Skip to main content

Advertisement

Log in

The LMDZ4 general circulation model: climate performance and sensitivity to parametrized physics with emphasis on tropical convection

  • Published:
Climate Dynamics Aims and scope Submit manuscript

Abstract

The LMDZ4 general circulation model is the atmospheric component of the IPSL–CM4 coupled model which has been used to perform climate change simulations for the 4th IPCC assessment report. The main aspects of the model climatology (forced by observed sea surface temperature) are documented here, as well as the major improvements with respect to the previous versions, which mainly come form the parametrization of tropical convection. A methodology is proposed to help analyse the sensitivity of the tropical Hadley–Walker circulation to the parametrization of cumulus convection and clouds. The tropical circulation is characterized using scalar potentials associated with the horizontal wind and horizontal transport of geopotential (the Laplacian of which is proportional to the total vertical momentum in the atmospheric column). The effect of parametrized physics is analysed in a regime sorted framework using the vertical velocity at 500 hPa as a proxy for large scale vertical motion. Compared to Tiedtke’s convection scheme, used in previous versions, the Emanuel’s scheme improves the representation of the Hadley–Walker circulation, with a relatively stronger and deeper large scale vertical ascent over tropical continents, and suppresses the marked patterns of concentrated rainfall over oceans. Thanks to the regime sorted analyses, these differences are attributed to intrinsic differences in the vertical distribution of convective heating, and to the lack of self-inhibition by precipitating downdraughts in Tiedtke’s parametrization. Both the convection and cloud schemes are shown to control the relative importance of large scale convection over land and ocean, an important point for the behaviour of the coupled model.

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. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20

Similar content being viewed by others

Notes

  1. The pressure p l at level l is defined as a function of surface pressure p s as \(p_{l}=A_l p_{s} +B_l.\) The values of A l and B l are chosen in such a way that the \(A_l p_s\) part dominates near the surface (where A l reaches 1), so that the coordinate follows the surface topography, and B l dominates above several km, making the coordinate equivalent to a pressure coordinate.

  2. A 50-layer version is also used for stratospheric studies (Lott et al. 2005).

  3. TOGA-COARE: Tropical Ocean Global Atmosphere Coupled Ocean-Atmosphere Response Experiment.

  4. In the bucket model, the soil water content is described as a single reservoir height h which evolves according to the net water budget PE (Precipitation minus Evaporation); E = β E p , where E p is the potential evaporation (that of a free surface of water) and β = min(1,h/h p ) with h p  = 75 mm. Water in excess of the maximum content (h max = 150 mm) is lost through run-off.

  5. Note that a numerical problem in the surface scheme was identified after the realisation of the IPCC simulations. It produces occasionally very cold temperatures over one time-step in very dry continental regions in the tropics. Although a more robust and improved version of this surface model is now available, the same version as for the IPCC simulations is used here, on purpose.

  6. The AMIP II experimental protocol requirements are all fulfilled for the simulations presented here, except that the model was not explicitly spun-up at the beginning of the AMIP period. Instead, each simulation starts from a "quasi-equilibrium" state corresponding to a previous AMIP II simulation for which there was no perceptible trend in deep soil temperature and moisture.

  7. The difference is the same when comparing CLOUDSB with CLOUDSA or CONTROL simulations.

  8. In the tropics, nearly all of the upward motion associated with ensemble-average ascent occurs within cumulus clouds, and gentle subsidence occurs in-between clouds. The rate of subsidence in-between clouds being strongly constrained by the clear-sky radiative cooling (which is nearly invariant), an increase of the large-scale mean ascent corresponds, to first order, to an increase of the mass flux in cumulus clouds (Emanuel et al. 1994).

