and 13 residual Mt wetter mass climate. and radiative leads In forcing to positive and residual hence radiative cooler forcing and and hence Fig. 11. (a) error of the respectivesimulations for variables. temperature Results and shown precipitationare are while used the averages for last radiative of 20 forcing yr the calculations. of last 40 yr 20 yr fixed-SST simulations of 50 yr SOM Sensitivity of Simulated Climate to latitudinal distribution of solar insolation reduction in Solar Radiation Management

Angshuman Modak* and Govindasamy Bala Divecha Centre for Climate Change & Centre for Atmospheric and Oceanic Sciences Indian Institute of Science Bangalore – 560 012, India

*Corresponding author Email: [email protected]

1

1 Abstract 2 3 Solar radiation management (SRM) geoengineering has been proposed as a potential 4 option to counteract climate change. We perform a set of idealized geoengineering simulations 5 using NCAR CAM3.1 to understand the global hydrological implications of varying the 6 latitudinal distribution of solar insolation reduction in SRM methods. To reduce the solar 7 insolation we have prescribed sulfate aerosols in the stratosphere. The radiative forcing in the

8 geoengineering simulations is the net forcing from a doubling of CO2 and the prescribed

9 stratospheric aerosols. We find that for a fixed total mass of sulfate aerosols (12.6 Mt of SO4), 10 relative to a uniform distribution which nearly offsets changes in global mean temperature from a

11 doubling of CO2, global mean radiative forcing is larger when aerosol concentration is maximum 12 at the poles leading to a warmer global mean climate and consequently an intensified 13 hydrological cycle. Opposite changes are simulated when aerosol concentration is maximized in 14 the tropics. We obtain a range of 1K in global mean temperature and 3% in precipitation changes 15 by varying the distribution pattern in our simulations: this range is about 50 % of the climate

16 change from a doubling of CO2. Hence, our study demonstrates that a range of global mean 17 climate states, determined by the global mean radiative forcing, are possible for a fixed total 18 amount of aerosols but with differing latitudinal distribution. However, it is important to note that 19 this is an idealized study and thus not all important realistic climate processes are modeled. 20 21 (Key words: Climate Change, Geoengineering, Solar Radiation Management, Hydrological 22 Cycle, Sulfate aerosols, Climate Model) 23

2

24 1. Introduction 25 26 Atmospheric concentrations of the greenhouse gases (GHGs) such as carbon dioxide

27 (CO2), methane (CH4) and nitrous oxide (N2O) have been increasing since pre-industrial periods 28 primarily because of fossil fuel use and land-use change (IPCC, 2007). Their increase has the 29 potential to cause long term climate change by altering the planetary radiation budget. To 30 moderate future climate change and its impacts, several geoengineering proposals have been 31 made recently. By definition, geoengineering is an intentional large-scale manipulation of the 32 environment, particularly intended to counteract the undesired consequences of anthropogenic 33 climate change (Keith, 2000). 34 Proposed geoengineering methods are classified into two main groups: Solar Radiation 35 Management (SRM) methods and Carbon dioxide Removal (CDR) methods (Shepherd et al., 36 2009). In the first approach, the amount of solar absorption by the planet is reduced by 37 artificially enhancing the planetary albedo so that the reduced insolation compensates the 38 radiative forcing due to rising GHGs. Some proposed methods are injecting sulfate aerosols in 39 the stratosphere (Budkyo,1982; Crutzen, 2006; Wigley, 2006) and placing space based sun 40 shields in between the Sun and the Earth (Early, 1989). CDR methods propose to accelerate the

41 removal of CO2 from the atmosphere and thus they deal with the root cause of global warming 42 (Shepherd et al., 2009). 43 Past climate modeling studies have modeled the effects of space-based SRM methods by 44 reducing the solar constant (Govindasamy and Caldeira, 2000; Matthews and Caldeira, 2007; 45 Caldeira and Wood, 2008; Lunt et al., 2008) or modeled the effects of stratospheric aerosol 46 methods (Robock et al., 2008; Rasch et al., 2008a; Rasch et al., 2008b; Heckendorn et al., 2009; 47 Jones et al., 2010). It has been shown (e.g., Bala et al., 2008) that SRM geoengineering would 48 lead to a weakening of the global water cycle when the global mean temperature change is offset 49 exactly. A recent study (Tilmes et al., 2013) using 12 models from Geoengineering Model 50 Intercomparison Project (GeoMIP) confirms this weakening of hydrological cycle under multi- 51 model framework. Further, it has been shown (Robock et al., 2008; Ricke et al., 2010; Tilmes et 52 al., 2013) that the level of compensation will vary with residual changes larger in some regions 53 than others. Therefore, some recent studies (Ban-Weiss and Caldeira, 2010; MacMartin et al., 54 2012) determine an optimal reduction in solar radiation in both space and time so the

