Skip to main content

Advertisement

Log in

Anthropogenic climate change for 1860 to 2100 simulated with the HadCM3 model under updated emissions scenarios

  • Published:
Climate Dynamics Aims and scope Submit manuscript

Abstract.

In this study we examine the anthropogenically forced climate response over the historical period, 1860 to present, and projected response to 2100, using updated emissions scenarios and an improved coupled model (HadCM3) that does not use flux adjustments. We concentrate on four new Special Report on Emission Scenarios (SRES) namely (A1FI, A2, B2, B1) prepared for the Intergovernmental Panel on Climate Change Third Assessment Report, considered more self-consistent in their socio-economic and emissions structure, and therefore more policy relevant, than older scenarios like IS92a. We include an interactive model representation of the anthropogenic sulfur cycle and both direct and indirect forcings from sulfate aerosols, but omit the second indirect forcing effect through cloud lifetimes. The modelled first indirect forcing effect through cloud droplet size is near the centre of the IPCC uncertainty range. We also model variations in tropospheric and stratospheric ozone. Greenhouse gas-forced climate change response in B2 resembles patterns in IS92a but is smaller. Sulfate aerosol and ozone forcing substantially modulates the response, cooling the land, particularly northern mid-latitudes, and altering the monsoon structure. By 2100, global mean warming in SRES scenarios ranges from 2.6 to 5.3 K above 1900 and precipitation rises by 1%/K through the twenty first century (1.4%/K omitting aerosol changes). Large-scale patterns of response broadly resemble those in an earlier model (HadCM2), but with important regional differences, particularly in the tropics. Some divergence in future response occurs across scenarios for the regions considered, but marked drying in the mid-USA and southern Europe and significantly wetter conditions for South Asia, in June–July–August, are robust and significant.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1.
Fig. 2a–d.
Fig. 3a–f.
Fig. 4a, b.
Fig. 5.
Fig. 6.
Fig. 7.
Fig. 8.
Fig. 9.
Fig. 10.
Fig. 11.
Fig. 12a–f.
Fig. 13a–f.
Fig. 14a–f.
Fig. 15a–f.

Similar content being viewed by others

References

  • Allen MR, Ingram WJI (2002) Constraints on future changes in climate and the hydrological cycle. Nature 419:224–232

    Google Scholar 

  • Albrecht BA (1989) Aerosols, cloud microphysics and fractional cloudiness. Science 245: 1227–1230

    Google Scholar 

  • Andres RJ, Kasgnoc AD (1998) A time-averaged inventory of subaerial volcanic sulfur emissions. J Geophys Res 103: 25,251–25,261

    Google Scholar 

  • Barnett TP, Hasselmann K, Chelliah M, Delworth T, Hegerl G, Jones P, Rasmussen E, Roeckner E, Ropelewski C, Santer B, Tett S (1999) Detection and attribution of climate change: a status report. Bull Am Meteorol Soc 12: 2631–2659

    Article  Google Scholar 

  • Boer GJ (1993) Climate change and the regulation of the surface moisture and energy budgets. Clim Dyn 8: 225–239

    Google Scholar 

  • Bohren CF (1980) Multiple scattering of light and some of its observable consequences. Am J Phys 55: 524–533

    Google Scholar 

  • Bohren CF, Huffman DR (1983) Absorption and scattering of light by small particles. John Wiley and Sons, New York, pp 530

  • Boville BA, Gent PR (1998) The NCAR Climate System Model, version one. J Clim 11: 1115–1130

    Article  Google Scholar 

  • Boville BA, Kiehl JT, Rasch PJ, Bryan FO (2001) Improvements to the NCAR CSM-1 for transient climate simulations. J Clim 14: 164–179

    Article  Google Scholar 

  • Bower KN, Choularton TW (1992) A parametrization of the effective radius of ice free clouds for use in global climate models. Atmos Res 27: 305–339

    Google Scholar 

  • Cattle H, Crossley J (1995) Modelling Arctic climate change. Philos Trans R Soc London A352: 201–213

    Google Scholar 

  • Collins M, Tett SFB, Cooper C (2001) The internal climate variability of HadCM3, a version of the Hadley Centre coupled model without flux adjustments. Clim Dyn 17: 61–81

    Article  Google Scholar 

  • Collins WJ, Stevenson DS, Johnson CE, Derwent RG (1997) Tropospheric ozone in global-scale three-dimensional model and its response to NOx emission controls. J Atmos Chem 26: 223–274