References

  • EPICA community members (2004) Eight glacial cycles from an Antarctic ice core. Nature 429:623–628

    Article  Google Scholar 

  • Arctic Climatology Project (2000) Environmental working group arctic meteorology and climate atlas, CD-Rom. In: Fetterer F, Radionov V (eds) National Snow and Ice Data Center, Boulder

  • Automatic Weather Stations Greenland Project (2004) Greenland aws data, digital data available on http://www.amrc.ssec.wisc.edu/greenland.html

  • Automatic Weather Stations Project (2004) Archive aws data, digital data available on http://www.amrc.ssec.wisc.edu/aws.html

  • Barkstrom BR (1984) The earth radiation budget experiment (ERBE). Bull Am Meteorol Soc 65:1170–1185

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Bony S, Dufresne J-L (2005) Marine boundary layer clouds at the heart of cloud feedback uncertainties in climate models. Geophys Res Lett 32(20), L20806, doi: 10.1029/2005GL023851

  • Bony S, Emanuel J-L (2001) A parameterization of the cloudiness associated with cumulus convection; evaluation using TOGA COARE data. J Atmos Sci 58:3158–3183

    Article  Google Scholar 

  • Bony S, Lau K-M, Sud YC (1997) Sea surface temperature and large-scale circulation influences on tropical greenhouse effect and cloud radiative forcing. J Clim 10:2055–2077

    Article  Google Scholar 

  • Bony S, Dufresne J-L, Le Treut H, Morcrette J-J, Senior C (2004) On dynamic and thermodynamic components of cloud changes. Clim Dyn 22:71–86

    Article  Google Scholar 

  • Boucher O, Pham M (2002) History of sulfate aerosol radiative forcings. Geophys Res Lett 29:(9):

    Google Scholar 

  • Braconnot P (1998) Tests de sensibilité, Note technique 0076, IPSL, http://www.ipsl.jussieu.fr/poles/ Modelisation/NotesTechniques.htm

  • Braconnot P, Hourdin F, Bony S, Dufresne J-L, Grandpeix J-Y, Marti O (2006) Impact of different convective cloud schemes on the simulation of the tropical seasonal cycle with a coupled ocean-atmosphere model. Clim Dyn (submitted)

  • Chen B, Bromwich DH, Hines KM, Pan X (1995) Simulations of the 1979–1988 polar climates by global climate models. Ann Glaciol 21:83–90

    Google Scholar 

  • Cosme E, Hourdin F, Genthon C, Martinerie P (2005) Origin of dimethylsulfide, non-sea-salt sulfate, and methanesulfonic acid in eastern Antarctica. J Geophys Res 110, (D9), D03302, doi:10.1029/2004JD004881

  • Deardorff JW (1966) The counter-gradient heat-flux in the lower atmosphere and in the laboratory. J Atmos Sci 23:503–506

    Article  Google Scholar 

  • Dufresne J, Grandpeix J (1996) Raccordement des modèles thermodynamiques de glace, d’océan et d’atmosphère, Note Interne 205, L.M.D., juin

  • Dufresne J-L, Quaas J, Boucher O, Denvil F, Fairhead L (2005) Contrasts in the effects on climate of anthropogenic sulfate aerosols between the 20th and the 21st century, Geophys Res Lett 32, L21703, doi: 10.1029/2005GL023619

  • Emanuel KA (1991) A scheme for representing cumulus convection in large-scale models. J Atmos Sci 48:2313–2335

    Article  Google Scholar 

  • Emanuel KA (1993) A cumulus representation based on the episodic mixing model: the importance of mixing and microphysics in predicting humidity. AMS Meteorol Monogr 24(46):185–192

    Google Scholar 

  • Emanuel KA, Neelin JD, Bretherton CS (1994) On large-scale circulations in convective atmospheres. Q J R Meteorol Soc 120:1111–1143

    Article  Google Scholar 

  • ERA-40 by ECMWF (2002) ERA-40, forty-year european re-analysis of the global atmosphere, 2002, http://www.ecmwf.int/products/data/archive/descriptions/e4

  • Fichefet T, Maqueda MM (1997) Sensitivity of a global sea ice model to the treatment of ice thermodynamics and dynamics. J Geophys Res 102:12609–12646