3

55 geoengineered world is more similar to the control climate while other studies (Irvine et al., 56 2010; Ricke et al., 2010) analyze the effect of different levels of uniform SRM forcing on 57 regional climate response. Ban-Weiss and Caldeira (2010) vary both the amount and latitudinal 58 distribution of aerosols to offset either the zonally averaged changes in surface temperature or the 59 water budget. However, a simple and clear understanding of the effects of systematically varying 60 the latitudinal distribution of aerosols and hence solar insolation reduction (e.g., more 61 concentration in the tropics or high latitudes) on the hydrological cycle and surface temperature 62 is lacking. In this study, we perform multiple idealized SRM geoengineering simulations with 63 constant total amount of sulfate aerosols but with systematically varying latitudinal distribution. 64 We caution that our simulations are highly idealized and they are not meant to represent 65 realistic latitudinal distribution of aerosols in geoengineering scenarios. Rather, they are designed 66 to elucidate the fundamental properties of the climate system when the latitudinal distribution of 67 aerosols and hence solar insolation reduction is systematically altered. We believe that our study 68 should be considered as complementary to a previous work (Ban-Weiss and Caldeira, 2010), 69 because not only we vary the latitudinal distribution of aerosols but we also provide a constraint 70 by fixing the total amount of aerosols which facilitates a clear insight on the effects of varying 71 the latitudinal distribution of aerosols. 72 73 2. Model and Experiments 74 75 We used the atmospheric general circulation model, CAM3.1 developed at the National 76 Center for Atmospheric Research (NCAR) (Collins et al., 2004). It is coupled to the land model 77 CLM3.0 and to a slab ocean model (SOM) with a thermodynamic sea ice model to represent the 78 interactions with the ocean and sea ice components of the climate system. The model can be also 79 configured with prescribed sea surface temperature and sea ice fraction. The horizontal resolution 80 is 2° latitude and 2.5° longitude and the model has 26 vertical levels and the top of the model 81 (TOM) is at 3hPa. 82 We performed two sets of simulations: 1) fixed-SST (sea surface temperature) 83 simulations to estimate the radiative forcing which is measured as the net radiative flux change at 84 the top of the atmosphere (Hansen et al., 1997). This method allows the rapid adjustment of the 85 atmosphere and land components before radiative forcing is evaluated. The other set include the

4

86 SOM simulations to study the climate change. For both set of simulations, fixed-SST and SOM,

87 we performed twelve cases: a control (1xCO2),doubled CO2 climate (2xCO2) and ten 88 geoengineering simulations each with differing latitudinal distribution of sulfate aerosol

89 concentrations but with fixed total amount. The concentration of atmospheric CO2 in 1xCO2 is

90 390ppm and is 780ppm in 2xCO2 and geoengineering simulations. The concentrations of other 91 greenhouse gases are kept constant in all simulations. The background sulfate aerosol amount in

92 this version of the model is 1.38 Mt SO4. The fixed–SST simulations lasted for 30 years and the 93 last 20 years are used to calculate the radiative forcing. The SOM simulations lasted for 60 years 94 and the last 30 years are used for climate change analysis since all SOM simulations reach a 95 near-equilibrium climate state in approximately 25 years. 96 As in Ban-Weiss and Caldeira (2010), the additional sulfate is prescribed in the 97 geoengineering cases (Table 1, Figure 1a) and hence it is not transported around. However, in 98 contrast to Ban-Weiss and Caldeira (2010), we introduce the constraint that the total amount of

99 aerosol is constant (12.6 Mt SO4) while latitudinal distributions are varied. Since aerosols are 100 prescribed at TOM, the effect is essentially equivalent to making latitudinal changes to the solar 101 constant. Sulfate aerosol particle size is prescribed and is assumed to be log-normally distributed 102 with dry median radius ≈ 0.05μm and geometric standard deviation ≈ 2.0 (as used in a 103 geoengineering scenario in a previous study (Rasch et al., 2008b)). The aerosol indirect effects 104 are not modeled and aerosol loadings for other species like sea-salt, soil dust, black and organic 105 carbon are unchanged in each of the simulations. 106 Besides a simulation with uniform aerosol concentration, our geoengineering simulations 107 can be grouped into two categories: 1) Three “Tropics” simulations with maximum aerosol 108 concentrations at the equator and 2) Six “Polar” cases with maximum concentrations at the poles. 109 The latitudinal distribution of the stratospheric sulfate aerosol concentration are developed using 110 the expression: 111 Q(φ) = a + bcos(φ) 112 where Q is the concentration of the additional mass of sulfate aerosols, a and bcos(φ) are the 113 uniform and non-uniform components of the distributions and φ represents the latitude. Both a 114 and b are varied to obtain various distributions of concentrations (Table 1, Figure 1a). However, 115 when Q is integrated over the sphere, the result is 12.6Mt in all the cases. Our choice of 12.6 Mt 116 for Q is dictated by the uniform distribution case which had near-zero global mean temperature

5

117 change relative to the control case. In each of the geoengineering simulations aerosol mass is 118 added to the model background concentration at the TOM as was done in a recent study (Ban- 119 Weiss and Caldeira, 2010). An experiment where the same total mass (12.6Mt) of aerosol is 120 distributed uniformly over the globe between 61hPa to 9.8hPa (15 – 30km) with a maximum at 121 30hPa (22km) showed that the radiative forcing is nearly the same as in our uniform distribution 122 geoengineering case and hence the main conclusions reached in this study are unlikely to be 123 altered. 124 125 3. Results 126 3.1 Global mean Temperature and Precipitation Response 127

128 We find that the radiative forcing for doubling the atmospheric CO2 (2xCO2) to be 3.5 129 W/m2 while the global mean surface temperature rise is about 2.1K and the precipitation increase 130 is about 4.3% (i.e. ≈ 2%/K) in agreement with previous studies using the same model (Rasch et 131 al., 2008b; Bala et al., 2009). The slopes in Figures 1c and 1d indicate a climate sensitivity of 132 0.53K/Wm−2 and precipitation sensitivity (% change in precipitation for unit change in radiative 133 forcing) of 1.5%/Wm−2 respectively, values that are similar to Bala et al. (2009). 134 The slight warming in the geoengineering case where forcing is close to zero (the case

