Hostname: page-component-848d4c4894-mwx4w Total loading time: 0 Render date: 2024-07-02T02:28:06.674Z Has data issue: false hasContentIssue false

The effect of Reynolds number on mixing in Kelvin–Helmholtz billows

Published online by Cambridge University Press:  28 October 2014

M. Rahmani*
Affiliation:
Department of Civil Engineering, University of British Columbia, Vancouver, BC V6T 1Z4, Canada
G. A. Lawrence
Affiliation:
Department of Civil Engineering, University of British Columbia, Vancouver, BC V6T 1Z4, Canada
B. R. Seymour
Affiliation:
Department of Mathematics, University of British Columbia, Vancouver, BC V6T 1Z2, Canada
*
Email address for correspondence: mona.rahmani@alumni.ubc.ca

Abstract

Mixing induced through the life-cycle of Kelvin–Helmholtz (KH) billows is studied for a range of low and intermediate Reynolds numbers using direct numerical simulations (DNS). The amount of stirring, and therefore mixing, is significantly controlled by the process of vortex pairing of two KH billows. For low Reynolds numbers, vortex pairing of the billows is complete in the pre-turbulent stage or early stages of turbulence, generating a high amount of stirring. At higher Reynolds numbers, vortex pairing is suppressed by the growth of three-dimensional instabilities, and the amount of stirring is significantly reduced. For single KH billows, as the Reynolds number increases, there is a transition in the characteristics of the mixing, similar to the laboratory measurements of Breidenthal (J. Fluid Mech., vol. 109, 1981, pp. 1–24) and Koochesfahani & Dimotakis (J. Fluid Mech., vol. 170, 1986, pp. 83–112). The transition in mixing is associated with the growth and sustainability of three-dimensional motions at sufficiently high Reynolds numbers. We examine this ‘mixing transition’ and the influence of vortex pairing on it by examining the flow properties at different stages and the exchange between the energy partitions. As the Reynolds number increases, three-dimensional motions develop over a wider range of length scales, and smaller scale eddies form. However, this does not necessarily result in a greater amount of mixing. The maximum total amount of mixing induced over the lifetime of a KH instability, for billows both with and without vortex pairing, occurs when the large-scale eddies that cause the stirring are the most energetic. The mixing efficiency reveals a non-monotonic dependence on the Reynolds number.

Type
Papers
Copyright
© 2014 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