    Article  Google Scholar 

  • Forget F, Hourdin F, Fournier R, Hourdin C, Talagrand O, Collins M, Lewis SR, Read PL, Huot J-P (1999) Improved general circulation models of the Martian atmosphere from the surface to above 80 km. J Geophys Res 104:24155–24176

    Article  Google Scholar 

  • Fouquart Y, Bonnel B (1980) Computations of solar heating of the Earth’s atmosphere: a new parametrization. Contrib Atmos Phys 53:35–62

    Google Scholar 

  • Gates WL (1992) AMIP: the atmospheric model intercomparison project. Bull Am Meteorol Soc 73:1962–1970

    Article  Google Scholar 

  • Genthon C, Krinner G, Cosme E (2002) Free and laterally-nudged Antarctic climate of an Atmospheric General Circulation Model. Mon Weather Rev 130:1601–1616

    Article  Google Scholar 

  • Grandpeix JY, Phillips V, Tailleux R (2004) Improved mixing representation in Emanuel’s convection scheme. Q J R Meteorol Soc 130:3207–3222

    Article  Google Scholar 

  • Gregory D, Morcrette JJ, Jakob C, Beljaars ACM, Stockdale T (2000) Revision of convection, radiation and cloud schemes in the ECMWF Integrated Forecasting System. Q J R Meteorol Soc 126:1685–1710

    Article  Google Scholar 

  • Gregory J, Stouffer R, Raper S, Stott P, Rayner N (2005) An observationally based estimate of the climate sensitivity. J Clim 15(22):3117–3121

    Article  Google Scholar 

  • Grenier H, Le Treut H, Fichefet T (2000) Ocean–atmosphere interactions and climate drift in a coupled general circulation model. Clim Dyn 16:701–717

    Article  Google Scholar 

  • Guichard F, Petch JC, Redelsperger J-L, Bechtold P, Chaboureau J-P, Cheinet S, Grabowski W, Grenier H, Jones CG, Köhler M, Piriou J-M, Tailleux R, Tomasini M (2004) Modelling the diurnal cycle of deep precipitating convection over land with cloud-resolving models and single-column models. Q J R Meteorol Soc 130(604C):3139–3172

    Article  Google Scholar 

  • Hauglustaine DA, Hourdin F, Jourdain F, Filiberti M-A, Walters S, Lamarque J-F, Holland EA (2004) Interactive chemistry in the Laboratoire de Météorologie Dynamique general circulation model: description and background tropospheric chemistry evaluation. J Geophys Res 109:4314–4357

    Article  Google Scholar 

  • Heymsfield AJ, Platt C (1984) A parameterization of the particle size spectrum of ice clouds in terms of the ambient temperature and the ice water content. J Atmos Sci 41:846–855

    Article  Google Scholar 

  • Horinouchi T, Pawson S, Shibata K, Langematz U, Manzini E, Giorgetta MA, Sassi F, Wilson RJ, Hamilton K, de Grandpré J, Scaife AA (2003) Tropical cumulus convection and upward-propagating waves in middle-atmospheric GCMs. J Atmos Sci 60:2765–2782

    Article  Google Scholar 

  • Hosking JRM, Wallis JR (1997) Regional frequency analysis: an approach based on L-monents. Cambridge University Press, Cambridge, 224 pp

  • Hoskins BJ, Hsu HH, James IN, Masutani M, Sardeshmuck PD, White GH (1996) Diagnostics of the global atmospheric circulation based on ecmwf analysis 1979–1989, Technical document, WCRP/WMO

  • Hourdin F, Armengaud A (1999) The use of finite-volume methods for atmospheric advection of trace species. Part i: test of various formulations in a general circulation model. Mon Weather Rev 127:822–837

    Article  Google Scholar 

  • Hourdin F, Issartel J-P (2000) Sub-surface nuclear tests monitoring through the CTBT xenon network. Geophys Res Lett 27:2245–2248