135 Polar 1 in Figure 1c) is because of the CO2 physiological forcing (Betts et al., 2007; Cao et al.,

136 2010) which is not counteracted by a decrease in solar flux. CO2 physiological forcing refers to

137 the direct physiological response of plants to elevated CO2: the plant stomata open less widely 138 and thus decrease the canopy transpiration which in turn reduces evapotranspiration and causes

139 surface warming. Therefore, in the zero radiative forcing case where CO2 radiative forcing is

140 countered by the reduction in solar radiation, the CO2-physiological forcing leads to a slight 141 warming. 142 In agreement with past studies (e.g., Lunt et al., 2008; Bala et al., 2008; Tilmes et al. 143 2013), we find that in the geoengineering scenario with uniform distribution of aerosol there is a 144 decline in precipitation though there is a near cancellation of surface temperature change (Figure 145 1b). This occurs because of differing fast response (changes that occur before global mean

146 surface temperature) in precipitation for solar and CO2-forcing (Allen and Ingram, 2002; Bala et

147 al., 2008; Bala et al., 2009; Andrews et al., 2009): longwave absorption by CO2 in the

6

148 atmosphere can contribute to increased vertical stability and suppress precipitation but this fast 149 response mechanism is nearly absent for solar forcing because the atmosphere is nearly 150 transparent to solar radiation. However, since the slow response (changes that are associated with

151 global mean surface temperature change) is same for CO2 and solar forcings (Bala et al. 2010),

152 the total changes in rainfall are larger to solar forcing than to equivalent CO2 forcing. Because of

153 this differing hydrological sensitivity to solar and CO2 forcing, insolation reductions (in 154 geoengineering scenarios) sufficient to offset the entirety of global-scale temperature increases 155 would lead to a decrease in global mean precipitation. This suppression of precipitation is 156 simulated in all geoengineering simulations (the regression line does not pass through the origin 157 in Figure 1b). 158 Our geoengineering simulations with varying aerosol distributions indicate a linear 159 relationship between the global mean surface temperature change and the precipitation change 160 (Figure 1b). The regression lines do not pass through the origin which implies that none of the 161 distribution can offset global mean temperature and precipitation simultaneously. Though the 162 total amount of aerosols in each of the geoengineering simulation is fixed, we obtain a range of 163 1K (residual cooling of 0.3 K for the Tropics3 case to residual warming of 0.7 K for the Polar6 164 case) in global mean temperature and 3 % (residual drying of 2 % for Tropics3 case to residual 165 increase of 1 % for the Polar6 case) in precipitation changes which are about 50 % or more of the

166 changes that result from doubling of CO2. This indicates that a range of climate states are 167 possible for a constant amount of aerosols. 168 As the polar maximum of the aerosol concentration increases the global mean 169 temperature increases with concomitant increase in global mean precipitation as implied by the 170 linear relationship in Figure 1b. One of the polar maximum SRM simulations (Polar3) almost 171 offsets the changes in global mean precipitation but it has a residual warming of 0.4°C. Our 172 results are broadly in agreement with other modeling studies: in an Arctic geoengineering study 173 (Caldeira and Wood, 2008) with reduced solar constant only over arctic, residual global mean 174 warming and enhancements of global precipitation are found. 175 In contrast, as magnitude of the tropical maximum concentration increases both global 176 mean temperature and precipitation decreases. One of the Tropics cases (Tropics1) where the 177 temperature change is nearly zero shows a reduction in the global mean precipitation. The 178 reduction in precipitation in our “Tropics” simulations are consistent with observed decline in

7

179 precipitation over land, runoff and river discharge into the ocean following the tropical volcanic 180 eruption Mount Pinatubo (15°N) in 1991 (Trenberth and Dai, 2007). Our “Tropics” simulations 181 can be compared to Mount Pinatubo eruption because the distribution of aerosols in „Tropics‟ 182 simulations have reasonable resemblance to the distribution of volcanic aerosols after few weeks 183 of the eruption (the volcanic aerosols occupied a latitude band of 20o S to 30o N (McCormick et 184 al., 1995)). Interestingly, we find that in none of the geoengineering scenarios considered in this 185 study, changes in global mean surface temperature and precipitation can be offset simultaneously 186 over either land or ocean. We also notice that the hydrological sensitivity (% change in 187 precipitation per unit change in temperature) is almost same over both land and ocean (Figure 188 1b). Here, we have defined the hydrological sensitivity over land (ocean) as the ratio of change in 189 land (ocean) averaged precipitation to change in land (ocean) averaged surface temperature. 190 We find that the prescribed aerosols with different latitudinal distributions along with

191 doubled CO2 concentrations (geoengineering simulations) lead to different global mean forcings 192 (Figure 1c and 1d). Since there are linear relationships between radiative forcing and the changes 193 in global mean temperature (Figure 1c) and between temperature and precipitation changes 194 (Figure 1b), we find a linear relationship between radiative forcing and precipitation changes 195 (Figure 1d). The Polar geoengineering scenarios have positive residual radiative forcing while 196 the Tropics scenarios have negative residual forcing because solar forcing is less effective over 197 the poles relative to the tropics (Figure 1c). This is further confirmed in Figure 2 which shows 198 that the Polar cases have smaller increase in planetary albedo compared to the Tropics cases. The 199 radiative forcing associated with planetary albedo changes drive the temperature changes and 200 thus the Polar cases have lower albedo changes relative to the uniform case and hence are 201 warmer and wetter while opposite is true for Tropics cases. 202 The variation of global mean surface temperature and precipitation with global mean 203 radiative forcing (Figure 1c and 1d) shows that as the maximum aerosol concentration over the 204 poles increases (Polar1 to Polar6) the residual forcing increases and hence the global mean 205 temperature and precipitation increase. Similarly, as the maximum aerosol concentration over 206 the equator increases (Tropics1 to Tropics3), an opposite variation is noticed. 207 To further investigate the degree of departure of the different geoengineering simulations 208 from the control, we calculate the root mean square difference between the spatial patterns in 209 geoengineered climates and the control climate and normalize this root mean square difference