    Article  CAS  Google Scholar 

  • Collins WJ, Stevenson DS, Johnson CE, Derwent RG (1999) The role of convection in determining the budget of odd hydrogen in the upper troposphere. J Geophys Res 104: 26,927–26,941

    Article  Google Scholar 

  • Cox P, Betts R, Bunton C, Essery R, Rowntree PRR, Smith J (1999) The impact of new land surface physics on the GCM simulation of climate and climate sensitivity. Clim Dyn 15: 183–203

    Article  Google Scholar 

  • Cox PM, Betts RA, Jones CD, Spall SA, Totterdell IJ (2000) Acceleration of global warming by carbon cycle feedbacks in a 3D coupled model. Nature 408: 184–187

    CAS  PubMed  Google Scholar 

  • Cubasch U, Meehl GA, Boer GJ, Stouffer RJ, Dix M, Noda A, Senior CA, Raper S, Yap KS (2001) Projections of Future Climate Change. In: Houghton JT, Ding Y, Griggs DJ, Noguer M, van der Linden PJ, Dai X, Maskell K, Johnson CA (eds) Climate change 2001. The scientific basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, pp 881

  • Cusack S, Slingo A, Edwards JM, Wild M (1998) The radiative impact of a simple aerosol climatology on the Hadley Centre GCM. Q J R Meteorol Soc 124: 2517–2526

    Article  Google Scholar 

  • Dai A, Wigley TML, Boville BA, Kiehl JT, Buja LE (2001) Climates of the 20th and 21st centuries simulated by the NCAR Climate System Model. J Clim 14: 485–519

    Article  Google Scholar 

  • Edwards JM, Slingo A (1996) Studies with a flexible new radiation code. I: choosing a configuration for a large scale model. Q J R Meteorol Soc 122: 689–719

    Google Scholar 

  • Fitzgerald JW (1975) Approximation formulas for the equilibrium size of an aerosol particle as a function of its dry size and composition and the ambient relative humidity. J Appl Meteorol 14: 1044–1049

    Article  Google Scholar 

  • Forster PM de F, Shine KP (1997) Radiative forcing and temperature trends from stratospheric ozone changes. J Geophys Res 102: 10,841–10,855

    Google Scholar 

  • Gent PR, McWilliams JC (1990) Isopycnal mixing in ocean circulation models. J Phys Oceanogr 20: 150–155

    Article  Google Scholar 

  • Gordon C, Cooper C, Senior CA, Banks HT, Gregory JM, Johns TC, Mitchell JFB, Wood RA (2000) The simulation of SST, sea ice extents and ocean heat transports in a version of the Hadley Centre coupled model without flux adjustments. Clim Dyn 16: 147–168

    Article  Google Scholar 

  • Gregory D, Rowntree PR (1990) A mass flux convection scheme with representation of ensemble characteristics and stability dependent closure. Mon Weather Rev 118: 1483–1506

    Article  Google Scholar 

  • Gregory D, Kershaw R, Inness PM (1997) Parametrization of momentum transport by convection. II: tests in single column and general circulation models. Q J R Meteorol Soc 123: 1153–1183

    Article  Google Scholar 

  • Gregory D, Shutts GJ, Mitchell JR (1998) A new gravity wave drag scheme incorporating anisotropic orography and low level wave breaking: impact upon the climate of the UK Meteorological Office Unified Model. Q J R Meteorol Soc 124: 463–493

    Article  Google Scholar 

  • Gregory JM, Lowe JA (2000) Predictions of global and regional sea-level rise using AOGCMs with and without flux adjustment. Geophys Res Lett 27: 3069–3072

    Google Scholar 

  • Gregory JM, Church JA, Boer GJ, Dixon KW, Flato GM, Jackett DR, Lowe JA, O'Farrell SP, Roeckner E, Russell GL, Stouffer RJ, Winton M (2001) Comparison of results from several AOGCMs for global and regional sea-level change 1900–2100. Clim Dyn 18: 225–240

    Article  Google Scholar 

  • Hough AM (1991) Development of a two-dimensional global tropospheric model: model chemistry. J Geophys Res 96: 7325–7326

    CAS  Google Scholar 

  • Houghton JT, Jenkins GJ, Ephraums JJ (eds) (1990) Climate change: the IPCC scientific assessment. Cambridge University Press, Cambridge, UK, pp 365

    Google Scholar 

  • Houghton JT, Meira Filho LG, Bruce J, Lee H, Callander BA, Haites E, Harris N, Maskell K (eds) (1996) Climate change 1995: the science of climate change. Cambridge University Press, Cambridge, UK, pp 572

