Skip to main content

Advertisement

Log in

Changes in storm tracks and energy transports in a warmer climate simulated by the GFDL CM2.1 model

  • Published:
Climate Dynamics Aims and scope Submit manuscript

Abstract

Storm tracks play a major role in regulating the precipitation and hydrological cycle in midlatitudes. The changes in the location and amplitude of the storm tracks in response to global warming will have significant impacts on the poleward transport of heat, momentum and moisture and on the hydrological cycle. Recent studies have indicated a poleward shift of the storm tracks and the midlatitude precipitation zone in the warming world that will lead to subtropical drying and higher latitude moistening. This study agrees with this key feature for not only the annual mean but also different seasons and for the zonal mean as well as horizontal structures based on the analysis of Geophysical Fluid Dynamics Laboratory (GFDL) CM2.1 model simulations. Further analyses show that the meridional sensible and latent heat fluxes associated with the storm tracks shift poleward and intensify in both boreal summer and winter in the late twenty-first century (years 2081–2100) relative to the latter half of the twentieth century (years 1961–2000). The maximum dry Eady growth rate is examined to determine the effect of global warming on the time mean state and associated available potential energy for transient growth. The trend in maximum Eady growth rate is generally consistent with the poleward shift and intensification of the storm tracks in the middle latitudes of both hemispheres in both seasons. However, in the lower troposphere in northern winter, increased meridional eddy transfer within the storm tracks is more associated with increased eddy velocity, stronger correlation between eddy velocity and eddy moist static energy, and longer eddy length scale. The changing characteristics of baroclinic instability are, therefore, needed to explain the storm track response as climate warms. Diagnosis of the latitude-by-latitude energy budget for the current and future climate demonstrates how the coupling between radiative and surface heat fluxes and eddy heat and moisture transport influences the midlatitude storm track response to global warming. Through radiative forcing by increased atmospheric carbon dioxide and water vapor, more energy is gained within the tropics and subtropics, while in the middle and high latitudes energy is reduced through increased outgoing terrestrial radiation in the Northern Hemisphere and increased ocean heat uptake in the Southern Hemisphere. This enhanced energy imbalance in the future climate requires larger atmospheric energy transports in the midlatitudes which are partially accomplished by intensified storm tracks. Finally a sequence of cause and effect for the storm track response in the warming world is proposed that combines energy budget constraints with baroclinic instability theory.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Notes

  1. The projected surface warming has a local minimum in the North Atlantic region due to the weakening of the Atlantic Meridional Overturning Circulation (MOC) predicted by climate models.

  2. The GFDL CM2.1 model included both stratospheric ozone depletion in the twentieth century simulation and ozone recovery in the A1B scenario simulation. Son et al. (2008, 2009) found that the recent poleward shift of the midlatitude westerly jet and the poleward expansion of the Hadley Cell in the Southern Hemisphere were partially forced by ozone depletion and will be likely to be weakened or even reversed (depending on models) by anticipated ozone recovery in the current century. Ozone change and its effect are almost negligible in the Northern Hemisphere.

  3. The zonal mean meridional temperature gradient reduces despite an increase in temperature gradient in the North Atlantic region.

  4. In order to pick out the major feature of the change in \(F_{atm}^{net}\), a cubic smoothing spline has been applied to it in the meridional direction to remove small scale noises.

References

  • Bengtsson L, Hodges KI, Roeckner E (2006) Storm tracks and climate change. J Clim 19:3518–3543

    Article  Google Scholar 

  • Blackmon ML (1976) A climatological spectral study of the 500 mb geopotential height of the Northern Hemisphere. J Atmos Sci 33:1607–1623

    Article  Google Scholar 

  • Carissimo BC, Oort AH, Haar THV (1985) Estimating the meridional energy transports in the atmosphere and ocean. J Phys Oceanogr 15:82–91

    Article  Google Scholar 

  • Chang EK (1993) Downstream development of baroclinic waves as inferred from regression analysis. J Atmos Sci 50:2038–2053

    Article  Google Scholar 

  • Chang EKM, Lee S, Swanson KL (2002) Storm track dynamics. J Clim 15:2163–2183

    Article  Google Scholar 

  • Charney JG (1947) The dynamics of long waves in a baroclinic westerly current. J Meteorol 4:135–162