    Article  Google Scholar 

  • Hourdin F, Le Van P, Forget F, Talagrand O (1993) Meteorological variability and the annual surface pressure cycle on Mars. J Atmos Sci 50:3625–3640

    Article  Google Scholar 

  • Hourdin F, Talagrand O, Sadourny R, Régis C, Gautier D, McKay D (1995) General circulation of the atmosphere of Titan. Icarus 117:358–374

    Article  Google Scholar 

  • Hourdin F, Lebonnois S, Luz D, Rannou P (2004) Titan’s stratospheric composition driven by condensation and dynamics. J Geophys Res 109

  • Hourdin F, Idelkadi A, Talagrand O (2005) Eulerian backtracking of atmospheric tracers: II numerical aspects, QJRMS (in press)

  • Jacquart C, Choisnel E (1995) Un modèle de bilan hydrique simplifié à deux réservoirs utilisable en agrométéorologie. La Météorologie 8:4–17

    Google Scholar 

  • Jakob C, Siebesma AP (2003) A new subcloud model for mass-flux convection schemes: influence on triggering, updraft properties, and model climate. Mon Weather Rev 131:2765–2778

    Article  Google Scholar 

  • James IN (1989) The Antarctic drainage flow: implications for hemispheric flow on the Southern hemisphere. Antarct Sci 1:279–290

    Google Scholar 

  • Kasahara A (1977) Computational aspects of numerical models for weather prediction and climate simulation. In: Chang J (ed) Methods in computational physics, vol 17, Academic, Amsterdam, pp 1–66

  • King J, Connolley W (1997) Validation of the surface energy balance over the antarctic ice sheets in the UK meteorological office unified climate model. J Clim 1273–1287

  • Klein SA, Jakob C (1999) Validation and sensitivities of frontal clouds simulated by the ECMWF model. Mon Weather Rev 127:2514–2531

    Article  Google Scholar 

  • Krinner G, Genthon C (1998) GCM simulations of the Last Glacial Maximum surface climate of Greenland and Antarctica. Clim Dyn 14:741–758

    Article  Google Scholar 

  • Krinner G, Genthon C (1999) Altitude dependence of the surface climate over the ice sheets. Geophys Res Lett 26:2227–2230

    Article  Google Scholar 

  • Krinner G, Genthon C (2003) Tropospheric transport of continental tracers towards Antarctica under varying climatic conditions. Tellus 53:54–70

    Google Scholar 

  • Krinner G, Genthon C, Li Z-X, Le Van P (1997) Studies of the Antarctic climate with a stretched-grid general circulation model. J Geophys Res 102:13731–13745

    Article  Google Scholar 

  • Krinner G, Mangerud J, Jakobsson M, Crucifix M, Ritz C, Svendsen J (2004) Enhanced ice sheet growth in Eurasia owing to adjacent ice dammed lakes. Nature 427:429–432

    Article  Google Scholar 

  • Krinner G, Viovy N, de Noblet-Ducoudré N, Ogée J, Polcher J, Friedlingstein P, Ciais P, Sitch S, Prentice C (2005) A dynamic global vegetation model for studies of the coupled atmosphere–biosphere system, Global Biogeochem Cycles 19, GB1015, 2005, doi:10.1029/2003GB002199

  • Krinner G, Magand O, Simmonds I, Genthon C, Dufresne J (2006) Simulated antarctic precipitation and surface mass balance of the end of the 20th and 21st centuries. Clim Dyn (submitted)

  • Laval K, Sadourny R, Serafini Y (1981) Land surface processes in a simplified general circulation model. Geophys Astrophys Fluid Dyn 17:129–150

    Google Scholar 

  • Le Treut H, Li ZX (1991) Sensitivity of an atmospheric general circulation model to prescribed SST changes: feedback effects associated with the simulation of cloud optical properties. Clim Dyn 5:175–187

    Google Scholar 

  • Le Treut H, Li Z, Forichon M (1994) Sensitivity study of the LMD GCM to greenhouse forcing associated with two different cloud water parametrizations. J Clim 7:1827–1841