    Google Scholar 

  • Houghton JT, Ding Y, Griggs DJ, Noguer M, van der Linden PJ, Dai X, Maskell K, Johnson CA (eds) (2001) Climate change 2001. The scientific basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, pp 881

  • Hulme M, Jenkins GJ (1998) Climate change scenarios for the United Kingdom. UK Climate Impacts Programme Tech Note 1, (61 + xv)pp

  • Hulme M, Mitchell JFB, Ingram WJ, Johns TC, New M, Viner D (1999) Climate change scenarios for global impact studies. Global Environ Change 9: S3–S19

    Google Scholar 

  • Johns TC, Carnell RE, Crossley JF, Gregory JM, Mitchell JFB, Senior CA, Tett SFB, Wood RA (1997) The Second Hadley Centre Coupled Ocean–Atmosphere GCM: model description, spinup and validation. Clim Dyn 13: 103–134

    Article  Google Scholar 

  • Johnson CE, Collins WJ, Stevenson DS, Derwent RG (1999) Relative roles of climate and emissions changes on future tropospheric oxidant concentrations. J Geophys Res 104: 18,631–18,645

    Google Scholar 

  • Johnson CE, Stevenson DS, Collins WJ, Derwent RG (2001) Role of climate feedback on methane and ozone studied with a coupled ocean–atmosphere-chemistry model. Geophys Res Lett 28: 1723–1726

    CAS  Google Scholar 

  • Jones A, Roberts DL, Slingo A (1994) A climate model study of indirect radiative forcing by anthropogenic sulphate aerosols. Nature 370: 450–453

    CAS  Google Scholar 

  • Jones A, Roberts DL, Woodage MJ (1999) The indirect effects of anthropogenic sulphate aerosol simulated using a climate model with an interactive sulphur cycle. Hadley Centre Technical Note 14, Met Office, Bracknell, UK, pp 39

  • Jones A, Roberts DL, Woodage MJ, Johnson CE (2001) Indirect sulphate aerosol forcing in a climate model with an interactive sulphur cycle. J Geophys Res 106 D17: 20,293–20,310

    Google Scholar 

  • Joos F, Bruno M, Fink R, Siegenthaler U, Stocker T, Le Quéré C, Sarmiento JL (1996) An efficient and accurate representation of complex oceanic and biospheric models of anthropogenic carbon uptake. Tellus 48B: 397–417

    CAS  Google Scholar 

  • Kattenberg A, Georgi F, Grassl H, Meehl GA, Mitchell JFB, Stouffer RJ, Tokioka T, Weaver AJ, Wigley TML (1996) Climate models – projections of future climate. In: Houghton JT, Meira Filho LG, Callander BA, Harris N, Kattenberg A, Maskell K (eds) Climate change 1995. The science of climate change. Cambridge University Press, Cambridge, UK, pp 572

  • Kettle AJ, Andreae MO, Amouroux D, Andreae TW, Bates TS, Berresheim H, Bingemer H, Boniforti R, Helas G, Leck C, Maspero M, Matrai P, McTaggart AR, Mihalopoulos N, Nguyen BC, Novo A, Putaud JP, Rapsomanikas S, Roberts G, Schebeske G, Sharma S, Simo R, Staubes R, Turner S, Uher G (1999) A global database of sea surface dimethylsulfide (DMS) measurements and a procedure to predict sea surface DMS as a function of latitude, longitude and month. Global Biogeochem Cycles 13: 399–444

    CAS  Google Scholar 

  • Kraus EB, Turner JS (1967) A one dimensional model of the seasonal thermocline. Part II. Tellus 19: 98–105

    Google Scholar 

  • Levitus S, Boyer TP (1994) World Ocean Atlas 1994, vol 4: temperature. NOAA/NESDIS E/OC21, US Department of Commerce, Washington DC, pp 117

    Google Scholar 

  • Levitus S, Burgett R, Boyer TP (1995) World Ocean Atlas 1994, Vol 3: salinity. NOAA/NESDIS E/OC21, US Department of Commerce, Washington DC, pp 99

    Google Scholar 

  • Li D, Shine KP (1995) A 4-dimensional ozone climatology for UGAMP models. UGAMP Internal Report 35

  • Liss PS, Merlivat L (1986) Air–sea gas exchange rates: introduction and synthesis. In: Buat-Menard P (ed) The role of air–sea exchange in geochemical cycling. Reidel, Dordrecht, Netherlands

  • Mahlman JD (2001) Projections of future changes in climate. In: MacCracken MC, Perry JS (eds) Encyclopaedia of global environmental change, vol 1. John Wiley, Chichester, UK, pp 688

