Hostname: page-component-8448b6f56d-c47g7 Total loading time: 0 Render date: 2024-04-24T03:35:11.561Z Has data issue: false hasContentIssue false

Optimal perturbations in time-dependent variable-density Kelvin–Helmholtz billows

Published online by Cambridge University Press:  30 August 2016

Adriana Lopez-Zazueta
Affiliation:
Institut Supérieur de l’Aéronautique et de l’Espace (ISAE-SUPAERO), Université de Toulouse, 10 avenue Édouard Belin - BP 54032 - 31055 Toulouse CEDEX 4, France
Jérôme Fontane*
Affiliation:
Institut Supérieur de l’Aéronautique et de l’Espace (ISAE-SUPAERO), Université de Toulouse, 10 avenue Édouard Belin - BP 54032 - 31055 Toulouse CEDEX 4, France
Laurent Joly
Affiliation:
Institut Supérieur de l’Aéronautique et de l’Espace (ISAE-SUPAERO), Université de Toulouse, 10 avenue Édouard Belin - BP 54032 - 31055 Toulouse CEDEX 4, France
*
Email address for correspondence: j.fontane@isae.fr

Abstract

We analyse the influence of the specific features of time-dependent variable-density Kelvin–Helmholtz (VDKH) roll-ups on the development of three-dimensional secondary instabilities. Due to inertial (high Froude number) baroclinic sources of spanwise vorticity at high Atwood number (up to 0.5 here), temporally evolving mixing layers exhibit a layered structure associated with a strain field radically different from their homogeneous counterpart. We use a direct-adjoint non-modal linear approach to determine the fastest growing perturbations over a single period of the time-evolving two-dimensional base flow during a given time interval $[t_{0},T]$. When perturbations are seeded at the initial time of the primary KH mode growth, i.e. $t_{0}=0$, it is found that additional mechanisms of energy growth are onset around a bifurcation time $t_{b}$, a little before the saturation of the primary two-dimensional instability. The evolution of optimal perturbations is thus observed to develop in two distinct stages. Whatever the Atwood number, the first period $[t_{0},t_{b}]$ is characterised by a unique route for optimal energy growth resulting from a combination of the Orr and lift-up transient mechanisms. In the second period $[t_{b},T]$, the growing influence of mass inhomogeneities raises the energy gain over the whole range of spanwise wavenumbers. As the Atwood number increases, the short spanwise wavelength perturbations tend to benefit more from the onset of variable-density effects than large wavelength ones. The extra energy gain due to increasing Atwood numbers relies on contributions from spanwise baroclinic sources. The resulting vorticity field is structured into two elongated dipoles localised along the braid on either side of the saddle point. In return they yield two longitudinal velocity streaks of opposite sign which account for most of the energy growth. This transition towards three-dimensional motions is in marked contrast with the classic streamwise rib vortices, so far accepted as the paradigm for the transition of free shear flows, either homogeneous or not. It is argued that the emergence of these longitudinal velocity streaks is generic of the transition in variable-density shear flows. Among them, the light round jet is known to display striking side ejections as a result of the loss of axisymmetry. The present analysis helps to renew the question of the underlying flow structure behind side jets, otherwise based on radial induction between pairs of counter-rotating longitudinal vortices (Monkewitz & Pfizenmaier, Phys. Fluids A, vol. 3 (5), 1991, pp. 1356–1361). Instead, it is more likely that side ejections would result from the convergence of the longitudinal velocity streaks near the braid saddle point. When the injection time is delayed so as to suppress the initial stage of energy growth, a new class of perturbations arises at low wavenumber with energy gains far larger than those observed so far. They correspond to the two-dimensional Kelvin–Helmholtz secondary instability of the baroclinically enhanced vorticity braid discovered by Reinaud et al. (Phys. Fluids, vol. 12 (10), pp. 2489–2505), leading potentially to another route to turbulence through a two-dimensional fractal cascade.