    Article  Google Scholar 

  • Le Treut H, Forichon M, Boucher O, Li Z (1998) Sulfate aerosol indirect effect and CO 2 greenhouse forcing: equilibrium response of the LMD GCM and associated cloud feedbacks. J Climate 11:1673–1684

    Article  Google Scholar 

  • Levrard B, Forget F, Montmessin F, Laskar J (2004) Recent ice-rich deposits formed at high latitudes on Mars by sublimation of unstable equatorial ice during low obliquity. Nature 431:1072–1075

    Article  Google Scholar 

  • Li Z (1999) Ensemble atmospheric GCM simulation of climate interannual variability from 1979 to 1994. J Climate 12:986–1001

    Article  Google Scholar 

  • Li ZX, Conil S (2003) A 1000-year simulation with the IPSL ocean-atmosphere coupled model. Ann Geophys 46(1):39–46

    Google Scholar 

  • Lindzen RS (1974) Wave-CISK in the tropics. J Atmosph Sci 31:156–179

    Article  Google Scholar 

  • Lott F (1999) Alleviation of stationary biases in a gcm through a mountain drag parametrization scheme and a simple representation of mountain lift forces. Mon Weather Rev 127:788–801

    Article  Google Scholar 

  • Lott F, Miller M (1997) A new sub-grid scale orographic drag parametrization: its formulation and testing. Q J R Meteorol Soc 123:101–128

    Article  Google Scholar 

  • Lott F, Fairhead L, Hourdin F, Levan P (2005) The stratospheric version of LMDz: dynamical climatologies, arctic oscillation, and impact on the surface climate. Clim Dyn 25:851–868

    Article  Google Scholar 

  • Louis JF (1979) A parametric model of vertical eddy fluxes in the atmosphere. Boundary Layer Meteorol 17:187–202

    Article  Google Scholar 

  • Madec G, Delecluse P, Imbard M, Lévy M (1998) OPA 8.1, ocean general circulation model reference manual, Notes du pôle de modélisation, n. 11, Institut Pierre-Simon Laplace (IPSL), Paris, France

  • Marti O, Braconnot P, Bellier J, Benshila R, Bony S, Brockmann P, Cadule P, Caubel A, Denvil S, Dufresne JL, Fairhead L, Filiberti MA, Foujols MA, Fichefet T, Friedlingstein P, Grandpeix JY, Hourdin F, Krinner G, Lévy C, Madec G, Musat I, de Nolbet N, Polcher J, Talandier C (2005) The new IPSL climate system model: IPSL-CM4, Technical note, IPSL, available at http://www.dods.ipsl.jussieu.fr/omamce/IPSLCM4/DocIPSLCM4/

  • Morcrette JJ, Smith L, Fouquart Y (1986) Pressure and temperature dependence of the absorption in longwave radiation parametrizations. Contrib Atmos Phys 59(4):455–469

    Google Scholar 

  • Myneni R, Hoffman S, Glassy J, Zhang Y, Votava P, Nemani R, Running S, Privette J (2002) Global products of vegetation leaf area and fraction absorbed PAR from year one of MODIS data. Remote Sens Environ 83:214–231

    Article  Google Scholar 

  • Quaas J, Boucher O, Bréon FM (2004) Aerosol indirect effects in POLDER satellite data and the Laboratoire de Météorologie Dynamique-Zoom (LMDZ) general circulation model. J Geophys Res 109

  • Quadrelli R, Wallace JM (2004) A simplified linear framework for interpreting patterns of northern hemisphere wintertime climate variability. J Climate 17:3728–3744

    Article  Google Scholar 

  • Rannou P, Hourdin F, McKay CP (2002) A wind origin for Titan’s haze structure. Nature 418:853–856

    Article  Google Scholar 

  • de Rosnay P, Polcher J, Bruen M, Laval K (2002) Impact of a physically based soil water flow and soil–plant interaction representation for modeling large scale land surface processes. J Geophys Res 107, 10.1029/2001JD000634