  • Manabe S, Wetherald RT (1975) The effect of doubling CO2 concentrations on the climate of a GCM. J Atmos Sci 32: 3–15

    Article  CAS  Google Scholar 

  • Martin GM, Johnson DW, Spice A (1994) The measurement and parametrization of effective radius of droplets in warm stratocumulus clouds. J Atmos Sci 51: 1823–1842

    Article  Google Scholar 

  • Milton SF, Wilson CA (1996) The impact of parametrized sub-grid scale orographic forcing on systematic errors in a global NWP model. Mon Weather Rev 124: 2023–2045

    Article  Google Scholar 

  • Mitchell JFB, Johns TC (1997) On modification of global warming by sulphate aerosols. J Clim 10: 245–267

    Article  Google Scholar 

  • Mitchell JFB, Johns TC, Senior CA (1999) Transient response to increasing greenhouse gases using models with and without flux adjustments. Hadley Centre Technical Note 2, Met Office, Bracknell, 15pp + 20 figs

  • Mitchell JFB, Karoly DJ, Hegerl GC, Zwiers FW, Allen MR, Marengo J (2001) Detection of climate change and attribution of causes. In: Houghton JT, Ding Y, Griggs DJ, Noguer M, van der Linden PJ, Dai X, Maskell K, Johnson CA (eds) Climate change 2001. The scientific basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, USA, pp 881

  • Murphy JM, Mitchell JFB (1995) Transient response of the Hadley Centre coupled ocean atmosphere model to increasing carbon dioxide. Part II: spatial and temporal structure of response. J Clim 8: 57–80

    Article  Google Scholar 

  • Nakicenovic N, and 27 co-authors (2000) IPCC Special report on emission scenarios. Cambridge University Press, Cambridge, UK, pp 599

  • Noda A, Yukimoto S, Maeda S, Uchiyama T, Shibata K, Yamaki S (2001) A New Meterological Research Institute Coupled GCM (MRI-CGCM2) Transient response to greenhouse gas and aerosol scenarios. Centre for Global Environmental Research Report, NIES, Japan, pp 63

  • Nozawa T, Emori S, Numagati A, Tsushima Y, Takemura T, Nakajima T, Abe-Ouchi A, Kimoto M (2001) Projections of future climate change in the 21st century simulated by the CCSR/NIES CGCM under the IPCC SRES scenarios. In: Matsuno T, Kida H (eds) Present and future of modelling global environment change – toward integrated modelling. Terra Scientific, Tokyo, Japan, pp 15–28

  • Olivier JGJ, Bouwman AF, van der Maas CWM, Berdowski JJM, Veldt C, Bloos JPJ, Visschedijk AJH, Zandveld PYK, Haverlag JL (1996) Description of EDGAR Version 2.0: a set of global emission inventories of greenhouse gases and ozone-depleting substances for all anthropogenic and most natural sources on a per country basis and on a 1 deg x 1 deg grid. National Institute of Public Health and the Environment, The Netherlands. RIVM Rep 771060 002, TNO-MEP Rep R96/119, pp 141

  • Orn G, Hansson U, Rodhe H (1996) Historical worldwide emissions of anthropogenic sulfur: 1860–1985. International Meteorological Institute in Stockholm, Rep CM-91, Stockholm University, 5pp + 15 plates

    Google Scholar 

  • Pacanowski RC, Philander SG (1981) Parametrization of vertical mixing in numerical models of tropical oceans. J Phys Oceanogr 11: 1443–1451

    Article  Google Scholar 

  • Parker DE, Jones PD, Folland CK, Bevan A (1994) Interdecadal changes of surface temperature since the late nineteenth century, J Geophys Res 99: 14,373–14,399

    Google Scholar 

  • Penner JE, Andreae M, Annegarn H, Barrie L, Feichter J, Hegg D, Jayaraman A, Leaitch R, Murphy D, Nganga J, Pitari G (2001) Aerosols, their direct and indirect effects. In: Houghton JT, Ding Y, Griggs DJ, Noguer M, van der Linden PJ, Dai X, Maskell K, Johnson CA (eds) Climate change 2001. The scientific basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, pp 881

  • Platnick S, Twomey S (1994) Determining the susceptibility of cloud albedo to changes in droplet concentration with the advanced very high resolution radiometer. J Appl Meteorol 33: 334–347

    Article  Google Scholar 

  • Pope VD, Gallani ML, Rowntree PR, Stratton RA (2000) The impact of new physical parametrizations in the Hadley Centre climate model: HadAM3. Clim Dyn 16: 123–146