    Article  Google Scholar 

  • Chen G, Lu J, Frierson DM (2008) Phase speed spectra and the latitude of surface westerlies: interannual variability and global warming trend. J Clim 21:5942–5959

    Article  Google Scholar 

  • Delworth TL et al (2006) GFDL’s CM2 global coupled climate models. Part I. Formulation and simulation characteristics. J Clim 19:643–674

    Article  Google Scholar 

  • Eady ET (1949) Long waves and cyclone waves. Tellus 1:33–52

    Article  Google Scholar 

  • Emanuel KA, Fantini M, Thorpe AJ (1987) Baroclinic instability in an environment of small stability to slantwise moist convection. Part I. Two-dimensional models. J Atmos Sci 44:1559–1573

    Article  Google Scholar 

  • Frierson DMW, Held IM, Zurita-Gotor P (2006) A gray-radiation aquaplanet moist GCM. Part I. Static stability and eddy scale. J Atmos Sci 63:2548–2566

    Article  Google Scholar 

  • Frierson DMW, Held IM, Zurita-Gotor P (2007) A gray-radiation aquaplanet moist GCM. Part II. Energy transports in altered climates. J Atmos Sci 64:1680–1693

    Article  Google Scholar 

  • Fyfe JC (2003) Extratropical Southern Hemisphere cyclones: Harbingers of climate change. J Clim 16:2802–2805

    Article  Google Scholar 

  • Gastineau G, Soden BJ (2009) Model projected changes of extreme wind events in response to global warming. Geophys Res Lett 36:L10 810. doi:10.1029/2009GL037500

    Article  Google Scholar 

  • Geng Q, Sugi M, (2003) Possible change of extratropical cyclone activity due to enhanced greenhouse gases and sulfate aerosols—study with a high-resolution AGCM. J Clim 16:2262–2274

    Article  Google Scholar 

  • Green JSA (1970) Transfer properties of the large-scale eddies and the general circulation of the atmosphere. Quart J Roy Meteorol Soc 96:157–185

    Article  Google Scholar 

  • Gregory JM, Mitchell JFB, Brady AJ (1996) Summer drought in Northern Midlatitudes in a time-dependent CO2 climate experiment. J Clim 10:662–686

    Article  Google Scholar 

  • Hall NMJ, Hoskins BJ, Valdes PJ, Senior CA (1994) Storm tracks in a high-resolution GCM with doubled carbon dioxide. Quart J Roy Meteorol Soc 120:1209–1230

    Article  Google Scholar 

  • Hansen JE, Lacis A, Rind D, Russell G, Stone P, Fung I, Ruedy R, Lerner J (1984) Climate sensitivity: analysis of feedback mechanisms. In: Hansen JE, Takahashi T (eds) Climate processes and climate sensitivity, AGU Geophysical Monograph 29, Maurice Ewing vol 5. American Geophysical Union, pp 130–163

  • Harnik N, Lindzen RS (1997) The effect of basic-state potential vorticity gradients on the growth of baroclinic waves and the height of the tropopause. J Atmos Sci 55:344–360

    Article  Google Scholar 

  • Hazeleger W, (2005) Can global warming affect tropical ocean heat transport?. Geophys Res Lett 32:L22 701. doi:10.1029/2005GL023450

    Article  Google Scholar 

  • Held IM (1993) Large-scale dynamics and global warming. Bull Am Meteorol Soc 74:228–241

    Article  Google Scholar 

  • Held IM, O’Brien E (1992) Quasigeostrophic turbulence in a three-layer model: effects of vertical structure in the mean shear. J Atmos Sci 49:1861–1870

    Article  Google Scholar 

  • Held IM, Soden BJ (2006) Robust response of the hydrological cycle to global warming. J Clim 19:5686–5699

    Article  Google Scholar 

  • Hoskins BJ, Valdes PJ (1990) On the existence of storm-tracks. J Atmos Sci 47:1854–1864

    Article  Google Scholar 

  • Kalnay E et al (1996) The NCEP/NCAR 40-year reanalysis project. Bull Am Meteorol Soc 77:437–471

    Article  Google Scholar 

  • Kunz T (2008) The role of breaking synoptic scale Rossby waves for the North Atlantic oscillation and its coupling with the stratosphere. Ph.D. thesis, der Universitat Hamburg

  • Lambert SJ (1995) The effect of enhanced greenhouse warming on winter cyclone frequencies and strengths. J Clim 8:1447–1452

