Hostname: page-component-848d4c4894-nmvwc Total loading time: 0 Render date: 2024-07-02T06:35:37.614Z Has data issue: false hasContentIssue false

Three-dimensional baroclinic instability of a Hadley cell for small Richardson number

Published online by Cambridge University Press:  20 April 2006

Basil N. Antar
Affiliation:
The University of Tennessee Space Institute, Tullahoma, TN 37388
William W. Fowlis
Affiliation:
Space Science Laboratory, NASA Marshall Space Flight Center, Huntsville, AL 35812

Abstract

A three-dimensional linear stability analysis of a baroclinic flow for Richardson number Ri of order unity is presented. The model considered is a thin, horizontal, rotating fluid layer which is subjected to horizontal and vertical temperature gradients. The basic state is a Hadley cell which is a solution of the Navier–Stokes and energy equations and contains both Ekman and thermal boundary layers adjacent to the rigid boundaries; it is given in closed form. The stability analysis is also based on the Navier–Stokes and energy equations; and perturbations possessing zonal, meridional and vertical structures were considered. Numerical methods were developed for the solution of the stability problem, which results in an ordinary differential eigenvalue problem. The objectives of this work were to extend the previous theoretical work on three-dimensional baroclinic instability for small Ri to a more realistic model involving the Prandtl number σ and the Ekman number E, and to finite growth rates and a wider range of the zonal wavenumber. The study covers ranges of 0.135 [les ] Ri [les ] 1.1, 0.2 [les ] σ [les ] 5.0, and 2 × 10−4 [les ] E [les ] 2 σ 10−3. For the cases computed for E = 10−3 and σ ≠ 1, we found that conventional baroclinic instability dominates for Ri > 0.825 and symmetric baroclinic instability dominates for Ri < 0.675. However, for E [ges ] 5 × 10−4 and σ = 1 in the range 0.3 [les ] Ri [les ] 0.8, conventional baroclinic instability always dominates. Further, we found in general that the symmetric modes of maximum growth are not purely symmetric but have weak zonal structure. This means that the wavefronts are inclined at a small angle to the zonal direction. The results also show that as E decreases the zonal structure of the symmetric modes of maximum growth rate also decreases. We found that when zonal structure is permitted the critical Richardson number for marginal stability is increased, but by only a small amount above the value for pure symmetric instability. Because these modes do not substantially alter the results for pure symmetric baroclinic instability and because their zonal structure is weak, it is unlikely that they represent a new type of instability.

Type
Research Article
Copyright
© 1983 Cambridge University Press

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Antar, B. N. & Fowlis, W. W. 1981 Baroclinic instability of a rotating Hadley cell J. Atmos. Sci. 38, 21302141.Google Scholar
Antar, B. N. & Fowlis, W. W. 1982 Symmetric baroclinic instability of a Hadley cell J. Atmos. Sci. 39, 12801289.Google Scholar
Bennets, D. A. & Hoskins, B. J. 1979 Conditional symmetric instability a possible explanation for frontal rainbands Q. J. R. Met. Soc. 105, 945962.Google Scholar
Busse, F. H. & Chen, W. L. 1981 On the (nearly) symmetric instability J. Atmos. Sci. 38, 877880.Google Scholar
Calman, J. 1977 Experiments on high Richardson number instability of a rotating stratified shear flow Dyn. Atmos. Oceans 1, 277297.Google Scholar
Charney, J. G. 1947 The dynamics of long waves in a baroclinic westerly current J. Met. 4, 135162.Google Scholar
Conte, S. D. 1966 The numerical solution of linear boundary value problems S.I.A.M. Rev. 8, 309320.Google Scholar
Eady, E. T. 1949 Long waves and cyclone waves Tellus 1, 3552.Google Scholar
Emanuel, K. A. 1979 Inertial instability and mesoscale convective systems. Part I: Linear theory of inertial instability in rotating viscous fluids. J. Atmos. Sci. 36, 2425810.Google Scholar
Hadlock, R. K., Na, J. Y. & Stone, P. H. 1972 Direct thermal verification of symmetric baroclinic instability J. Atmos. Sci. 29, 13911393.Google Scholar
Hide, R. & Mason, P. J. 1975 Sloping convection in a rotating fluid Adv. Phys. 24, 47100.Google Scholar
Kuo, H. L. 1956 Forced and free axially symmetric convection produced by differential heating in a rotating fluid J. Met. 13, 521527.Google Scholar
Lorenz, E. N. 1967 The Nature and Theory of the General Circulation of the Atmosphere. World Meteorological Organization.
Mcintyre, M. E. 1970 Diffusive destabilization of the baroclinic circular vortex Geophys. Fluid Dyn. 1, 1957.Google Scholar
Pedlosky, J. 1979 Geophysical Fluid Dynamics. Springer.
Solberg, H. 1936 Le Mouvement d'inertie de l'atmosphère stable et son rôle dans le théorie des cyclones. In Proc. Union Géodesique et Géophysique Internationale VIième Assemblée, Edinburgh, vol. II, pp. 6682.
Stone, P. H. 1966 On non-geostrophic baroclinic stability J. Atmos. Sci. 23, 390400.Google Scholar
Stone, P. H. 1967 An application of baroclinic stability theory to the dynamics of the Jovian atmosphere J. Atmos. Sci. 24, 642652.Google Scholar
Stone, P. H. 1970 On non-geostrophic baroclinic stability: Part II. J. Atmos. Sci. 27, 721810.Google Scholar
Stone, P. H. 1971 Baroclinic stability under non-hydrostatic conditions J. Fluid Mech. 45, 659671.Google Scholar
Stone, P. H., Hess, S., Hadlock, R. & Ray, P. 1969 Preliminary results of experiments with symmetric baroclinic instability J. Atmos. Sci. 26, 991996.Google Scholar
Tokioka, T. 1970 Non-geostrophic and non-hydrostatic stability of a baroclinic fluid J. Met. Soc. Japan 48, 503520.Google Scholar