    Article  Google Scholar 

  • Prather M, Ehhalt D, Dentener F, Derwent R, Dlugokencky E, Holland E, Isaksen I, Katima J, Kirchhoff V, Matson P, Midgley P, Wang M (2001) Atmospheric chemistry and greenhouse gases. In: Houghton JT, Ding Y, Griggs DJ, Noguer M, van der Linden PJ, Dai X, Maskell K, Johnson CA (eds) Climate change 2001. The scientific basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, pp 881

  • Ramanathan V, Dickinson RE (1979) The role of stratospheric ozone in the zonal and seasonal radiative energy balance of the Earth–troposphere system. J Atmos Sci 36: 1084–1104

    CAS  Google Scholar 

  • Ramaswamy V, Boucher O, Haigh J, Hauglustaine D, Haywood J, Myhre G, Nakajima T, Shi GY, Solomon S (2001) Radiative forcing of climate change. In: Houghton JT, Ding Y, Griggs DJ, Noguer M, van der Linden PJ, Dai X, Maskell K, Johnson CA (eds) Climate change 2001. The scientific basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, pp 881

  • Rayner NA, Horton EB, Parker DE, Folland CK, Hackett RB (1996) Version 2.2 of the Global sea-Ice and Sea Surface Temperature dataset, 1903–1994. Climate Research Technical Note 74, Met Office, Bracknell, UK, 21pp + 13 figs

  • Roeckner E, Bengtsson L, Feichter J, Lelieveld J, Rodhe H (1999) Transient climate change simulations with a coupled atmosphere–ocean GCM including the tropospheric sulfur cycle. J Clim 12: 3004–3032

    Article  Google Scholar 

  • Roether W, Roussenov VM, Well R (1994) A tracer study of the thermohaline circulation of the eastern Mediterranean. In: Malanotte-Rizzoli P, Robinson AR (eds) Ocean processes in climate dynamics: global and Mediterranean example. Kluwer Academic, pp 371–394

    Google Scholar 

  • Santer BD, Wigley TML, Barnett TP, Anyamba E (1996) Detection of Climate Change and Attribution of Causes. In: Houghton JT, Meira Filho LG, Callander BA, Harris N, Kattenberg A, Maskell K (eds) Climate change 1995. The science of climate change. Cambridge University Press, Cambridge, UK, pp 572

  • Schimel D, and 26 co-authors (1996) Radiative forcing of climate change. In: Houghton JT, Meira Filho LG, Callander BA, Harris N, Kattenberg A, Maskell K (eds) Climate change 1995. The science of climate change. Cambridge University Press, Cambridge, UK, pp 572

  • Shine KP, Derwent RG, Wuebbles DJ, Morcrette J-J (1990) Radiative forcing of climate. In: Houghton JT, Jenkins GJ, Ephraums JJ (eds) Climate change. The IPCC scientific assessment. Cambridge University Press, Cambridge, UK, pp 365

  • Smith RNB (1990) A scheme for predicting layer clouds and their water content in a general circulation model. Q J R Meteorol Soc 116: 435–460

    Google Scholar 

  • Stendel M, Schmith T, Roeckner E, Cubasch U (2000) The climate of the 21st century: transient simulations with a coupled atmosphere–ocean general circulation model. Danmarks Klimacenter Report 00-6, Danish Meteorological Institute, Copenhagen, Denmark, pp 51

    Google Scholar 

  • Stevenson DS, Johnson CE, Collins WJ, Derwent RG, Edwards JM (2000) Future estimates of tropospheric ozone radiative forcing and methane turnover – the impact of climate change. Geophys Res Lett 27: 2073–2076

    CAS  Google Scholar 

  • Stott PA, Tett SFB, Jones GS, Allen MR, Ingram WJ, Mitchell JFB (2001) Attribution of twentieth century temperature change to natural and anthropogenic causes. Clim Dyn 17: 1–21

    CAS  Google Scholar 

  • Tett SFB, Stott PA, Allen MR, Ingram WJ, Mitchell JFB (1999) Causes of twentieth century temperature change. Nature 399: 569–572

    CAS  Google Scholar 

  • Tett SFB, Jones GS, Stott PA, Hill D, Mitchell JFB, Allen MR, Ingram WJ, Johns TC, Johnson CE, Jones A, Roberts DL, Sexton DMH, Woodage MJ (2002) Estimation of natural and anthropogenic contributions to 20th century climate temperature change. J Geophys Res 107 (D16), 10.1029/2000JD000028