Arratia, C., Caulfield, C. P. & Chomaz, J.-M. 2013 Transient perturbation growth in time-dependent mixing layers. J. Fluid Mech. 717, 90133.Google Scholar
Atsavapranee, P. & Gharib, M. 1997 Structures in stratified plane mixing layers and the effects of cross-shear small-scale variation of convected quantities like temperature in turbulent fluid. J. Fluid Mech. 342, 5386.Google Scholar
Barenblatt, G. I., Bertsch, M., Passo, R. D., Prostokishin, V. M. & Uchi, M. 1993 A mathematical model of turbulent heat and mass transfer in stably stratified shear flow. J. Fluid Mech. 253, 341358.Google Scholar
Bernal, L. P. & Roshko, A. 1986 Streamwise vortex structure in plane mixing layers. J. Fluid Mech. 170, 499525.Google Scholar
Breidenthal, R. 1981 Structure in turbulent mixing layers and wakes using a chemical reaction. J. Fluid Mech. 109, 124.Google Scholar
Carpenter, J. R., Lawrence, G. A. & Smyth, W. D. 2007 Evolution and mixing of asymmetric Holmboe instabilities. J. Fluid Mech. 582, 103132.CrossRefGoogle Scholar
Caulfield, C. P. & Peltier, W. R. 2000 The anatomy of the mixing transition in homogeneous and stratified free shear layers. J. Fluid Mech. 413, 147.Google Scholar
Corcos, G. M. & Sherman, F. S. 1976 Vorticity concentration and the dynamics of unstable free shear layers. J. Fluid Mech. 73, 241264.Google Scholar
Cortesi, A. B., Smith, B. L., Yadigaroglu, G. & Banerjee, S. 1999 Numerical investigation of the entrainment and mixing processes in neutral and stably-stratified mixing layers. Phys. Fluids 11, 162184.Google Scholar
DeSilva, I. P. D., Fernando, H. J. S., Eaton, F. & Hebert, D. 1996 Evolution of Kelvin–Helmholtz billows in nature and laboratory. Earth Planet. Sci. Lett. 143, 217231.CrossRefGoogle Scholar
Dimotakis, P. E. 2000 The mixing transition in turbulent flows. J. Fluid Mech. 409, 6998.CrossRefGoogle Scholar
Dimotakis, P. E. 2005 Turbulent mixing. Annu. Rev. Fluid Mech. 37, 329356.Google Scholar
Drazin, P. G. & Reid, W. H. 1981 Hydrodynamic Stability. Cambridge University Press.Google Scholar
Forryan, A., Martin, A. P., Srokosz, M. A., Popova, E. E., Painter, S. C. & Renner, A. H. H. 2013 A new observationally motivated Richardson number based mixing parametrization for oceanic mesoscale flow. J. Geophys. Res. 118 (3), 14051419.Google Scholar
Geyer, W. R., Lavery, A. C., Scully, M. E. & Trowbridge, J. H. 2010 Mixing by shear instability at high Reynolds number. Geophys. Res. Lett. 37, L22607.Google Scholar
Haigh, S. P.1995 Non-symmetric Holmboe waves. PhD thesis, the University of British Columbia.Google Scholar
Inoue, R. & Smyth, W. D. 2009 Efficiency of mixing forced by unsteady shear flow. J. Phys. Oceanogr. 39, 11501166.Google Scholar
Klaassen, G. P. & Peltier, W. R. 1985a The effect of Prandtl number on the evolution and stability of Kelvin–Helmholtz billows. Geophys. Astrophys. Fluid Dyn. 32, 2360.Google Scholar
Klaassen, G. P. & Peltier, W. R. 1985b The onset of turbulence in finite-amplitude Kelvin–Helmholtz billows. J. Fluid Mech. 155, 135.Google Scholar
Klaassen, G. P. & Peltier, W. R. 1991 The influence of stratification on secondary instability in free shear layers. J. Fluid Mech. 227, 71106.Google Scholar
Konrad, J. H.1976 An experimental investigation of mixing in two-dimensional turbulent shear flows with applications to diffusion-limited chemical reactions. PhD thesis, California Institute of Technology.Google Scholar
Koochesfahani, M. M. & Dimotakis, P. E. 1986 Mixing and chemical reaction in a turbulent liquid mixing layer. J. Fluid Mech. 170, 83112.CrossRefGoogle Scholar
Koop, C. G. & Browand, F. K. 1979 Instability and turbulence in a stratified layer with shear. J. Fluid Mech. 93, 135159.Google Scholar
Linden, P. F. 1979 Mixing in stratified fluids. Geophys. Astrophys. Fluid Dyn. 13, 323.Google Scholar
Lorenz, E. N. 1955 Available potential energy and the maintenance of the general circulation. Tellus VII 2, 157167.Google Scholar
Mashayek, A., Caulfield, C. P. & Peltier, W. R. 2013 Time-dependent, non-monotonic mixing in stratified turbulent shear flows: implications for oceanographic estimates of buoyancy flux. J. Fluid Mech. 736, 570593.Google Scholar
Mashayek, A. & Peltier, W. R. 2012a The ‘zoo’ of secondary instabilities precursory to stratified shear flow transition. Part 1. Shear aligned convection, pairing, and braid instabilities. J. Fluid Mech. 708, 544.Google Scholar
Mashayek, A. & Peltier, W. R. 2012b The ‘zoo’ of secondary instabilities precursory to stratified shear flow transition. Part 2. The influence of stratification. J. Fluid Mech. 708, 4570.Google Scholar