    Article  Google Scholar 

  • Lambert SJ, Fyfe JC (2006) Changes in winter cyclones frequencies and strengths simulated in enhanced greenhouse warming experiments: results from the models participating in the IPCC diagnostic exercise. Clim Dyn 26:713–728

    Article  Google Scholar 

  • Lapeyre G, Held IM (2004) The role of moisture in the dynamics and energetics of turbulent baroclinic eddies. J Atmos Sci 61:1693–1710

    Article  Google Scholar 

  • Levitus S, Antonov JI, Wang J, Delworth TL, Dixon KW, Broccoli AJ (2001) Anthropogenic warming of earth’s climate system. Sci 292:267–270

    Article  Google Scholar 

  • Lin SJ (2004) A vertically Lagrangian finite-volume dynamical core for global models. Mon Wea Rev 132: 2293–2307

    Article  Google Scholar 

  • Lindzen RS, Farrell BF (1980) A simple approximation result for maximum growth rate of baroclinic instabilities. J Atmos Sci 37:1648–1654

    Article  Google Scholar 

  • Lorenz EN (1955) Available potential energy and the maintenance of the general circulation. Tellus 7:157–167

    Article  Google Scholar 

  • Lorenz DJ, DeWeaver ET (2007) Tropopause height and zonal wind response to global warming in the IPCC scenario integrations. J Geophys Res 112:D10 119. doi:10.1029/2006JD008087

    Article  Google Scholar 

  • Lu J, Vecchi GA, Reichler T (2007) Expansion of the Hadley cell under global warming. Geophys Res Lett 34:L06 805. doi:10.1029/2006GL028443

    Google Scholar 

  • Lunkeit F, Fraedrich K, Bauer SE (1998) Storm tracks in a warmer climate: sensitivity studies with a simplified global circulation model. Clim Dyn 14:813–826

    Article  Google Scholar 

  • McCabe GJ, Clark MP, Serreze MC (2001) Trends in Northern Hemisphere surface cyclone frequency and intensity. J Clim 14:2763–2768

    Article  Google Scholar 

  • Meehl GA et al (2007) Global climate projections. In: Solomon S, Qin D, Manning M, Chen Z, Marquis M, Averyt KB, Tignor M, Miller HL (eds) Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, Cambridge University Press, Chap. 10, pp 747–846

  • Nakamura H, Sampe T, Tanimoto Y, Shimpo A (2004) Observed associations among storm tracks, jet streams and midlatitude oceanic fronts, earth’s climate: the ocean–atmosphere interaction. In: Wang C, Xie S-P, Carton JA (eds) Geophys Monogr, 147, American Geophysical Union, Washington, pp 329–346

  • O’Gorman PA, Schneider T (2008) Energy of midlatitude transient eddies in idealized simulations of changed climates. J Clim 21:5797–5806

    Article  Google Scholar 

  • Pavan V (1995) Sensitivity of a multi-layer quasi-geostrophic β-channel to the vertical structure of the equilibrium meridional temperature gradient. Quart J Roy Meteorol Soc 122:55–72

    Google Scholar 

  • Pavan V, Hall N, Valdes P, Blackburn M (1999) The importance of moisture distribution for the growth and energetics of mid-latitude systems. Ann Geophys 17:242–256

    Article  Google Scholar 

  • Randel WJ, Held IM (1991) Phase speed spectra of transient eddy fluxes and critical layer absorption. J Atmos Sci 48:688–697

    Article  Google Scholar 

  • Russell JL, Dixon KW, Gnanadesikan A, Stouffer RJ, Toggweiler JR (2006a) The Southern Hemisphere westerlies in a warming world: propping open the door to the deep ocean. J Clim 19:6382–6390

    Article  Google Scholar 

  • Russell JL, Souffer RJ, Dixon KW (2006b) Intercomparison of the Southern Ocean circulations in the IPCC coupled model control simulations. J Clim 19:4560–4575

    Article  Google Scholar 

  • Seager R, Murtugudde R, Clement AC, Herweijer C (2003) Why is there an evaporation minimum at the Equator. J Clim 16:3792–3801

    Google Scholar 

  • Seager R, et al (2007) Model projections of an imminent transition to a more arid climate in southwestern North America. Science 316: 1181–1184

