Hostname: page-component-76fb5796d-22dnz Total loading time: 0 Render date: 2024-04-25T12:27:34.294Z Has data issue: false hasContentIssue false

Balance dynamics in rotating stratified turbulence

Published online by Cambridge University Press:  22 April 2016

Hossein A. Kafiabad*
Affiliation:
Department of Atmospheric and Oceanic Sciences, McGill University, 805 Sherbrooke ouest, Montreal, Quebec H3A 0B9, Canada
Peter Bartello
Affiliation:
Department of Atmospheric and Oceanic Sciences, McGill University, 805 Sherbrooke ouest, Montreal, Quebec H3A 0B9, Canada Department of Mathematics and Statistics, McGill University, 805 Sherbrooke ouest, Montreal, Quebec H3A 0B9, Canada
*
Email address for correspondence: hossein.aminikafiabad@mail.mcgill.ca

Abstract

If classical quasigeostrophic (QG) flow breaks down at smaller scales, it gives rise to questions of whether higher-order nonlinear balance can be maintained, to what scale and for how long. These are naturally followed by asking how this is affected by stratification and rotation. To address these questions, we perform non-hydrostatic Boussinesq simulations where the initial data is balanced using the Baer–Tribbia nonlinear normal mode initialization scheme (NNMI), which is accurate to second order in the Rossby number, as the next-order improvement to first-order QG theory. The NNMI procedure yields an ageostrophic contribution to the energy spectrum that has a very steep slope. However, as time passes, a shallow range emerges in the ageostrophic spectrum when the Rossby number is large enough for a given Reynolds number. It is argued that this shallow range is the unbalanced part of the motion that develops spontaneously in time and eventually dominates the energy at small scales. If the initial flow is not nonlinearly balanced, the shallow range emerges at even lower Rossby number and it appears at larger scales. Through numerous simulations at different rotation and stratification, this study gives a clear picture of how energy is cascaded in different initially balanced regimes of rotating stratified flow. We find that at low Rossby number the flow mainly consists of a geostrophic part and a balanced ageostrophic part with a steep spectrum. As the Rossby number increases, the unbalanced part of the ageostrophic energy increases at a rate faster than the balanced part. Hence, the total energy spectrum displays a shallow range above a transition wavenumber. This wavenumber evolves to smaller values as rotation weakens.

Type
Papers
Copyright
© 2016 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

