Hostname: page-component-7c8c6479df-7qhmt Total loading time: 0 Render date: 2024-03-28T18:31:22.724Z Has data issue: false hasContentIssue false

Vorticity inversion and action-at-a-distance instability in stably stratified shear flow

Published online by Cambridge University Press:  14 January 2011

A. RABINOVICH
Affiliation:
Department of Geophysics and Planetary Sciences, Tel-Aviv University, Tel-Aviv 69978, Israel
O. M. UMURHAN
Affiliation:
School of Mathematical Sciences, Queen Mary, University of London, London E1 4NS, UK Department of Astronomy, City College of San Francisco, San Francisco, CA 94112, USA
N. HARNIK
Affiliation:
Department of Geophysics and Planetary Sciences, Tel-Aviv University, Tel-Aviv 69978, Israel
F. LOTT
Affiliation:
Laboratoire de Météorologie Dynamique, École Normale Supérieure, 24 rue Lhomond, 75231 Paris CEDEX 05, France
E. HEIFETZ*
Affiliation:
Department of Geophysics and Planetary Sciences, Tel-Aviv University, Tel-Aviv 69978, Israel
*
Email address for correspondence: eyalh@post.tau.ac.il

Abstract

The somewhat counter-intuitive effect of how stratification destabilizes shear flows and the rationalization of the Miles–Howard stability criterion are re-examined in what we believe to be the simplest example of action-at-a-distance interaction between ‘buoyancy–vorticity gravity wave kernels’. The set-up consists of an infinite uniform shear Couette flow in which the Rayleigh–Fjørtoft necessary conditions for shear flow instability are not satisfied. When two stably stratified density jumps are added, the flow may however become unstable. At each density jump the perturbation can be decomposed into two coherent gravity waves propagating horizontally in opposite directions. We show, in detail, how the instability results from a phase-locking action-at-a-distance interaction between the four waves (two at each jump) but can as well be reasonably approximated by the interaction between only the two counter-propagating waves (one at each jump). From this perspective the nature of the instability mechanism is similar to that of the barotropic and baroclinic ones. Next we add a small ambient stratification to examine how the critical-level dynamics alters our conclusions. We find that a strong vorticity anomaly is generated at the critical level because of the persistent vertical velocity induction by the interfacial waves at the jumps. This critical-level anomaly acts in turn at a distance to dampen the interfacial waves. When the ambient stratification is increased so that the Richardson number exceeds the value of a quarter, this destructive interaction overwhelms the constructive interaction between the interfacial waves, and consequently the flow becomes stable. This effect is manifested when considering the different action-at-a-distance contributions to the energy flux divergence at the critical level. The interfacial-wave interaction is found to contribute towards divergence, that is, towards instability, whereas the critical-level–interfacial-wave interaction contributes towards an energy flux convergence, that is, towards stability.