Mashayek, A. & Peltier, W. R. 2013 Shear-induced mixing in geophysical flows: does the route to turbulence matter to its efficiency? J. Fluid Mech. 725, 216261.Google Scholar
Moin, P. & Mahesh, K. 1998 Direct numerical simulation: a tool in turbulence research. Annu. Rev. Fluid Mech. 30, 539578.Google Scholar
Moser, R. D. & Rogers, M. M. 1991 Mixing transition and the cascade to small scales in a plane mixing layer. Phys. Fluids A 3, 11281134.Google Scholar
Moser, R. D. & Rogers, M. M. 1993 The three-dimensional evolution of a plane mixing layer: pairing and transition to turbulence. J. Fluid Mech. 247, 275320.Google Scholar
Moum, J. N. 1996 Energy-containing scales of turbulence in the ocean thermocline. J. Geophys. Res. 101 (C6), 1409514109.Google Scholar
Moum, J. N., Farmer, D. M., Smyth, W. D., Armi, L. & Vagle, S. 2003 Structure and generation of turbulence at interfaces strained by solitary waves propagating shoreward over the continental shelf. J. Phys. Oceanogr. 33, 20932112.Google Scholar
Osborn, T. R. 1980 Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr. 10 (1), 8389.Google Scholar
Patterson, M. D., Caulfield, C. P., McElwaine, J. N. & Dalziel, S. B. 2006 Time-dependent mixing in stratified Kelvin–Helmholtz billows: experimental observations. Geophys. Res. Lett. 33, L15608.Google Scholar
Peltier, W. R. & Caulfield, C. P. 2003 Mixing efficiency in stratified shear flows. Annu. Rev. Fluid Mech. 35, 135167.Google Scholar
Peters, H., Gregg, M. C. & Toole, J. M. 1988 On the parameterization of equatorial turbulence. J. Geophys. Res. 93 (C2), 11991218.Google Scholar
Phillips, O. M. 1972 Turbulence in a strongly stratified fluid – is it unstable? Deep-Sea Res. 19, 7981.Google Scholar
Posmentier, E. S. 1977 The generation of salinity finestructure by vertical diffusion. J. Phys. Oceanogr. 7, 298300.2.0.CO;2>CrossRefGoogle Scholar
Rahmani, M., Seymour, B. & Lawrence, G. 2014 The evolution of large and small-scale structures in Kelvin–Helmholtz instabilities. Environ. Fluid Mech. 127.Google Scholar
Scinocca, J. F. 1995 The mixing of mass and momentum by Kelvin–Helmholtz billows. J. Atmos. Sci. 52, 25092530.Google Scholar
Seim, H. E. & Gregg, M. C. 1994 Detailed observations of a naturally occuring shear instability. J. Geophys. Res. 99, 1004910073.Google Scholar
Seim, H. E. & Gregg, M. C. 1995 Energetics of a naturally occuring shear instability. J. Geophys. Res. 100 (C3), 49434958.CrossRefGoogle Scholar
Smyth, W. D. 1999 Dissipation range geometry and scalar spectra in sheared, stratified turbulence. J. Fluid Mech. 401, 209242.Google Scholar
Smyth, W. D., Carpenter, J. R. & Lawrence, G. A. 2007 Mixing in symmetric Holmboe waves. J. Phys. Oceanogr. 37, 15661583.Google Scholar
Smyth, W. D. & Moum, J. N. 2000 Length scales of turbulence in stably stratified mixing layers. Phys. Fluids 12, 13271342.Google Scholar
Smyth, W. D. & Moum, J. N. 2012 Ocean mixing by Kelvin–Helmholtz instability. Oceanography 25 (2), 140149.Google Scholar
Smyth, W., Moum, J. & Caldwell, D. 2001 The efficiency of mixing in turbulent patches: inferences from direct simulations and microstructure observations. J. Phys. Oceanogr. 31, 19691992.Google Scholar
Smyth, W. D., Nash, J. D. & Moum, J. N. 2005 Differential diffusion in breaking Kelvin–Helmholtz billows. J. Phys. Oceanogr. 35, 10041022.Google Scholar
Smyth, W. D. & Winters, K. B. 2003 Turbulence and mixing in Holmboe waves. J. Phys. Oceanogr. 33, 694711.Google Scholar
Staquet, C. 2000 Mixing in a stably stratified shear layer: two- and three-dimensional numerical experiments. Fluid Dyn. Res. 27, 367404.Google Scholar
Tailleux, R. 2013 Available potential energy and exergy in stratified fluids. Annu. Rev. Fluid Mech. 45 (1), 3558.Google Scholar
Tennekes, H. & Lumley, J. L. 1972 A First Course in Turbulence. The MIT Press.Google Scholar
Thorpe, S. A. 1973 Experiments on instability and turbulence in stratified shear flows. J. Fluid Mech. 61, 731751.Google Scholar
Wesson, J. C. & Gregg, M. C. 1994 Mixing at camarinal sill in the Strait of Gibraltar. J. Geophys. Res. 99 (C5), 98479878.Google Scholar
Winant, C. D. & Browand, F. K. 1974 Vortex pairing: the mechanism of turbulent mixing layer growth at moderate Reynolds number. J. Fluid Mech. 63, 237255.Google Scholar
Winters, K. B., Lombard, P. N., Riley, J. J. & D’asaro, E. A. 1995 Available potential energy and mixing in density-stratified fluids. J. Fluid Mech. 289, 115128.Google Scholar
Winters, K. B., MacKinnon, J. A. & Mills, B. 2004 A spectral model for process studies of rotating, density-stratified flows. J. Atmos. Ocean. Technol. 21, 6994.Google Scholar
Woods, J. D. 1968 Wave-induced shear instability in the summer thermocline. J. Fluid Mech. 32, 791800.Google Scholar