Baer, F. & Tribbia, J. J. 1977 On complete filtering of gravity modes through nonlinear initialization. Mon. Weath. Rev. 105 (12), 15361539.2.0.CO;2>CrossRefGoogle Scholar
Bartello, P. 1995 Geostrophic adjustment and inverse cascades in rotating stratified turbulence. J. Atmos. Sci. 52 (24), 44104428.Google Scholar
Bartello, P. 2010 Quasigeostrophic and stratified turbulence in the atmosphere. In IUTAM Symposium on Turbulence in the Atmosphere and Oceans, pp. 117130. Springer.CrossRefGoogle Scholar
Bartello, P., Métais, O. & Lesieur, M. 1996 Geostrophic versus wave eddy viscosities in atmospheric models. J. Atmos. Sci. 53 (4), 564571.2.0.CO;2>CrossRefGoogle Scholar
Bartello, P. & Tobias, S. M. 2013 Sensitivity of stratified turbulence to the buoyancy Reynolds number. J. Fluid Mech. 725, 122.Google Scholar
Brethouwer, G., Billant, P., Lindborg, E. & Chomaz, J. M. 2007 Scaling analysis and simulation of strongly stratified turbulent flows. J. Fluid Mech. 585, 343368.CrossRefGoogle Scholar
Callies, J., Ferrari, R. & Bühler, O. 2014 Transition from geostrophic turbulence to inertia–gravity waves in the atmospheric energy spectrum. Proc. Natl Acad. Sci. USA 111 (48), 1703317038.CrossRefGoogle ScholarPubMed
Charney, J. G. 1971 Geostrophic turbulence. J. Atmos. Sci. 28 (6), 10871095.2.0.CO;2>CrossRefGoogle Scholar
Cole, J. D. 1968 Perturbation Methods in Applied Mathematics. Blaisdell.Google Scholar
Deusebio, E., Vallgren, A. & Lindborg, E. 2013 The route to dissipation in strongly stratified and rotating flows. J. Fluid Mech. 720, 66103.Google Scholar
Dritschel, D. G. & Mckiver, W. J. 2015 Effect of Prandtl’s ratio on balance in geophysical turbulence. J. Fluid Mech. 777, 569590.Google Scholar
Dritschel, D. G. & Viúdez, A. 2003 A balanced approach to modelling rotating stably stratified geophysical flows. J. Fluid Mech. 488, 123150.CrossRefGoogle Scholar
Evans, K. J., Lauritzen, P. H., Mishra, S. K., Neale, R. B., Taylor, M. A. & Tribbia, J. J. 2013 AMIP simulation with the CAM4 Spectral Element Dynamical Core. J. Clim. 26 (3), 689709.Google Scholar
Ford, R., McIntyre, M. E. & Norton, W. A. 2000 Balance and the slow quasimanifold: some explicit results. J. Atmos. Sci. 57 (9), 12361254.Google Scholar
Gage, K. S. 1979 Evidence for a k -5/3 law inertial range in mesoscale two-dimensional turbulence. J. Atmos. Sci. 36 (10), 19501954.2.0.CO;2>CrossRefGoogle Scholar
Gage, K. S. & Nastrom, G. D. 1986 Theoretical interpretation of atmospheric wavenumber spectra of wind and temperature observed by commercial aircraft during GASP. J. Atmos. Sci. 43 (7), 729740.Google Scholar
Hakim, G. J. 2000 Climatology of coherent structures on the extratropical tropopause. Mon. Weath. Rev. 128 (2), 385406.Google Scholar
Hamilton, K., Takahashi, Y. O. & Ohfuchi, W. 2008 Mesoscale spectrum of atmospheric motions investigated in a very fine resolution global general circulation model. J. Geophys. Res. 113 (D18), D18110.Google Scholar
Hoskins, B. J. & Hodges, K. I. 2002 New perspectives on the Northern Hemisphere winter storm tracks. J. Atmos. Sci. 59 (6), 10411061.Google Scholar
Kreiss, H.-O. & Lorenz, J. 1994 On the existence of slow manifolds for problems with different timescales. Phil. Trans. R. Soc. Lond. A 346 (1679), 159171.Google Scholar
Leith, C. E. 1980 Nonlinear normal mode initialization and quasi-geostrophic theory. J. Atmos. Sci. 37 (5), 958968.Google Scholar
Lindborg, E. 2006 The energy cascade in a strongly stratified fluid. J. Fluid Mech. 550, 207242.Google Scholar
Lorenz, E. N. 1980 Attractor sets and quasi-geostrophic equilibrium. J. Atmos. Sci. 37 (8), 16851699.Google Scholar
Lorenz, E. N. 1986 On the existence of a slow manifold. J. Atmos. Sci. 43 (15), 15471558.2.0.CO;2>CrossRefGoogle Scholar
Lorenz, E. N. 1992 The slow manifold-what is it? J. Atmos. Sci. 49 (24), 24492451.Google Scholar
Lorenz, E. N. & Krishnamurthy, V. 1987 On the nonexistence of a slow manifold. J. Atmos. Sci. 44 (20), 29402950.Google Scholar
Marino, R., Mininni, P. D., Rosenberg, D. & Pouquet, A. 2013 Inverse cascades in rotating stratified turbulence: fast growth of large scales. Europhys. Lett. 102 (4), 44006.Google Scholar