  • Toon OB, Pollack JB, Khare BN (1976) The optical constants of several atmospheric aerosol species: ammonium sulfate, aluminium oxide, and sodium chloride. J Geophys Res 81: 5733–5748

    CAS  Google Scholar 

  • US National Assessment synthesis team (2000) Climate change impacts on the United States. US Global Change Programme, Cambridge University Press, pp 154

  • Visbeck M, Marshall J, Haine T, Spall M (1997) On the specification of eddy transfer coefficients in coarse resolution ocean circulation models. J Phys Oceanogr 27: 381–402

    Article  Google Scholar 

  • Wang HJ, Cunnold DM, Bao X (1996) A critical analysis of stratospheric aerosol and gas experiment ozone trends. J Geophys Res 101 D7: 12,495–12,514

    Google Scholar 

  • Williams KD, Senior CA, Mitchell JFB (2001a) Transient climate change in the Hadley Centre models: the role of physical processes. J Clim 14: 2659–2674

    Article  Google Scholar 

  • Williams KD, Jones A, Roberts DL, Senior CA, Woodage MJ (2001b) The response of the climate system to the indirect effects of anthropogenic sulfate aerosol. Clim Dyn 17: 845–856

    Article  Google Scholar 

  • Wood RA, Keen AB, Mitchell JFB, Gregory JM (1999) Changing spatial structure of the thermohaline circulation in response to atmospheric CO2 forcing in a climate model. Nature 399: 572–575

    CAS  Google Scholar 

  • Woodward S (2001) Modelling the atmospheric life cycle and radiative impact of mineral dust in the Hadley Centre climate model. J Geophys Res 106: 18,155–18,166

    Google Scholar 

  • Wright DK (1997) A new eddy mixing parametrization and ocean general circulation model. International WOCE News 26: 27–29

    Google Scholar 

  • Xie P, Arkin PA (1997) Global precipitation: a 17-year monthly analysis based on gauge observations, satellite estimates and numerical model outputs. Bull Am Meteorol Soc 78: 2539–2558

    Article  Google Scholar 

Download references

Acknowledgements.

This work was carried out under the UK Department of the Environment, Transport and Regions Climate Prediction Programme, DETR-CPP (PECD 7/12/37) and UK Meteorological Office Public Meteorological Service Research Programme (MSG 2/98) with supplementary support for some experiments from the Commission of the European Communities (contract ENV4-CT95-0102 – SIDDACLICH). Supercomputing resources for this work were funded by the DETR-CPP. We thank Denise Cresswell and Gareth Jones for running the A1FI experiment, Thomas Toniazzo for running the B1 experiment, and Chris Colloff and Tristan Tuftnell for assistance with data analysis, preparing figures and tables.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to T. C. Johns.

Appendices

Appendix A: Calibration of indirect sulfate aerosol forcing

Annual-mean distributions of sulfate aerosol were generated from HadAM3 model runs which included the sulfur cycle. Only natural emissions were used for the pre-industrial era, namely emissions of DMS (Kettle et al. 1999; Liss and Merlivat 1986) and volcanic emissions (Andres and Kasgnoc 1998). Anthropogenic emission scenarios were added to the natural emissions for the years 1860, 1900, 1950, 1975, 2000, 2050 and 2100 (see Appendix C). Successive pairs of these sulfate distributions (e.g. pre-industrial and 1860; 1860 and 1900) were then used as input to 1-year runs of HadAM3 with climatological (present-day) sea surface temperatures and sea ice. The version of HadAM3 used was identical to the atmospheric component of HadCM3 except that it used the relation of Jones et al. (1994) to predict cloud droplet number concentration (N d ) rather than the fixed N d values used in the HadCM3 control run. Concentrations of sulfate and (over the oceans) sea-salt aerosols were used as input to this relation to determine N d (Jones et al. 1999), subject to a minimum value of 5 cm–3. For each model run the shortwave radiation scheme was called twice, once with the earlier sulfate distribution and once with the later. In each case a simple measure of cloud albedo α was calculated (Bohren 1980):

$$ \alpha = \tau (1 - g)[2 + \tau (1 - g)]^{- 1} $$
(1)

This assumes conservative scattering, and in approximating the indirect effect we also assume a fixed asymmetry parameter g (=0.85) and use the fact that the optical (extinction) cross section of a particle much larger than the wavelength is twice its geometrical cross section (Bohren and Huffman 1980).