    Article  Google Scholar 

  • Seager R, Naik N, Vecchi G (2010) Thermodynamic and dynamic causes of large-scale changes in the hydrological cycle in response to global warming. J Clim (submitted)

  • Seager R. Vecchi G (2010) Global warming and the 21st century climate of southwestern North America. Proc Natl Acad Sci USA (submitted)

  • Son S-W et al (2008) The impact of stratospheric ozone recovery on the Southern Hemisphere westerly jet. Science 320: 1486–1489

    Article  Google Scholar 

  • Son S-W, Tandon NF, Polvani LM, Waugh DW (2009) Ozone hole and Southern Hemisphere climate change. Geophys Res Lett 36:L15 705. doi:10.1029/2009GL038671

    Article  Google Scholar 

  • Stephenson DB, Held IM, (1993) GCM response of northern winter stationary waves and storm tracks to increasing amounts of carbon dioxide. J Clim 6:1859–1870

    Article  Google Scholar 

  • Trenberth KE, Caron JM (2001) Estimates of meridional atmosphere and ocean heat transports. J Clim 14:3433–3443

    Article  Google Scholar 

  • Trenberth KE, Stepaniak DP (2003) Seamless poleward atmospheric energy transports and implications for the Hadley circulation. J Clim 16:3705–3721

    Google Scholar 

  • Trenberth KE, Caron JM, Stepaniak DP (2001) The atmospheric energy budget and implications for surface fluxes and ocean heat transports. Clim Dyn 17:259–276

    Article  Google Scholar 

  • Trenberth KE, Fasullo JT, Kiehl J (2009) Earth’s global energy budget. Bull Am Meteorol Soc 90:311–323

    Article  Google Scholar 

  • Vallis GK, (2006) Atmospheric and oceanic fluid dynamics. Cambridge University Press, Cambridge, p 745

    Book  Google Scholar 

  • Wittman MAH, Charlton AJ, Polvani LM (2007) The effect of lower stratospheric shear on baroclinic instability. J Atmos Sci 64:479–496

    Article  Google Scholar 

  • Yin H (2005) A consistent poleward shift of the storm tracks in simulations of 21st century climate. Geophys Res Lett 32: L18 701. doi:10.1029/2005GL023684

    Article  Google Scholar 

Download references

Acknowledgments

The authors would like to thank two anonymous reviewers for their insightful comments which lead to significant improvement of the paper. We also thank Nili Harnik for helpful discussions. This work was supported by NOAA CPPA Program and by NSF grant ATM-0543256. YW was also supported by NASA Headquarters under the NASA Earth and Space Science Fellowship Program -Grant NNX08AU80H.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yutian Wu.

Appendix

Appendix

1.1 Change in oceanic heat transport

The meridional oceanic heat transport can be derived from the surface energy balance and is written as

$$ S_o = R_{surface} - SH - LH - \nabla\cdot F_o $$
(8)

where S o denotes the rate of ocean heat storage, R surface is the net downward radiative flux at the surface, SH and LH are the surface upward sensible and latent heat fluxes, and F o denotes the oceanic heat transport. The difference from the derivation in atmospheric energy transport is that the ocean heat storage can’t be neglected in the long term as climate warms. The heat storage rate is defined as the oceanic potential temperature difference between years 1961–2000 and years 2081–2100 integrated over the ocean depth, multiplying by the sea water density and heat capacity, divided by the time period of 110 years. It is quite uniformly distributed over latitude and is calculated to be about 1.5 W/m2 on global average (consistent with Russell et al. (2006a)). Increased heat storage is relatively larger over the Southern Ocean and smaller over the North Atlantic region (not shown). Figure 13a shows the climatological meridional total, atmospheric and oceanic energy transport derived from the energy balance requirement. Taking the differential ocean heat storage rate into account, Fig. 13b shows the change in total, atmospheric and oceanic heat transport in the future climate. It is noted that the oceanic heat transport decreases almost everywhere whereas the atmospheric heat transport increases. Held and Soden (2006) pointed out that the reduction in meridional ocean heat transport serves as a compensation for the increased poleward latent heat transport in the atmosphere.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Wu, Y., Ting, M., Seager, R. et al. Changes in storm tracks and energy transports in a warmer climate simulated by the GFDL CM2.1 model. Clim Dyn 37, 53–72 (2011). https://doi.org/10.1007/s00382-010-0776-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00382-010-0776-4

Keywords

Navigation