Hostname: page-component-7c8c6479df-fqc5m Total loading time: 0 Render date: 2024-03-28T14:45:10.396Z Has data issue: false hasContentIssue false

Low Rossby limiting dynamics for stably stratified flow with finite Froude number

Published online by Cambridge University Press:  27 April 2011

BETH A. WINGATE*
Affiliation:
MS D413, Los Alamos National Laboratory, Los Alamos, NM 87544, USA
PEDRO EMBID
Affiliation:
Department of Mathematics and Statistics, the University of New Mexico, Albuquerque, NM 87131, USA
MIRANDA HOLMES-CERFON
Affiliation:
New York University, Courant Institute of Mathematical Sciences, New York, NY 10012, USA
MARK A. TAYLOR
Affiliation:
Sandia National Laboratories, Albuquerque, NM 87185, USA
*
Email address for correspondence: wingate@lanl.gov
Rights & Permissions [Opens in a new window]

Abstract

Core share and HTML view are not available for this content. However, as you have access to this content, a full PDF is available via the ‘Save PDF’ action button.

In this paper, we explore the strong rotation limit of the rotating and stratified Boussinesq equations with periodic boundary conditions when the stratification is order 1 ([Rossby number] Ro = ε, [Froude number] Fr = O(1), as ε → 0). Using the same framework of Embid & Majda (Geophys. Astrophys. Fluid Dyn., vol. 87, 1998, p. 1), we show that the slow dynamics decouples from the fast. Furthermore, we derive equations for the slow dynamics and their conservation laws. The horizontal momentum equations reduce to the two-dimensional Navier–Stokes equations. The equation for the vertically averaged vertical velocity includes a term due to the vertical average of the buoyancy. The buoyancy equation, the only variable to retain its three-dimensionality, is advected by all three two-dimensional slow velocity components. The conservation laws for the slow dynamics include those for the two-dimensional Navier–Stokes equations and a new conserved quantity that describes dynamics between the vertical kinetic energy and the buoyancy. The leading order potential enstrophy is slow while the leading order total energy retains both fast and slow dynamics. We also perform forced numerical simulations of the rotating Boussinesq equations to demonstrate support for three aspects of the theory in the limit Ro → 0: (i) we find the formation and persistence of large-scale columnar Taylor–Proudman flows in the presence of O(1) Froude number; after a spin-up time, (ii) the ratio of the slow total energy to the total energy approaches a constant and that at the smallest Rossby numbers that constant approaches 1 and (iii) the ratio of the slow potential enstrophy to the total potential enstrophy also approaches a constant and that at the lowest Rossby numbers that constant is 1. The results of the numerical simulations indicate that even in the presence of the low wavenumber white noise forcing the dynamics exhibit characteristics of the theory.

Type
Papers
Copyright
Copyright © Cambridge University Press 2011. The online version of this article is published within an Open Access environment subject to the conditions of the Creative Commons Attribution-NonCommercial-ShareAlike licence <http://creativecommons.org/licenses/by-nc-sa/2.5/>. The written permission of Cambridge University Press must be obtained for commercial re-use.