8

210 by the standard deviation of the control scenario (NRMSD). A value less than 1 for NRMSD 211 would suggest that the geoengineered climate is indistinguishable from the control climate. 212 Further, the geoengineering simulation with the smallest value for this quantity is the one that is 213 closest to the control. In our study, we find that the NRMSD for temperature increases as the 214 maximum concentration of aerosols at the poles increases and the NRMSD for precipitation 215 increases as tropical maximum is increased (Figure 3). When all grids in the latitude-longitude 216 domain are considered the NRMSD (Figure 3a) shows large variations: 0.40 to 1.4 for surface 217 temperature and 0.25 to 0.40 for precipitation. In case of NRMSD for zonal means (Figure 3b), 218 the spread is relatively less: 0.30 to 0.95 for surface temperature and 0.27 to 0.38 for 219 precipitation. The uniform case has the least distance from the origin in Figure 3, suggesting that 220 it has the least NRMSD if the objective is to minimize root mean square difference in both 221 temperature and precipitation simultaneously. 222 223 3.2 Precipitation and Temperature Response in Tropics and Poles 224 225 The change in zonal-mean surface temperature between the geoengineering cases and the 226 control case (1xCO2) show, similar to changes in global annual mean values, a monotonic 227 increase at each latitude with increased polar weighting (Figure 4a). We notice a similar 228 monotonic increase in zonal-mean land and zonal-mean ocean surface temperature (Figures 5a 229 and 5b). Further, we find that almost all geoengineering simulation show residual high latitude 230 warming. In the Tropics cases, we find smaller residual warming in the high latitudes and cooler 231 tropics. Similar to temperature changes, the change in zonal-mean precipitation between the 232 geoengineering cases and the control case show a monotonic increase at each latitude with 233 increased polar weighting (Figures 4b, 5c and 5d).We find large changes in precipitation in the 234 tropics which is likely to be seen as shifts in the intertropical convergence zone (ITCZ) but closer 235 examination (Figure 6) shows that the position of ITCZ remains the same in all the cases and the 236 monotonic increase in precipitation with poleward weighting is clearly seen. The changes in 237 zonal mean precipitation minus evaporation (water budget) are similar to changes in zonal mean 238 precipitation (Figures 4c, 5e and 5f).

239 Figure 7 shows the spatial pattern of radiative forcing in selected simulations: 2xCO2, 240 Uniform, Polar3, Tropics1, Polar6 and Tropics3 cases. We notice that the radiative forcing in the

9

241 2xCO2 case is significant over the whole globe but not significant in most regions in the 242 geoengineering cases. The radiative forcing is positive in most locations in Polar cases. In the 243 Tropics cases, the forcing is negative in the tropical regions and positive in polar regions.

244 In the 2xCO2 case, both temperature and precipitation changes are large and significant 245 over the whole globe (Figure 8). The temperature increase over poles is much larger than in the 246 tropics, in agreement with previous studies (Caldeira and Wood, 2008; Lunt et al., 2008; 247 Matthews and Caldeira, 2007; Robock et al., 2008; Rasch et al., 2008b). The uniform 248 geoengineering case (Uniform) shows mitigation in the temperature with reduced precipitation

249 relative to 1xCO2. This is because of the different nature of the CO2 forcing and solar forcing:

250 solar forcing is larger in the tropics and smaller in the poles whereas the CO2 forcing is uniform 251 over the whole globe. In Polar3 case, the change in precipitation is largely offset but there is 252 significant warming over large regions. However, temperature is largely offset in Tropics1 but 253 there is decrease in precipitation relative to the uniform distribution case. The last four panels of 254 Figure 8 shows the extreme cases; the case with largest polar weighting (Polar6) significantly 255 warms the planet while the case with largest tropical weighting (Tropics3) overcools the planet 256 with large reduction in precipitation. 257 258 4. Discussion and Conclusions 259 260 In this study, for a fixed total amount of sulfate aerosols which when distributed

261 uniformly nearly offsets the global mean temperature change from a doubling of CO2, there is a 262 residual cooling when the aerosol concentration is maximized near the tropical regions and 263 warming when concentration is maximized near the polar regions (Figure 1c). Consequent 264 changes in global mean precipitation are simulated as dictated by the hydrological sensitivity of 265 the model (Figure 1b). Our result that the global mean precipitation is reduced when aerosol 266 concentration is maximized at the equator is in agreement with a recent study that shows a drastic 267 reduction in tropical rainfall when aerosol concentration is maximum in the tropics (Ferraro et al. 268 2014). We also observe a similar monotonic increase in precipitation intensity as the maximum 269 aerosol concentration over the poles increases (Figure 9, 10 and 11). The increases are of the 270 order of 10% for low intensity (5th percentile) and 2-3% for large intensity (99th percentile) 271 between the extreme cases (Tropics3 and Polar6). In order to confirm that global mean radiative