Mckiver, W. J. & Dritschel, D. G. 2008 Balance in non-hydrostatic rotating stratified turbulence. J. Fluid Mech. 596, 201219.Google Scholar
Mcwilliams, J. C. 1989 Statistical properties of decaying geostrophic turbulence. J. Fluid Mech. 198, 199230.CrossRefGoogle Scholar
Mcwilliams, J. C., Weiss, J. B. & Yavneh, I. 1994 Anisotropy and coherent vortex structures in planetary turbulence. Science 264 (5157), 410413.Google Scholar
Mcwilliams, J. C., Weiss, J. B. & Yavneh, I. 1999 The vortices of homogeneous geostrophic turbulence. J. Fluid Mech. 401, 126.Google Scholar
Melander, M. V., Zabusky, N. J. & Mcwilliams, J. C. 1988 Symmetric vortex merger in two dimensions: causes and conditions. J. Fluid Mech. 195, 303340.Google Scholar
Nadiga, B. T. 2014 Nonlinear evolution of a baroclinic wave and imbalanced dissipation. J. Fluid Mech. 756, 9651006.Google Scholar
Polvani, L. M. 1991 Two-layer geostrophic vortex dynamics. Part 2. Alignment and two-layer V-states. J. Fluid Mech. 225, 241270.Google Scholar
Polvani, L. M., Zabusky, N. J. & Flierl, G. R. 1989 Two-layer geostrophic vortex dynamics. Part 1. Upper-layer V-states and merger. J. Fluid Mech. 205, 215242.Google Scholar
Pouquet, A. & Marino, R. 2013 Geophysical turbulence and the duality of the energy flow across scales. Phys. Rev. Lett. 111 (23), 234501.Google Scholar
Reznik, G. M., Zeitlin, V. & Ben Jelloul, M 2001 Nonlinear theory of geostrophic adjustment. Part 1. Rotating shallow-water model. J. Fluid Mech. 445, 93120.Google Scholar
Skamarock, W. C. 2004 Evaluating mesoscale nwp models using kinetic energy spectra. Mon. Weath. Rev. 132 (12), 30193032.Google Scholar
Snyder, C., Muraki, D. J., Plougonven, R. & Zhang, F. 2007 Inertia-gravity waves generated within a dipole vortex. J. Atmos. Sci. 64 (12), 44174431.Google Scholar
Tulloch, R. & Smith, K. S. 2006 A theory for the atmospheric energy spectrum: depth-limited temperature anomalies at the tropopause. Proc. Natl Acad. Sci. USA 103 (40), 1469014694.Google Scholar
Vallgren, A., Deusebio, E. & Lindborg, E. 2011 Possible explanation of the atmospheric kinetic and potential energy spectra. Phys. Rev. Lett. 107 (26), 268501.Google Scholar
Vanneste, J. 2013 Balance and spontaneous wave generation in geophysical flows. Annu. Rev. Fluid Mech. 45, 147172.CrossRefGoogle Scholar
Vanneste, J. & Yavneh, I. 2004 Exponentially small inertia-gravity waves and the breakdown of quasigeostrophic balance. J. Atmos. Sci. 61 (2), 211223.2.0.CO;2>CrossRefGoogle Scholar
Viúdez, Á. 2007 The origin of the stationary frontal wave packet spontaneously generated in rotating stratified vortex dipoles. J. Fluid Mech. 593, 359383.CrossRefGoogle Scholar
Viúdez, Á. 2008 The stationary frontal wave packet spontaneously generated in mesoscale dipoles. J. Phys. Oceanogr. 38 (1), 243256.Google Scholar
Viúdez, Á. & Dritschel, D. G. 2004 Optimal potential vorticity balance of geophysical flows. J. Fluid Mech. 521, 343352.Google Scholar
Waite, M. L. & Bartello, P. 2004 Stratified turbulence dominated by vortical motion. J. Fluid Mech. 517, 281308.Google Scholar
Waite, M. L. 2011 Stratified turbulence at the buoyancy scale. Phys. Fluids 23 (6), 066602.Google Scholar
Waite, M. L. & Snyder, C. 2009 The mesoscale kinetic energy spectrum of a baroclinic life cycle. J. Atmos. Sci. 66 (4), 883901.Google Scholar
Warn, T. 1997 Nonlinear balance and quasi-geostrophic sets. Atmos.-Ocean 35 (2), 135145.Google Scholar
Warn, T., Bokhove, O., Shepherd, T. G. & Vallis, G. K. 1995 Rossby number expansions, slaving principles, and balance dynamics. Q. J. R. Meteorol. Soc. 121 (523), 723739.CrossRefGoogle Scholar
Warn, T. & Menard, R. 1986 Nonlinear balance and gravity-inertial wave saturation in a simple atmospheric model. Tellus A 38 (4), 285294.Google Scholar
Whitehead, J. P. & Wingate, B. A. 2014 The influence of fast waves and fluctuations on the evolution of the dynamics on the slow manifold. J. Fluid Mech. 757, 155178.CrossRefGoogle Scholar
Zeitlin, V., Reznik, G. M. & Jelloul, M. 2003 Nonlinear theory of geostrophic adjustment. Part 2. Two-layer and continuously stratified primitive equations. J. Fluid Mech. 491, 207228.Google Scholar