  • Rossow WB, Schiffer RA (1999) Advances in understanding clouds from isccp. Bull Am Meteorol Soc 80:2261–2287

    Article  Google Scholar 

  • Sadourny R (1975a) Compressible model flows on the sphere. J Atmos Sci 32:2103–2110

    Article  Google Scholar 

  • Sadourny R (1975b) The dynamics of finite-difference models of the shallow-water equations. J Atmos Sci 32:680–689

    Article  Google Scholar 

  • Sadourny R, Laval K (1984) January and July performance of the LMD general circulation model. In: Berger A, Nicolis C (eds) New perspectives in climate modeling. Elsevier, Amsterdam, pp 173–197

    Google Scholar 

  • Sawyer JS (1976) Observational characteristics of atmospheric fluctuations with a time scale of a month. Q J R Meteorol Soc 96:610–625

    Google Scholar 

  • Shuman C, Alley R, Anandakrishnan S, White J, Grootes P, Stearns C (1995) Temperature and accumulation at the Greenland Summit: comparison of high-resolution isotope profiles and satellite passive microwave brightness temperature trends. J Geophys Res 100:9165–9177

    Article  Google Scholar 

  • Slingo JM (1987) The development and verification of a cloud prediction scheme for the ecmwf model. Q J R Meteorol Soc 113(477):899–927

    Article  Google Scholar 

  • Smith SD (1988) Coefficients for sea surface wind stress, heat flux, and wind profiles as a function of wind speed and temperature. J Geophys Res 93:15467–15472

    Article  Google Scholar 

  • Sommeria G, Deardorff JW (1977) Subgrid-scale condensation in models of nonprecipitating clouds. J Atmos Sci 34:344–355

    Article  Google Scholar 

  • Suzuki T, Tanaka M, Nakajima T (1993) The microphysical feedback of cirrus cloud in climate-change. J Meteor Soc Jpn 71(6):701–714

    Google Scholar 

  • Taylor KE, Williamson D, Zwiers F (2000) The sea surface temperature and sea–ice concentration boundary conditions for AMIP II simulations, PCMDI Report No. 60 60 P.C.M.D.I.

  • Thompson DWJ, Wallace JM (1998) The Arctic Oscillation signature in the wintertime geopotential height and temperature fields. Geophys. Res. Lett. 25:1297–1300

    Article  Google Scholar 

  • Tiedtke M (1989) A comprehensive mass flux scheme for cumulus parameterization in large-scale models. Mon Weather Rev 117:1179–1800

    Article  Google Scholar 

  • Van Leer B (1977) Towards the ultimate conservative difference scheme : IV. A new approach to numerical convection. J Comput Phys 23:276–299

    Article  Google Scholar 

  • Vaughan DG, Bamber J, Giovinetto M, Russel J, Cooper A (1999) Reassessment of net surface mass balance in Antarctica. J Clim 12:933–946

    Article  Google Scholar 

  • Webb M, Senior C, Bony S, Morcrette JJ (2001) Combining ERBE and ISCCP data to assess clouds in the Hadley Centre, ECMWF and LMD atmospheric climate models. Clim Dyn 17:905–922

    Article  Google Scholar 

  • Wigley T, Jones P, Raper S (1997) The observed global warming record: what does it tell us?. Proc Natl Acad Sci 94(16):8314–8320

    Article  Google Scholar 

  • Wyant MC, Bretherton CS, Bacmeister JT, Kiehl JT, Held IM, Zhao MZ, Klein SA, Soden BJ (2006) A comparison of tropical cloud properties and responses in gcms using mid-tropospheric vertical velocity. Clim Dyn (submitted)

  • Xie P, Arkin PA (1997) 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 

  • Xu K-M, Randall DA (1996) Evaluation of statistically based cloudiness parameterisations used in climate models. J Atmosph Sci 53:3103–3119