10

272 forcing is sufficient to infer the global mean climate change we performed four additional 273 geoengineering simulations with total amount of aerosols varied (10Mt, 11Mt, 13Mt, and 14Mt) 274 for the uniform distribution case. We find that the global mean temperature and precipitation 275 changes follow the changes in global mean forcing (Figure 12) for this set of simulations too. 276 In agreement with earlier studies (e.g., Bala et al., 2008), we find that both temperature 277 and precipitation changes cannot be offset simultaneously. In agreement with this, not only in a 278 simulation with uniform distribution but in all the geoengineering simulation with different 279 latitudinal distribution (that is, even with non-uniform distribution of solar insolation reduction), 280 we find that it is not possible to offset both temperature and precipitation changes 281 simultaneously. The latitudinal distribution which offsets the warming leads to a drier climate 282 while the distribution which offsets the precipitation results in a relatively warmer world (note 283 that Bala et al. (2008) used a uniform solar insolation reduction). For a fixed total amount of 284 aerosols but with different latitudinal distribution it is possible to achieve a range of global mean 285 radiative forcing and thus a range of climate states.

286 Our findings should be viewed in the light of the limitations and uncertainties involved in 287 this study. Our simulations are highly idealized as we have prescribed sulfate aerosol (to reduce 288 the solar insolation) instead of injecting and transporting them. We have prescribed a fixed 289 particle size distribution but particle size distribution would evolve with time and is shown to be 290 important in precisely estimating the effects on different climate variables (Rasch et al., 2008b). 291 Some modeling studies (Robock et al., 2008) have injected aerosol precursors into the 292 stratosphere with fixed particle size distribution while other studies (Heckendorn et al., 2009; 293 Pierce et al., 2010; Niemeier et al., 2010; Hommel and Graf, 2011; English et al., 2012) have 294 demonstrated the importance of including the microphysics of particle growth. Further, we have 295 focused our investigation primarily on global mean climate while several other studies (e.g., 296 Robock et al., 2008; Irvine et al., 2010; Ricke et al., 2010) focused on regional disparities.

297 In this study, we have not considered the consequences of detailed stratospheric dynamics 298 and sulfate aerosol chemistry on the ozone layer (Tilmes et al., 2009). Our model lacks a 299 dynamic ocean and sea ice components, and thus the effects of deep ocean circulation are not 300 modeled here. Further, in this model an interactive land carbon cycle is not included and hence 301 the impact of changes in the diffuse fraction of surface solar radiation due to stratospheric

11

302 aerosols could not be investigated. We intend to use a later version of the model that includes 303 carbon cycle to investigate the impacts of altered diffuse radiation in a future study. However, we 304 believe our results on temperature and precipitation is so fundamental that they would be 305 unchanged when additional components and feedbacks are included.

306 In summary, for a fixed total mass of aerosols, we find that the global mean climate is 307 warmer and wetter when aerosol concentration is maximum over the poles relative to the uniform 308 distribution case (which offsets global mean temperature change) because the global mean 309 residual radiative forcing is positive in these cases when compared to the uniform case. The 310 opposite is true when aerosol concentration is maximum in the tropics. Further, our study clearly 311 indicates that knowledge of global mean radiative forcing, not the details of latitudinal 312 distribution of aerosols, is sufficient to infer the global mean climate change. 313 314 Acknowledgements. Financial support for A. Modak was provided by the Divecha Centre for 315 Climate Change, Indian Institute of Science. We thank the Supercomputer Education and 316 Research Centre, Indian Institute of Science for providing the computational resources. 317 318 References 319 1. Allen, M.R. and Ingram, W. J.: Constraints on future changes in climate and the hydrologic 320 cycle, Nature, 419, 224-232, doi:10.1038/nature01092, 2002. 321 2. Andrews, T., Forster, P. M., and Gregory, J. M.:A surface energy perspective on climate 322 change, J. Climate., 22, 2557-2570, doi: 10.1175/2008JCLI2759.1, 2009. 323 3. Bala, G., Caldeira K., and Nemani, R.: Fast versus slow response in climate change: 324 implications for the global hydrological cycle, Clim. Dynam. 35, 423-434, doi: 325 10.1007/s00382-009-0583-y, 2009. 326 4. Bala, G., Duffy P. B., and Taylor, K. E.: Impact of geoengineering schemes on the global 327 hydrological cycle, P. Natl. Acad. Sci. USA, 105, 7664-7669, doi:10.1073/pnas.0711648105, 328 2008. 329 5. Ban-Weiss, G. A. and Calderia, K.: Geoengineering as an Optimization Problem, Environ. 330 Res. Lett., 5,034009, doi:10.1088/1748-9326/5/3/034009, 2010. 331 6. Betts, R. A., Boucher, O., Matthew, C., Cox, P. M., Falloon, P. D., Gedney, N., Hemming, D. 332 L., Huntingford, C., Jones, C. D., Sexton, D. M. H., and Webb, M. J.: Projected increase in 333 continental runoff due to plant responses to increasing carbon dioxide, Nature, 448, 1037– 334 1041, doi:10.1038/nature06045, 2007. 335 7. Budkyo, M.I.: The Earth's Climate Past and Future, Academic, 1982. 336 8. Caldeira, K. and Wood, L.: Global and Arctic climate engineering: Numerical model studies, 337 Philos. T. Roy. Soc. A, 366, 4039-4056, doi: 10.1098/rsta.2008.0132, 2008.