For each 1-year HadAM3 model run using a pair of sulfate scenarios, the distribution of the change in α caused by the changes in aerosol was stored and used to build up a history of the change in cloud albedo (Δα). If dα1, dα2 etc. represent the changes in cloud albedo for the sulfate distributions pairs {pre-industrial,1860}, {1860, 1900} etc. respectively, then the cloud albedo change at each "marker" year is calculated by summation:

$$\eqalign{& \Delta \alpha (1860) = d\alpha _1 \cr & \Delta \alpha (1900) = d\alpha _1 + d\alpha _2 \cr & etc.} $$

Cloud albedo changes between these dates were obtained by linear interpolation.

These cloud albedo changes were used to simulate indirect forcing in the HadCM3 runs as follows. The change in α due to the change in aerosol distribution corresponds to changing τ by a factor λ:

$$ \Delta \alpha = \lambda \tau (1 - g)[2 + \lambda \tau (1 - g)]^{- 1} - \tau (1 - g)[2 + \tau (1 - g)]^{- 1} $$
(2)

Re-arranging Eq. (1) to give an expression for τ and then substituting into Eq. (2) yields:

$$\Delta \alpha = \lambda \alpha [(1 - \alpha) + \lambda \alpha ]^{- 1} - \alpha $$
(3)

Following Platnick and Twomey (1994), we define a factor χ, the fractional change of N d corresponding to the fractional change λ in τ. As τ is inversely proportional to the cloud droplet effective radius r e , and as r e is proportional to N d –1/3 (Martin et al. 1994), then we can replace λ by χ1/3. Re-arranging Eq. (3) then gives:

$$\chi = [\Delta \alpha (\alpha - 1) - \alpha (1 - \alpha)]^3 [\alpha (\Delta \alpha + \alpha - 1)]^{- 3} $$
(4)

At each time step in the HadCM3 run the fixed values of N d and the predicted cloud water content were used to derive an initial value for r e using the parametrization of Martin et al. (1994). Cloud optical depth τ was then calculated and used to derive an initial distribution of α using Eq. (1). The distribution of Δα for the appropriate time was available from the offline history produced from the HadAM3 runs and used in conjunction with α to calculate χ using Eq. (4). Note that at this point in the model r e has only been used to determine the initial distribution of α. Before r e was used in the radiation code for a full calculation of cloud radiative properties it was modified to r e ′ using χ to approximate the indirect albedo effect:

$$r^{\prime}_e = r_e \chi ^{- 1/3} $$
(5)

It is possible for χ to be infinite, zero or negative, implying either: (1) that the decrease in cloud albedo required is greater than (χ < 0) or equal to (χ = 0) the current value of α, and in such cases the cloud was made as dark as possible by setting r e to the maximum value allowed (37 μm); or (2) that the increase in cloud albedo required to produce the desired forcing, when added to α, produces a result which is greater than (χ < 0) or equal to (χ → ∞) unity, and in such cases the cloud was made as bright as possible by setting r e to the minimum permitted value (0.35 μm). Both these factors may exert a limiting influence on the forcing actually produced in the coupled run, as may differences between the cloud distribution of the coupled run and of the atmosphere-only runs used to generate the albedo-change history. It is too computationally expensive to calculate the actual forcing experienced by the coupled model as the simulation evolves as this would require multiple calls to the (computationally intensive) radiation scheme. Consequently, 1-year portions of the simulations were re-run with extra radiation calls and diagnostics for the scenario years 1860, 1900, 1950, 1975, 2000, 2050 and 2100. The ideal forcing for these years was available from the separate HadAM3 runs used to generate the albedo-change history, and this was compared with the actual forcing produced in HadCM3 in these years.

Appendix B: Calculation of ozone trends

In the troposphere the ozone trends are constructed by means of off-line 3-D chemical transport (STOCHEM, Collins et al. 1997, 1999) model simulations for specific epochs 1990, 2030, 2060 and 2100; each simulation being run for 15 months. Intermediate results are obtained by linear interpolation in time. Initial methane concentrations are provided from an earlier run of the 2-D TROPOS model. Emissions for 1990 are based on those from the EDGAR database (Olivier et al. 1996), with future emissions taken from the SRES scenario definitions themselves. The 3-D STOCHEM model runs are also meteorologically and dynamically forced with six-hourly data from a previous HadCM3 coupled experiment which to first order reflects the climate conditions (e.g. global warming) expected at the appropriate time.