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

Arratia, C., Caulfield, C. P. & Chomaz, J.-M. 2013 Transient perturbation growth in time-dependent mixing layers. J. Fluid Mech. 717, 90133.CrossRefGoogle Scholar
Batchelor, G. 1967 An Introduction to Fluid Mechanics. Cambridge University Press.Google Scholar
Brancher, P., Chomaz, J.-M. & Huerre, P. 1994 Direct numerical simulations of round jets: vortex induction and side jets. Phys. Fluids 6 (5), 17681774.CrossRefGoogle Scholar
Caulfield, C. P. & Kerswell, R. R. 2000 The nonlinear development of three-dimensional disturbances at hyperbolic stagnation points: a model of the braid region in mixing layers. Phys. Fluids 12 (5), 10321043.Google 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
Corbett, P. & Bottaro, A. 2000 Optimal perturbations for boundary layers subject to stream-wise pressure gradient. Phys. Fluids 12, 120130.Google Scholar
Corbett, P. & Bottaro, A. 2001 Optimal linear growth in swept bounday layers. J. Fluid Mech. 435, 123.Google Scholar
Corcos, G. M. & Lin, S. J. 1984 The mixing layer: deterministic models of a turbulent flow. Part 2. the origin of three-dimensional motion. J. Fluid Mech. 139, 6795.Google Scholar
Cortesi, A. B., Yadigaroglu, G. & Banerjee, S. 1998 Numerical investigation of the formation of three-dimensional structures in stably-stratified mixing layers. Phys. Fluids 10 (6), 14491473.CrossRefGoogle Scholar
Davey, R. F. & Roshkom, A. 1972 The effect of a density difference on shear-layer instability. J. Fluid Mech. 53, 523543.Google Scholar
Dixit, H. N. & Govindarajan, R. 2010 Vortex-induced instabilities and accelerated collapse due to inertial effects of density stratification. J. Fluid Mech. 646, 415439.Google Scholar
Dritschel, D. G., Haynes, P. H., Juckes, M. N. & Shepperd, T. G. 1991 The stability of a two-dimensional vorticity filament under uniform strain. J. Fluid Mech. 230, 647665.CrossRefGoogle Scholar
Ellingsen, T. & Palm, E. 1975 Stability of linear flow. Phys. Fluids 18 (4), 487488.Google Scholar
Farrell, B. 1988 Optimal excitation of perturbations in vicous shear flow. Phys. Fluids 5 (11), 26002609.Google Scholar
Farrell, B. F. & Ioannou, P. J. 1993 Optimal excitation of three-dimensional perturbations in viscous constant shear flow. Phys. Fluids A 612 (6), 13901400.CrossRefGoogle Scholar
Fontane, J.2005 Transition des écoulements cisaillés libres à densité variable. PhD thesis, Institut National Polytechnique de Toulouse.Google Scholar
Fontane, J. & Joly, L. 2008 The stability of the variable-density Kelvin–Helmholtz billow. J. Fluid Mech. 5, 237260.Google Scholar
Fontane, J., Joly, L. & Reinaud, J. N. 2008 Fractal Kelvin–Helmholtz break-ups. Phys. Fluids 20 (9), 091109.Google Scholar
Foures, D. P. G., Caulfield, C. P. & Schmid, P. J. 2014 Optimal mixing in two-dimensional plane poiseuille flow at infinite Péclet number. J. Fluid Mech. 748, 241277.Google Scholar
Gunzburger, M. 2003 Perspectives in Flow Control and Optimization. SIAM.Google Scholar
Joly, L. 2002 The structure of some variable density shear flows. In Variable Density Fluid Turbulence (ed. Chassaing, P., Antonia, R. A., Anselmet, F., Joly, L. & Sarkar, S.), chap. 8, pp. 201234. Springer.Google Scholar
Joly, L., Fontane, J. & Chassaing, P. 2005 The Rayleigh–Taylor instability of two-dimensional high-density vortices. J. Fluid Mech. 537, 415431.Google Scholar
Joly, L. & Reinaud, J. N. 2007 The merger of two-dimensional radially stratified high-Froude-number vortices. J. Fluid Mech. 582, 133151.Google Scholar
Joly, L., Reinaud, J. N. & Chassaing, P. 2001 The baroclinic forcing of the shear layer three-dimensional instability. In 2nd International Symposium on Turbulence and Shear Flow Phenomena, Stockholm, vol. 3, pp. 5964.Google Scholar