12

338 9. Cao, L., Bala, G., and Caldeira, K.: Climate response to changes in atmospheric carbon 339 dioxide and solar irradiance on the time scale of days to weeks, Environ. Res. Lett.,7, 340 034015, doi:10.1088/1748-9326/7/3/034015, 2012. 341 10. Cao, L., Bala, G., Caldeira, K., Nemani, R., and Ban-Weiss, G. A.: Importance of carbon 342 dioxide physiological forcing to future climate change, P. Natl. Acad. Sci. 343 USA,doi:10.1073/pnas.0913000107, 2010. 344 11. Collins, W. D., Rasch, P. J., Boville, B. A., Hack, J. J., McCaa, J. R., Williamson, D. L., 345 Kiehl, J. T., Briegleb, B., Bitz, C., Lin, S-J., Zhang, M., and Dai, Y.: Description of the 346 NCAR community atmosphere model (CAM 3.0) NCAR Tech. Rep. NCAR/TN-464+STR 347 National Center for Atmospheric Research Boulder CO 226pp, 2004. 348 12. Crutzen, P.J.: Albedo enhancement by stratospheric sulfur injections: A contribution to 349 resolve a policy dilemma?, Climatic Change, 77, 211-220, doi: 10.1007/s10584-006-9101-y, 350 2006. 351 13. Early, J.T.: Space-based solar shield to offset greenhouse effect, Journal of the British 352 Interplanetary Society, 42, 567-569, 1989. 353 14. English, J. M., Toon, O. B., and Mills, M. J.: Microphysical simulations of sulfur burdens 354 from stratospheric sulfur geoengineering, Atmos. Chem. Phys.,12, 4775-4793, 355 doi:10.5194/acp-12-4775-2012, 2012. 356 15. Ferraro, A. J., Highwood, E. J., and Charlton-Perez, A. J.: Weakened tropical circulation and 357 reduced precipitation in response to geoengineering. Environ. Res. Lett., 9, 014001 doi: 358 10.1088/1748-9326/9/1/014001, 2014. 359 16. Govindasamy, B. and Caldeira, K.: Geoengineering Earth‟s radiation balance to mitigate 360 CO2-induced climate change, Geophys. Res. Lett., 27, 2141–2144, 2000. 361 17. Hansen, J., Sato, M., and Ruedy, R.: Radiative forcing and climate response J. Geophys. Res., 362 102(D6) 6831–6864, 1997. 363 18. Heckendorn, P., Weisenstein, D., Fueglistaler, S., Luo, B. P., Rozanov, E., Schraner, M., 364 Thomason, L. W., and Peter, T.: The impact of geoengineering aerosols on stratospheric 365 temperature and ozone, Environ. Res. Lett., 4, 045108, doi: 10.1088/1748-9326/4/4/045108, 366 2009. 367 19. Hommel, R. and Graf, H. F.: Modelling the size distribution of geoengineered stratospheric 368 aerosols, Atmos. Sci. Lett., 12, 168-175, doi: 10.1002/asl.285, 2011. 369 20. Irvine, P. J., Ridgwell, A., and Lunt, D. J.: Assessing the regional disparities in 370 geoengineering impacts, Geophys. Res. Lett., 37, L18702, doi: 10.1029/2010GL044447, 371 2010. 372 21. IPCC 2007 Climate Change 2007: The Physical Science Basis Contribution of Working 373 Group 1 to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change 374 ed S Solomon, D Qin, M Manning, Z Chen, M Marquis, K B Averyt, M Tignor and H L 375 Miller (Cambridge & New York: Cambridge University Press) 996 pp. 376 22. Jones, A., Haywood, J., Boucher, O., Kravitz, B., and Robock, A.: Geoengineering by 377 stratospheric SO2 injection: Results from the Met Office HadGEM2 climate model and 378 comparison with the Goddard Institute for Space Studies ModelE, Atmos. Chem. Phys. 379 Discuss., 10, 7421-7434, doi: 10.5194/acpd-10-7421-2010, 2010. 380 23. Keith, D. W.: Geoengineering the climate: history and prospect, Annu. Rev. Energ. Env., 25, 381 245–284, 2000.