All SRES scenarios assume the Montreal protocol is followed, the Cl and Br concentrations being determined via box modelling techniques. The resulting stratospheric ozone concentrations are then scenario-independent. For the purpose of extrapolating stratospheric ozone depletion into the future, 'effective equivalent stratospheric chlorine' (EESC = Cl + 40*Br) contributions were estimated for Cl and Br compounds not already included in the 2-D model used for computing well-mixed minor GHGs by means of a simple box model driven using SRES and natural emissions where appropriate. Tropospheric mixing ratio estimates from this and the 2-D model for years from 1950 to 2100 were then combined and used to calculate stratospheric ozone as a function of height, longitude and month, assuming:

  1. 1.

    That it takes 3 years for tropospheric Cl and Br to get into the stratosphere

  2. 2.

    That ozone is depleted linearly by EESC levels above the 1974 stratospheric level, any implied negative ozone values being replaced by a minimum ozone mass mixing ratio of 10–11 for numerical reasons

Ozone loss is at its largest in 2002 and recovers subsequently as EESC returns to preindustrial values. Beyond about 2060, this should have resulted in a stabilization of stratospheric ozone at preindustrial levels again, but due to a software error ozone in fact continued to increase. This leads to an erroneous small positive net forcing of +0.17 W/m2 from stratospheric ozone at 2100 relative to preindustrial in all the scenarios.

Further details of the method and the stratospheric ozone trends up to present day is given by Tett et al. (2002). Zero ozone change was imposed at the tropopause model level, as estimated from the 1871–1990 mean tropopause height of a HadAM3 integration forced by the observed sea surface temperatures and sea ice extents from the GISST3.1 data set, an updated version of that described in Rayner et al. (1996).

Appendix C: Anthropogenic sulfur emissions

Only anthropogenic emissions were included in the experiments. For the 'past' portion of the experiments, the basic datasets we used are the revised version of the 1985 GEIA 1B dataset (released in June 1997 – see http://www.geiacenter.org/ and http://www.ortech.ca/cgeic/ for details) and the University of Stockholm (Orn et al. 1996) historic emissions datasets for 1860, 1870,… 1970. The GEIA 1B dataset includes seasonal variations over some regions, e.g. Europe, but for convenience we employed an annually averaged version. In any case the Stockholm datasets are annual means. The GEIA 1B dataset also splits the emissions into those from near-surface and elevated sources such as power station chimneys; the latter were injected into the third model level above the surface (approximately 550–1000 m), to allow for buoyant plume rise.

For the 'future' portion of the experiments, we used the preliminary versions of two of the four possible marker scenarios produced by the IPCC SRES group, namely A2 and B2. We also used the published versions of the 'fossil fuel intensive' variant of the A1 scenario (known as A1FI) and the B1 marker scenario, both of which were developed at a later stage (Nakicenovic et al. 2000).

The SRES SO2 emissions are provided on a regular latitude–longitude grid at a resolution of 1°× 1°. Fields are available at ten year intervals from 1990 to 2100. Note that the emissions actually used at 1990 in the HadCM3 experiments B1 and A1FI were identical to those in A2 for simplicity in conducting the experiments, rather than the slightly different values tabulated in Table 2 which reflect revisions between the preliminary and final versions of the respective emissions scenarios. The complete sequence of base dates, between which emissions are found by interpolation, is therefore 1860, 1870, … 1980, 1985, 1990, 2000, 2010, … 2090, 2100.

Define a two-dimensional field \({\cal F}\) by

$${\cal F}(x,y) = {{H(x,y,1985)} \over {H(x,y,1985) + S(x,y,1985)}}$$

where H(x, y, 1985) is the GEIA 1B elevated emission field and S(x, y, 1985) is the GEIA 1B surface emission field. \({\cal F}\) is thus the fraction of the emissions from elevated sources. (Where there are no emissions in 1985 \({\cal F}\) is set to zero.) \({\cal F}\) was used to split the Stockholm historic datasets into surface and elevated components. Clearly the fraction of emissions from elevated sources has varied in time, so introduce a factor φ (t) to describe this: assume that φ(t) is zero before 1950, 1 after 1985 and varies linearly with time between 1950 and 1985. Emissions prior to 1985 were thus split by multiplying them by \(\phi (t){\cal F}(x,y)\) to generate an elevated source dataset, the remaining emissions being assigned to the surface source field.

We have no information about possible changes in \({\cal F}\) in the future, so we assumed that no changes take place after 1985. This is not ideal, because it is clear that changes will occur, and the emission height does affect the sulfate distribution.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Johns, T.C., Gregory, J.M., Ingram, W.J. et al. Anthropogenic climate change for 1860 to 2100 simulated with the HadCM3 model under updated emissions scenarios. Climate Dynamics 20, 583–612 (2003). https://doi.org/10.1007/s00382-002-0296-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00382-002-0296-y

Keywords

Navigation