    Article  Google Scholar 

  • Zhang MH, Lin WY, Klein SA, Bacmeister JT, Bony S, Cderwall RT, Del Genio AD, Hack JJ, Loeb NG, Lohmann v, Minnis P, Musat I, Pincus R, Stier P, Suarez MJ, Webb MJ, Wu JB, Xie MS, Zhang JH (2005) Comparing clouds and their seasonal variations in 10 atmospheric general circulation models with satellite measurements. J Geophys Res 110, D15S02, doi:10.1029/2004JD005021

  • Zhou TJ, Li ZX (2002) Simulation of the east asian summer monsoon using a variable resolution atmospheric gcm. Clim Dyn 19:167–180

    Article  Google Scholar 

Download references

Acknowledgements

The numerical simulations presented here were performed on the NEC-SX5 of the IDRIS/CNRS computer centre. The graphics have been made with the user-friendly and public domain graphical package GrADS originally developed by Brian Dotty (COLA, support@grads.iges.org). The authors thank the anonymous referees for their constructive comments which helped us to improve the original version of the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Frédéric Hourdin.

Appendix 1: about vertically integrated velocity potential

Appendix 1: about vertically integrated velocity potential

The purpose of this appendix is: (1) to express the vertical momentum of atmospheric columns as a z-weighted integral of the Laplacian of the velocity potential; (2) to show that it is also approximately the Laplacian of a z-weighted integral of the velocity potential. Only monthly mean velocity fields are considered and the scalar velocity potential at each level is chosen so that it is zero at the poles.

1.1 Vertical momentum

The vertical momentum \(\tilde{w}\) of atmospheric columns reads:

$$ \tilde{w} = \int\limits_{z_s}^{\infty} {\text{d}}z \rho w \simeq - \int\limits_{z_s}^{\infty} {\text{d}}z {\frac{\omega} {g}} $$
(7)

where w and ω (≃ −ρ g w for these monthly mean fields) are the vertical velocity expressed in z and pressure coordinates, respectively, and z s is the altitude of the surface.

Vertical integration of the continuity equation (taking into account \(\vec{\nabla} \vec{V} = \nabla^2 \varphi\)) yields an expression of ω in terms of the velocity potential:

$$ \omega_{(z)} - \omega_s = \int\limits_{z_s}^z {\text{d}}z^{\prime} \nabla^2 \varphi_{(z^{\prime})} \rho_{(z^{\prime})}g $$
(8)

For monthly mean fields, the term \({\omega_s = {{\partial p_s}/{\partial t}}}\) is negligible. Then, substitution of (8) in (7) yields:

$$ \tilde{w} = - \int\limits_0^{\infty} {\text{d}}z \int\limits_{z_s}^z {\text{d}}z^{\prime} \rho_{(z^{\prime})} \nabla^2 \varphi_{(z^{\prime})} $$

Let Z 0 be an altitude high enough so that \(\omega_{(Z_0)} \simeq 0.\) Then the integration may be limited to the triangle (z<Z 0, z′<z). Permuting the two integrations yields:

$$\begin{aligned} \tilde{w} &= - \int\limits_{z_s}^{Z_0} {\text{d}}z^{\prime} \int\limits_{z^{\prime}}^{Z_0} {\text{d}}z \rho_{(z^{\prime})} \nabla^2 \varphi_{(z^{\prime})} \\ &= - \int\limits_{z_s}^{Z_0} {\text{d}}z^{\prime} \rho_{(z^{\prime})} \nabla^2 \varphi_{(z^{\prime})} (Z_0 - z^{\prime})\\ \end{aligned} $$
(9)

The Z 0 term drops out since the vertical integral of ∇2 φ is close to zero:

$$ \tilde{w} = {\frac{1} {g}} \int\limits_{p_0}^{p_s} {\text{d}}p z_{(p)} \nabla^2 \varphi_{(p)} $$
(10)

1.2 Expressing the vertical momentum \(\tilde{w}\) in terms of the potential of the geopotential transport

We wish to express \(\tilde{w}\) as the Laplacian of some potential. In order to do this, one has to commute the Laplacian operator in formula (10) with the z term and with the integral operator.