13

382 24. Lunt, D. J., Ridgwell, A., Valdes, P. J., and Seale, A.: „Sunshade world‟: a fully coupled 383 GCM evaluation of the climatic impacts of geoengineering, Geophys. Res. Lett., 35, L12710, 384 doi: 10.1029/2008GL033674, 2008. 385 25. Matthews, H. D. and Caldeira, K.: Transient climate-carbon simulations of planetary 386 geoengineering, P. Natl Acad. Sci. USA, 104, 9949–9954, doi: 10.1073/pnas.0700419104, 387 2007. 388 26. MacMartin, D. G., Keith, D.W., Kravitz, B., and Caldeira, K.: Management of trade-offs in 389 geoengineering through optimal choice of non-uniform radiative forcing, Nature Climate 390 Change, doi: 10.1038/nclimate1722, 2012. 391 27. McCormick, M. P., Thomason, L. W., and Trepte, C. R.: Atmospheric effects of the Mt. 392 Pinatubo eruption, Nature, 373, 399–404, 1995. 393 28. Niemeier, U., Schmidt, H., and Timmreck, C.: The dependency of geoengineered sulfate 394 aerosol on the emission strategy, Atmos. Sci. Lett., 12,189-194, doi: 10.1002/asl.304, 2010. 395 29. Pierce, J. R., Weisenstein, D. K., Heckendorn, P., Peter, T., and Keith, D.W.: Efficient 396 formation of stratospheric aerosol for climate engineering by emission of condensible vapor 397 from aircraft, Geophys. Res. Lett., 37, L18805, doi: 10.1029/2010GL043975, 2010. 398 30. Rasch, P. J., Tilmes, S., Turco, R. P., Robock, A., Oman, L., Chen, C.C., Stenchikov, G.L., 399 and Garcia, R. R.: An overview of geoengineering of climate using stratospheric 400 sulphateaerosols, Philos. T. Roy. Soc. A, 366, 4007–4037, doi: 10.1098/rsta.2008.0131, 401 2008a. 402 31. Rasch, P. J., Crutzen, P. J., and Coleman, D. B.: Exploring the geoengineering of climate 403 using stratospheric sulfate aerosols: The role of particle size, Geophys. Res. Lett., 35, 404 L02809, doi: 10.1029/2007GL032179, 2008b. 405 32. Robock, A., Oman, L., and Stenchikov, G. L.: Regional climate responses to geoengineering 406 with tropical and Arctic SO2 injections, J. Geophys. Res., 113, D16101, doi: 407 10.1029/2008JD010050, 2008. 408 33. Ricke, K. L., Morgan, M. G., and Allen, M. R.: Regional climate response to solar-radiation 409 management, Nature Geosci., 3, 537 – 541, 2010. 410 34. Shepherd, J., Caldeira, K., Haigh, J., Keith, D., Launder, B., Mace, G., MacKerron, G., Pyle, 411 J., Rayner, S., Redgwell, C., and Watson, A.: Geoengineering the climate: science, 412 governance and uncertainty, The Royal Academy, 2009. 413 35. Tilmes, S., Garcia, R. R., Kinnison, D. E., Gettelman, A., and Rasch, P. J.: Impact of 414 geoengineered aerosols on the troposphere and stratosphere, J. Geophys. Res., 114, D12305, 415 doi: 10.1029/2008JD011420, 2009. 416 36. Tilmes et al.: The hydrological impact of geoengineering in the Geoengineering Model 417 Intercomparison Project (GeoMIP). J. Geophys. Res. Atmos., 118, 11,036-11,058, doi: 418 10.1002/jgrd.50868, 2013. 419 37. Trenberth, K.E. and Dai, A.: Effects of Mount Pinatubo volcanic eruption on the hydrological 420 cycle as an analog of geoengineering, Geophys. Res. Lett., 34, L15702, doi: 421 10.1029/2007GL030524, 2007. 422 38. Wigley, T. M. L.: A combined mitigation/geoengineering approach to climate stabilization, 423 Science, 314, 452-454, doi: 10.1126/science.1131728, 2006.

14

424 Figure Captions 425 426 Figure 1: (a) Latitudinal profiles of sulfate aerosol concentration in the SRM geoengineering 427 experiments. Polar1-6 have maximum concentration over the poles and Tropics1-3 have 428 maximum at the equator. (b) Surface temperature change (K) vs precipitation change (%) relative

429 to the 1xCO2 case from slab ocean simulations ( global mean values - squares, land mean values 430 - stars, ocean mean values - triangles) . There is warming in all Polar cases relative to the uniform 431 case and a concomitant increase in precipitation. Opposite is the case for Tropics cases. None of 432 the regression lines pass through origin; temperature and precipitation cannot be offset 433 simultaneously. In the case of land and ocean, ∆TS and ∆P represent the averages over the 434 respective domain. (c) Radiative forcing (RF) vs surface temperature change. Polar cases have 435 larger forcing relative to the uniform case and hence are warmer while opposite is true for 436 Tropics cases. (d) Radiative forcing vs % precipitation change. Precipitation increases with 437 residual RF (i.e. with increase in polar weighting) while decreases with increase in tropical 438 weighting. In (b), (c) and (d) the horizontal and vertical bars represent the standard error of the 439 respective variables which are calculated from the last 30 years of 60-year SOM simulations 440 while in case of radiative forcing it is calculated from the last 20 years of 30-year fixed-SST 441 simulations. 442 443 Figure 2: Change in planetary albedo in fixed-SST vs surface temperature change in slab ocean 444 geoengineering simulations. The radiative forcing associated with albedo changes drive the 445 temperature changes. Polar cases have lower albedo changes relative to the uniform case and 446 hence are warmer and wetter while opposite is true for Tropics cases. The horizontal and vertical 447 bars represent the standard error of the respective variables; temperature standard errors are 448 calculated from the last 30 years of 60-year SOM simulations while albedo standard errors are 449 calculated from the last 20 years of 30-year fixed-SST simulations. 450 451 452 Figure 3: Normalized root mean square difference (NRMSD) of surface temperature and 453 precipitation between geoengineering and control simulation normalized by respective standard 454 deviations computed for the global domain (top panel) and for the zonal averages (bottom panel).