Joly, L., Suarez, J. & Chassaing, P. 2003 The strain field of light jets. In 5th European Fluid Mechanics Conference.Google Scholar
Joseph, D. D. 1990 Fluid dynamics of two miscible liquids with diffusion and gradient stresses. Eur. J. Mech. (B/Fluids) 9, 565596.Google Scholar
Kaminski, A. K., Caulfield, C. P. & Taylor, J. R. 2014 Transient growth in strongly stratified shear layers. J. Fluid Mech. 758, R4.CrossRefGoogle Scholar
Klaassen, G. P. & Peltier, W. R. 1985 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
Knio, O. M. & Ghoniem, A. F. 1992 The three-dimensional structure of periodic vorticity layers under non-symmetric conditions. J. Fluid Mech. 243, 353392.Google Scholar
Landhal, M. T. 1975 Wave breakdown and turbulence. SIAM J. Appl. Maths 28 (4), 735756.Google Scholar
Landhal, M. T. 1980 A note on an algebraic instability of inviscid parallel shear flows. J. Fluid Mech. 98 (2), 243251.Google Scholar
Lasheras, J. C. & Choi, H. 1988 Three-dimensional instability of a plane free shear layer: an experimental study of the formation and evolution of streamwise vortices. J. Fluid Mech. 189, 5386.Google Scholar
Lopez-Zazueta, A.2015. Stabilité secondaire non-modale d’une couche de mélange inhomogène. PhD thesis, Université de Toulouse, Institut Supérieur de l’Aéronautique et de l’Espace.Google Scholar
Luchini, P. & Bottaro, A. 1998 Görtler vortices: a backward-in-time approach to the receptivity problem. J. Fluid Mech. 363, 123.CrossRefGoogle Scholar
Mathew, G., Mezic, I. & Petzold, L. 2005 A multiscale measure for mixing. Physica D 211, 2346.Google Scholar
Metcalfe, R. W., Orszag, S. A., Brachet, M. E., Menon, S. & Riley, J. J. 1987 Secondary instability of a temporally growing mixing layer. J. Fluid Mech. 184, 207243.Google Scholar
Monkewitz, P. A., Bechert, D. W., Barsikow, B. & Lehman, B. 1990 Self-excited oscillations and mixing in a heated round jet. J. Fluid Mech. 213, 611639.CrossRefGoogle Scholar
Monkewitz, P. A., Lehmann, B., Barsikow, B. & Bechert, D. W. 1989 The spreading of self-excited hot jets by side jets. Phys. Fluids A 1 (3), 446448.Google Scholar
Monkewitz, P. A. & Pfizenmaier, E. 1991 Mixing by ‘side-jets’ in strongly forced and self-excited round jets. Phys. Fluids A 3 (5), 13561361.Google Scholar
Orr, W. M. 1907a The stability or instability of the steady motions of a perfect liquid and of a viscous liquid. Part 1: a perfect liquid. Proc. R. Irish Acad. 27, 968.Google Scholar
Orr, W. M. 1907b The stability or instability of the steady motions of a perfect liquid and of a viscous liquid. part 2: a viscous liquid. Proc. R. Irish Acad. 27, 69138.Google Scholar
Pierrehumbert, R. T. & Widnall, S. E. 1982 The two- and three-dimensional instabilities of a spatially periodic shear layer. J. Fluid Mech. 144, 5982.Google Scholar
Reinaud, J., Joly, L. & Chassaing, P. 2000 The baroclinic secondary instability of the two-dimensional shear layer. Phys. Fluids 12 (10), 24892505.Google Scholar
Rogers, M. & Moser, R. 1992 The three-dimensional evolution of a plane mixing layer: the Kelvin–Helmholtz rollup. J. Fluid Mech. 243, 183226.Google Scholar
Sandoval, D. L.1995 The dynamics of variable-density turbulence. PhD thesis, University of Washington.Google Scholar
Schmid, P. J. 2007 Nonmodal stability theory. Annu. Rev. Fluid Mech. 39, 129162.CrossRefGoogle Scholar
Schowalter, D., Van Atta, C. & Lasheras, J. 1994 A study of streamwise vortex structure in a stratified shear layer. J. Fluids Mech. 281, 247292.CrossRefGoogle Scholar
Soteriou, M. C. & Ghoniem, A. F. 1995 Effect of the free-stream density ratio on free and forced spatially developing shear layers. Phys. Fluids A 7 (8), 20362051.Google Scholar