Type
Papers
Copyright
Copyright © Cambridge University Press 2011

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

REFERENCES

Baines, P. G. 1998 Topographic Effects in Stratified Flows. Cambridge University Press.Google Scholar
Baines, P. G. & Mitsudera, H. 1994 On the mechanism of shear flow instabilities. J. Fluid Mech. 276, 327342.Google Scholar
Booker, J. R. & Bretherton, F. P. 1967 The critical layer for internal gravity waves in a shear flow. J. Fluid Mech. 27, 513519.Google Scholar
Bretherton, F. P. 1966 Baroclinic instability, the short wave cutoff in terms of potential vorticity. Q. J. R. Meteorol. Soc. 92, 335345.Google Scholar
Charney, J. G. & Stern, M. E. 1962 On the stability of internal baroclinic jets in a rotating atmosphere. J. Atmos. Sci. 19, 159172.Google Scholar
Davies, H. C. & Bishop, C. H. 1994 Eady edge waves and rapid development. J. Atmos. Sci. 51, 19301946.Google Scholar
Eady, E. T. 1949 Long waves and cyclone waves. Tellus 1, 3352.Google Scholar
Eliassen, A. & Palm, E. 1960 On the transfer of energy in stationary mountain waves. Geophys. Publ. 22, 123.Google Scholar
Fjørtoft, R. 1950 Application of integral theorems in deriving criteria of stability for laminar flows and for the baroclinic circular vortex. Geofys. Publ. 17, 152.Google Scholar
Goldstein, S. 1931 On the stability of superposed streams of fluids of different densities. Proc. R. Soc. Lond. A 132, 524548.Google Scholar
Harnik, N. & Heifetz, E. 2007 Relating over-reflection and wave geometry to the counter propagating Rossby wave perspective: toward a deeper mechanistic understanding of shear instability. J. Atmos. Sci. 64, 22382261.Google Scholar
Harnik, N., Heifetz, E., Umurhan, O. M. & Lott, F. 2008 A buoyancy–vorticity wave interaction approach to stratified shear flow. J. Atmos. Sci. 65, 26152630.Google Scholar
Hayashi, Y. Y. & Young, W. R. 1986 Stable and unstable shear modes of rotating parallel flows in shallow water. J. Fluid Mech. 184, 477504.Google Scholar
Heifetz, E., Bishop, C. H., Hoskins, B. J. & Alpert, P. 1999 Counter-propagating Rossby waves in barotropic Rayleigh model of shear instability. Q. J. R. Meteorol. Soc. 125, 28352853.Google Scholar
Heifetz, E., Bishop, C. H., Hoskins, B. J. & Metheven, J. 2004 The counter-propagating Rossby-wave perspective on baroclinic instability. Part I. Mathematical basis. Q. J. R. Meteorol. Soc. 130, 211232.Google Scholar
Heifetz, E., Harnik, N. & Tamarin, T. 2009 A Hamiltonian counter-propagating Rossby wave perspective of pseudoenergy in the context of shear instability. Q. J. R. Meteorol. Soc. 135, 21612167.Google Scholar
Heifetz, E. & Methven, J. 2005 Relating optimal growth to counterpropagating Rossby waves in shear instability. Phys. Fluids 17, 064107.Google Scholar
Holmboe, J. 1962 On the behaviour of symmetric waves in stratified shear layers. Geofys. Publ. 24, 67113.Google Scholar
Hoskins, B. J., Mcintyre, M. E. & Robertson, A. W. 1985 On the use and significance of isentropic potential vorticity maps. Q. J. R. Meteorol. Soc. 111, 877946.Google Scholar
Howard, L. N. 1961 Note on paper of John W. Miles. J. Fluid Mech. 10, 509512.Google Scholar
Knessl, C. & Keller, J. B. 1992 Stability of rotating shear flows in shallow water. J. Fluid Mech. 244, 605614.Google Scholar
Lindzen, R. S. & Barker, J. W. 1985 Instability and wave over-reflection in stably stratified shear flow. J. Fluid Mech. 151, 189217.Google Scholar
Lindzen, R. S. & Rambaldi, S. 1986 A study of over-reflection in viscous poiseuille flow. J. Fluid Mech. 165, 355372.Google Scholar
Lindzen, R. S. & Rosenthal, A. J. 1981 A WKB asymptotic analysis of baroclinic instability. J. Atmos. Sci. 38, 619629.Google Scholar
Lindzen, R. S. & Tung, K. K. 1978 Wave overreflection and shear instability. J. Atmos. Sci. 35, 16261632.Google Scholar
Metheven, J., Hoskins, B. J., Heifetz, E. & Bishop, C. H. 2005 The counter-propagating Rossby-wave perspective on baroclinic instability. Part IV. Nonlinear life cycles. Q. J. R. Meteorol. Soc. 131, 14251440.Google Scholar
Miles, J. W. 1961 On the stability of heterogeneous shear flows. J. Fluid Mech. 10, 496508.Google Scholar
Rayleigh, Lord 1880 On the stability, or instability, of certain fluid motions. Proc. Lond. Math. Soc. 9, 5770.Google Scholar
Sakai, S. 1989 Rossby–Kelvin instability: a new type of ageostrophic instability caused by a resonance between Rossby waves and gravity waves. J. Fluid Mech. 202, 149176.Google Scholar
Shepherd, T. G. 1990 Symmetries, conservation laws, and Hamiltonian structure in geophysical fluid dynamics. Adv. Geophys. 32, 287352.Google Scholar
Sneddon, I. N. 1956 Special Functions of Mathematical Physics and Chemistry. Oliver and Boyd.Google Scholar
Taylor, G. I. 1931 Effect of variation in density on the stability of superposed streams of fluid. Proc. R. Soc. Lond. A 132, 499523.Google Scholar