References

REFERENCES

Babin, A., Mahalov, A. & Nicolaenko, B. 1996 Global splitting, integrability and regularity of 3D Euler and Navier–Stokes equations for uniformly rotating fluids. Eur. J. Mech. B/Fluids 15 (3), 291300.Google Scholar
Babin, A., Mahalov, A., Nicolaenko, B. & Zhou, Y. 1997 On the asymptotic regimes and the strongly stratified limit of rotating boussinesq equations. Theor. Comput. Fluid Dyn. 9 (3/4), 223251.Google Scholar
Babin, A., Mahalov, A. & Nicolaenko, 1998 On nonlinear baroclinic waves and adjustment of pancake dynamics. Theor. Comput. Fluid Dyn. 11 (3/4), 215235.Google Scholar
Babin, A., Mahalov, A. & Nicolaenko, B. 2002 Fast singular oscillating limits of stably stratified three-dimensional Euler and Navier–Stokes equations and ageostrophic fronts. In Large-Scale Atmosphere-Ocean Dynamics (ed. Hunt, J., Norbury, J. & Roulstone, I.) pp. 126201. Cambridge University Press.Google Scholar
Charney, J. G. 1948 On the scale of atmospheric motions. Geofysiske Publikasjoner 17 (2), 3.Google Scholar
Chasnov, J. R. 1994 Similarity states of passive scalar transport in isotropic turbulence. Phys. Fluids 6, 10361051.Google Scholar
Chen, Q., Chen, S., Eyink, G. L. & Holm, D. D. 2005 Resonant interactions in rotating homogeneous three-dimensional turbulence. J. Fluid Mech. 542, 139164.Google Scholar
Davidson, P. A., Staplehurst, P. J. & Dalziel, S. B. 2006 On the evolution of eddies in a rapidly rotating system. J. Fluid Mech. 557, 135144.Google Scholar
Davies, P. A. 1972 Experiments on Taylor columns in rotating, stratified fluids. J. Fluid Mech. 54, 691718.Google Scholar
Embid, P. F. & Majda, A. J. 1996 Averaging over fast gravity waves for geophysical flows with arbitrary potential vorticity. Commun. Part. Diff. Equ. 21 (3–4), 619658.Google Scholar
Embid, P. F. & Majda, A. J. 1998 Low Froude number limiting dynamics for stably stratified flow with small or finite Rossby numbers. Geophys. Astrophys. Fluid Dyn. 87 (1–2), 150.Google Scholar
Emery, W. J., Lee, W. G. & Magaard, L. 1984 Geographic and seasonal distributions of Brunt–Vaisala frequency and Rossby radii in the North Pacific and North Atlantic. J. Phys. Oceanogr. 14, 294317.Google Scholar
Greenspan, H. P. 1990 The Theory of Rotating Fluids. Breukelen Press.Google Scholar
Heywood, K. J., Garabato, A. N. & Stevens, D. P. 2002 High mixing rates in the abyssal Southern ocean. Nature 415, 10111014.Google Scholar
Jones, E. P., Rudels, B. & Anderson, L. G. 1995 Deep waters of the Arctic Ocean: origins and circulation. Deep-Sea Res. 42, 737760.Google Scholar
Kelvin, L. W. 1880 Vibrations of a columnar vortex. Phil. Mag. 10, 155168.Google Scholar
Klainerman, S. & Majda, A. 1981 Singular limits of quasilinear hyperbolic systems with large parameters and the incompressible limit of compressible fluids. Commun. Pure Appl. Math. 34 (4), 481524.Google Scholar
Lax, P. 2002 Functional Analysis. Wiley-Interscience.Google Scholar
LeBlond, P. H. & Mysak, L. A. 1978 Waves in the Ocean. Elsevier Oceanography Series 20. Elsevier Scientific Publishing Company.Google Scholar
Majda, A. 1984 Compressible Fluid Flow and Systems of Conservation Laws in Several Space Variables. Springer-Verlag.Google Scholar
Majda, A. J. & Grote, M. J. 1997 Model dynamics and vertical collapse in decaying strongly stratified flows. Phys. Fluids 9 (10), 29322940.Google Scholar
Majda, A. J. & Embid, P. 1998 Averaging over fast gravity waves for geophysical flows with unbalanced initial data. Theor. Comput. Fluid Dyn. 11 (3/4), 155169.Google Scholar
Otheguy, P., Chomaz, J. & Billant, P. 2006 Elliptic and zigzag instabilities on co-rotating vertical vortices in a stratified fluid. J. Fluid Mech. 553, 253272.Google Scholar
Potylitsin, P. G. & Peltier, W. R. 1998 Stratification effects on the stability of columnar vortices on the f-plane. J. Fluid Mech. 355, 4579.Google Scholar
Potylitsin, P. G. & Peltier, W. R. 2003 On the nonlinear evolution of columnar vortices in a rotating environment. Geophys. Astrophys. Fluid Dyn. 97, 365391.Google Scholar
Riley, J. J. & deBruynKops, S. M. 2003 Dynamics of turbulence strongly influenced by buoyancy. Phys. Fluids 15 (7), 20472059.Google Scholar
Riley, J. J. & Lelong, M. P. 2000 Fluid motions in the presence of strong stable stratification. Annu. Rev. Fluid Mech. 32, 613657.Google Scholar
Schochet, S. 1987 Singular limits in bounded domains for quasilinear symmetric hyperbolic systems having a vorticity equation. J. Diff. Equ. 68 (3), 400428.Google Scholar
Schochet, S. 1994 Fast singular limits of hyperbolic pde's. J. Dif. Equ. 114, 476512.Google Scholar
Smith, L. M. & Waleffe, F. 2002 Generation of slow large scales in forced rotating stratified turbulence. J. Fluid Mech. 451, 145168.Google Scholar
Sreenivasan, B. & Davidson, P. A. 2008 On the formation of cyclones and anticyclones in a rotating fluid. Phys. Fluids 20, 085104.Google Scholar
Staplehurst, P. J., Davidson, P. A. & Dalziel, S. B. 2008 Structure formation in homogeneous freely decaying rotating turbulence. J. Fluid Mech. 598, 81105.Google Scholar
Taylor, G. I. 1921 Experiments with rotating fluids. Proc. R. Soc. Lond. 100, 114121.Google Scholar
Timmermans, M. L., Melling, H. & Rainville, L. 2007 Dynamics in the deep Canada Basin, Arctic Ocean, inferred by thermistor-chain time series. J. Phys. Oceanogr. 37, 10661076.Google Scholar
Timmermans, M. L., Rainville, L., Thomas, L. & Proshutinsky, A. 2010 Moored observations of bottom-intensified motions in the deep Canada Basin. J. Mar. Res. 68 (1), 625641.Google Scholar
Woodgate, R., Aagaard, K., Muench, R., Gunn, J., Björk, G., Rudels, B., Roach, A. T. & Schauer, U. 2001 The Arctic Ocean boundary current along the Eurasian slope and the adjacent Lomonosov Ridge: Water mass properties, transports and transformations from moored instruments. Deep-Sea Res. 48, 17571892.Google Scholar
Van Haren, H. & Millot, C. 2005 Gyroscopic waves in the Mediterranean Sea. Geophys. Res. Lett. 32 (24), 14.Google Scholar