15

455 The annual means of the last 30 years of the 60-year control simulation are used to estimate the 456 standard deviation. Simulation nearest to x-axis represents the best precipitation mitigating 457 scenario while the one closest to y-axis represents the best surface temperature mitigating 458 scenario. Scenarios with maximum aerosol concentrations at the poles have larger NRMSD in 459 temperature and conversely simulations with maximum at the equator have larger NRMSD in 460 precipitation. 461 462 Figure 4: Zonal means of change in surface temperature (ΔTS), precipitation (ΔP) and 463 precipitation minus evaporation (ΔPmE). (a) Zonal mean ΔTS increases monotonically with 464 increase in maximum concentrations over the poles and decreases with increase in tropical 465 maxima. (b) Zonal mean ΔP: polar maximum causes enhanced precipitation. (c) Zonal mean 466 ΔPmE; polar maximum causes enhanced precipitation minus evaporation. Results shown are 467 averages of the last 30 years of 60-year simulations. 468 469 Figure 5: Changes in zonal mean surface temperature (ΔTS), precipitation (ΔP) and 470 precipitation minus evaporation (ΔPmE) over ocean (left panels) and land (right panels). (a) and 471 (b): Zonal mean ΔTS increases monotonically with increase in the magnitude of maximum 472 concentration of aerosols over poles and decreases with increase in the magnitude of tropical 473 maximum. (c) and (d): polar maximum causes enhanced precipitation. (e) and (f): polar 474 maximum causes enhanced precipitation minus evaporation. Results shown are averages of the 475 last 30 years of 60-year simulations. 476 477 Figure 6: Zonal mean precipitation over the globe. The position of intertropical convergence 478 zone (ITCZ) remains the same in all the geoengineering cases. The zonal mean precipitation 479 decreases monotonically over the equator as the global mean radiative forcing increases. Results 480 shown are averages of the last 30 years of 60-year simulations. 481

482 Figure 7: Spatial pattern of radiative forcing in the 2xCO2 , uniform, and some Polar and Tropics 483 geoengineering scenarios. In the Unifrom and Tropical cases, there is a residual positive forcing 484 in the high latitudes and negative forcing in the low latitudes indicating an inexact compensation. 485 Hatching indicates the region where the changes are significant at 1% level. Significance level

16

486 was estimated by Student‟s t test. Results shown are averages of the last 20 years of 30-year 487 simulations with fixed sea surface temperature and sea ice fraction. 488 489 Figure 8: Changes in annual-mean surface temperature (left panels) and precipitation (right

490 panels) in the 2xCO2 , uniform, and some Polar and Tropics geoengineering scenarios relative to

491 the control (1xCO2). Hatching indicates the region where the changes are significant at 1% level. 492 Significance level was estimated using Student‟s t-test. Both surface temperature and

493 precipitation changes are large and significant everywhere in the 2xCO2 and extreme scenarios 494 (Polar6 and Tropics3). Although significant over large regions, both temperature and 495 precipitation changes are small in the Uniform case. Polar3 scenario offsets global mean 496 precipitation but not global mean temperature while Tropics1 scenario offsets global mean 497 temperature but with reduced precipitation. Results shown are averages of the last 30 years of 60- 498 year simulations. 499 500 Figure 9: Percentile values (p5, p25, median, p75, p85, p90, p95 and p99)of precipitation 501 intensity over Globe. There is a monotonic increase in precipitation for all percentile values as 502 the maximum concentration of aerosols over poles increases. Grid-level monthly mean 503 precipitation are used to calculate the percentile values.The last 30 years of 60-year simulations 504 are used for the statistics. 505 506 Figure 10: Percentile values (p5, p25, median, p75, p85, p90, p95 and p99)of precipitation 507 intensity over Land. There is a monotonic increase in precipitation for all percentile values as the 508 maximum concentration of aerosols over poles increases. Grid-level monthly mean precipitation 509 over all land points are used to calculate the percentile values.The last 30 years of 60-year 510 simulations are used for the statistics. 511 512 Figure 11: Percentile values (p5, p25, median, p75, p85, p90, p95 and p99)of precipitation 513 intensity over Ocean. There is a monotonic increase in precipitation for all percentile values as 514 the maximum concentration of aerosols over poles increases. Grid-level monthly mean 515 precipitation over all ocean points are used to calculate the percentile values.The last 30 years of 516 60-year simulations are used for the statistics.

17

517 Figure 12: (a) Radiative forcing (RF) vs surface temperature change. (b) Radiative forcing vs % 518 precipitation change for uniform distribution scenarios with 10Mt, 11Mt, 12.6Mt, 13Mt and 519 14Mt. More aerosol mass leads to negative residual radiative forcing and hence cooler and drier 520 climate, and smaller aerosol mass leads to positive residual radiative forcing and hence warmer 521 and wetter climate. In (a) and (b) the horizontal and vertical bars represent the standard error of 522 the respective variables. Results shown are averages of the last 20 years of 50-year SOM 523 simulations for temperature and precipitation while the last 20 years of 40-year fixed-SST 524 simulations are used for radiative forcing calculations.

18

525 Table 1: Description of different geoengineering experiments. Total additional mass is 12.6 Mt 526 SO4 in all the geoengineering simulations but the distribution varies.

Name of the a b Total Mass Total Mass from Total Mass Experiments (mg/m2) (mg/m2) from uniform non-uniform (Mt) component component (Mt) (Mt) Uniform 24.70 - 12.60 - 12.60 Polar1 23.52 3.19 12.00 0.60 12.60 Polar2 21.56 8.55 11.00 1.60 12.60 Polar3 19.60 13.89 10.00 2.60 12.60 Polar4 17.64 19.22 9.00 3.60 12.60 Polar5 15.68 24.56 8.00 4.60 12.60 Polar6 13.72 29.90 7.00 5.60 12.60 Tropics1 26.66 -5.34 13.60 -1.00 12.60 Tropics2 28.62 -10.67 14.60 -2.00 12.60 Tropics3 30.58 -16.02 15.60 -3.00 12.60 527 528

19

529 Figure 1

530

20

531 Figure 2

532 533

21

534 Figure 3

535

22

536 Figure 4

537 538

23

539 Figure 5

540 541

24

542 Figure 6

543 544

25

545 Figure 7

546

26

547 Figure 8

548 549

27

550 Figure 9

551 552

28

553 Figure 10

554 555

29

556 Figure 11

557 558

30

559 Figure 12

560

31