We shall limit ourselves to the tropical band where the geopotential altitude has weak horizontal variations. With such an approximation, the Laplacian and the z term commute.

Now we want to commute the horizontal differentials with the vertical integration. Taking into account the fact that the velocity is zero at the surface (so that \(\vec{\nabla} \varphi(p_s) = 0\)), one may write:

$$\begin{aligned} {\frac{\partial} {\partial x}} \left(\int\limits_{p_0}^{p_s} {\text{d}}p z\varphi \right) &= \int\limits_{p_0}^{p_s} {\text{d}}p z {\frac{\partial \varphi} {\partial x}} + z_{(p_s)} \varphi_{(p_s)} {\frac{\partial p_s} {\partial x}}\\ {\frac{\partial^2} {\partial x^2}} \left(\int\limits_{p_0}^{p_s} {\text{d}}p z\varphi \right) & = \int\limits_{p_0}^{p_s} {\text{d}}p z {\frac{\partial^2 \varphi} {\partial x^2}} + z_{(p_s)} \varphi_{(p_s)} {\frac{\partial^2 p_s} {\partial x^2}} \end{aligned}$$
(11)

Adding the analog formula for the y derivative, one gets:

$$ \nabla^2 \left(\int\limits_{p_0}^{p_s} {\text{d}}p z\varphi \right) = \int\limits_{p_0}^{p_s} {\text{d}}p z \nabla^2 \varphi + z_{(p_s)} \varphi_{(p_s)} \nabla^2 p_s $$
(12)

Over oceans, the last rhs term is zero. Over continents, it is not necessarily zero, because of orography. However, with a spatial resolution of the order of 100 km, it stays several order of magnitude smaller than the first rhs term and we shall neglect it.

Then, one may write the vertical momentum \(\tilde{w}\) as the Laplacian of a function \(\tilde{\varphi}:\)

$$ \left\{ {\begin{array}{*{20}c} \tilde{w} \simeq {\frac{1} {g}} \nabla^2 \tilde{\varphi} \\ \tilde{\varphi} = \int\limits_{p_0}^{p_s} {\text{d}}p z \varphi \\ \end{array}} \right.$$
(13)

Finally, using the same technique and the same approximations one may prove that \(\tilde{\varphi}\) is close to the saclar potential of the horizontal transport \(\vec{G}\) of geopotential:

$$\begin{aligned} \vec{G} &= \int\limits_0^{\infty} {\text{d}}z \rho \vec{V} g z \\ & = \int\limits_0^{p_s} {\text{d}}p z \vec{V} \end{aligned}$$
(14)

As an illustration, Fig. 21 displays the potential \(\tilde{\varphi}\) of the annual mean geopotential transport and the mean vertical velocity. The similarity of \(\overline{\omega}\) and ω500 is obvious. However, \(\overline{\omega}\) is smoother than ω500 and might be a better indicator of dynamic regimes. Finally, the lowest pannel illustrates the strong link between large-scale ascent and precipitation.

Fig. 21
figure 21

Potential \(\tilde{\varphi}\) of the annual mean of the horizontal transport of geopotential (upper pannel) and mean vertical velocity \(\overline{\omega} = {\frac{1}{Z_0-z_s}} \int_{z_s}^{Z_0} {\text{d}}z \omega = {{-1}\over {Z_0-z_s}} {\nabla^2 \tilde{\varphi}}\) (with Z 0z s =15 km) (second pannel) for one of the AMIP simulations; to be compared with ω500 and annual precipitation (two lower pannels)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Hourdin, F., Musat, I., Bony, S. et al. The LMDZ4 general circulation model: climate performance and sensitivity to parametrized physics with emphasis on tropical convection. Clim Dyn 27, 787–813 (2006). https://doi.org/10.1007/s00382-006-0158-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00382-006-0158-0

Keywords

Navigation