Hostname: page-component-7479d7b7d-m9pkr Total loading time: 0 Render date: 2024-07-12T02:17:53.399Z Has data issue: false hasContentIssue false

On Lagrangian and vortex-surface fields for flows with Taylor–Green and Kida–Pelz initial conditions

Published online by Cambridge University Press:  01 October 2010

YUE YANG*
Affiliation:
Graduate Aerospace Laboratories, 205-45, California Institute of Technology, Pasadena, CA 91125, USA
D. I. PULLIN
Affiliation:
Graduate Aerospace Laboratories, 205-45, California Institute of Technology, Pasadena, CA 91125, USA
*
Email address for correspondence: yy@caltech.edu

Abstract

For a strictly inviscid barotropic flow with conservative body forces, the Helmholtz vorticity theorem shows that material or Lagrangian surfaces which are vortex surfaces at time t = 0 remain so for t > 0. In this study, a systematic methodology is developed for constructing smooth scalar fields φ(x, y, z, t = 0) for Taylor–Green and Kida–Pelz velocity fields, which, at t = 0, satisfy ω·∇φ = 0. We refer to such fields as vortex-surface fields. Then, for some constant C, iso-surfaces φ = C define vortex surfaces. It is shown that, given the vorticity, our definition of a vortex-surface field admits non-uniqueness, and this is presently resolved numerically using an optimization approach. Additionally, relations between vortex-surface fields and the classical Clebsch representation are discussed for flows with zero helicity. Equations describing the evolution of vortex-surface fields are then obtained for both inviscid and viscous incompressible flows. Both uniqueness and the distinction separating the evolution of vortex-surface fields and Lagrangian fields are discussed. By tracking φ as a Lagrangian field in slightly viscous flows, we show that the well-defined evolution of Lagrangian surfaces that are initially vortex surfaces can be a good approximation to vortex surfaces at later times prior to vortex reconnection. In the evolution of such Lagrangian fields, we observe that initially blob-like vortex surfaces are progressively stretched to sheet-like shapes so that neighbouring portions approach each other, with subsequent rolling up of structures near the interface, which reveals more information on dynamics than the iso-surfaces of vorticity magnitude. The non-local geometry in the evolution is quantified by two differential geometry properties. Rolled-up local shapes are found in the Lagrangian structures that were initially vortex surfaces close to the time of vortex reconnection. It is hypothesized that this is related to the formation of the very high vorticity regions.

Type
Papers
Copyright
Copyright © Cambridge University Press 2010

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

Anderson, E., Bai, Z., Bischof, C., Blackford, S., Demmel, J., Dongarra, J., Du Croz, J., Greenbaum, A., Hammarling, S., McKenney, A. & Sorensen, D. 1999 LAPACK Users' Guide, 3rd edn. Society for Industrial and Applied Mathematics.Google Scholar
Ashurst, W. T., Kerstein, A. R., Kerr, R. M. & Gibson, C. H. 1987 Alignment of vorticity and scalar gradient with strain rate in simulated Navier–Stokes turbulence. Phys. Fluids 30, 23432353.Google Scholar
Bermejo-Moreno, I. & Pullin, D. I. 2008 On the non-local geometry of turbulence. J. Fluid Mech. 603, 101135.CrossRefGoogle Scholar
Boratav, O. N. & Pelz, R. B. 1994 Direct numerical simulation of transition to turbulence from a high-symmetry initial condition. Phys. Fluids 6, 27572784.Google Scholar
Brachet, M. E., Meiron, D. I., Orszag, S. A., Nickel, B. G., Morf, R. H. & Frisch, U. 1983 Small-scale structure of the Taylor–Green vortex. J. Fluid Mech. 130, 411452.Google Scholar
Brachet, M. E., Meneguzzi, M., Vincent, A., Politano, H. & Sulem, P. L. 1992 Numerical evidence of smooth self-similar dynamics and possibility of subsequent collapse for three-dimensional ideal flows. Phys. Fluids A 4, 28452854.CrossRefGoogle Scholar
Branicki, M. & Wiggins, S. 2009 An adaptive method for computing invariant manifolds in non-autonomous, three-dimensional dynamical systems. Physica D 238, 16251657.Google Scholar
Cartes, C., Bustamante, M. D. & Brachet, M. E. 2007 Generalized Eulerian–Lagrangian description of Navier–Stokes dynamics. Phys. Fluids 19, 077101.CrossRefGoogle Scholar
Chakraborty, P., Balachandar, S. & Adrian, R. J. 2005 On the relationships between local vortex identification schemes. J. Fluid Mech. 535, 189214.CrossRefGoogle Scholar
Chong, M. S., Perry, A. E. & Cantwell, B. J. 1990 A general classification of three-dimensional flow fields. Phys. Fluids A 2, 765777.CrossRefGoogle Scholar
Clebsch, A. 1859 Ueber die Integration der hydrodynamischen Gleichungen. J. Reine Angew. Math. 56, 110.Google Scholar
Constantin, P. 2001 An Eulerian–Lagrangian approach to the Navier–Stokes equations. Commun. Math. Phys. 216, 663686.CrossRefGoogle Scholar
Constantin, P., Majda, A. J. & Tabak, E. 1994 Formation of strong fronts in the 2-D quasi-geostrophic thermal active scalar. Nonlinearity 7, 14951533.Google Scholar
Dieci, L., Lorenz, J. & Russell, R. D. 1991 Numerical calculation of invariant tori. SIAM J. Sci. Stat. Comput. 12, 607647.CrossRefGoogle Scholar
Dombre, T., Frisch, U., Greene, J. M., Henon, M., Mehr, A. & Soward, A. M. 1986 Chaotic streamlines in the ABC flows. J. Fluid Mech. 167, 353391.CrossRefGoogle Scholar
Ertel, H. 1942 Ein neuer hydrodynamischer Wirbelsatz. Meteorol Z. 59, 271281.Google Scholar
Golub, G. H. & Van Loan, C. F. 1996 Matrix Computations, 3rd edn. Johns Hopkins University Press.Google Scholar
Goto, S. & Kida, S. 2007 Reynolds-number dependence of line and surface stretching in turbulence: folding effects. J. Fluid Mech. 586, 5981.CrossRefGoogle Scholar
Haller, G. 2005 An objective definition of a vortex. J. Fluid Mech. 525, 126.CrossRefGoogle Scholar
Helmholtz, H. 1858 Über Integrale der hydrodynamischen Gleichungen welche den Wirbel-bewegungen ensprechen. J. Reine Angew. Math. 55, 2555.Google Scholar
Hou, T. Y. & Li, R. 2008 Blowup or no blowup? The interplay between theory and numerics. Physica D 237, 19371944.Google Scholar
Hunt, J. C. R., Wray, A. A. & Moin, P. 1988 Eddies, stream, and convergence zones in turbulent flows. Center for Turbulence Research Rep. CTR-S88, pp. 193208.Google Scholar
Jeong, J. & Hussain, F. 1995 On the identification of a vortex. J. Fluid Mech. 285, 6994.Google Scholar
Jiang, G. S. & Shu, C. W. 1996 Efficient implementation of weighted ENO schemes. J. Comput. Phys. 126, 202228.Google Scholar
Kerr, R. M. 1993 Evidence for a singularity of the three-dimensional, incompressible Euler equations. Phys. Fluids A 5, 17251746.Google Scholar
Kida, S. 1985 Three-dimensional periodic flows with high-symmetry. J. Phys. Soc. Japan 54, 21322136.CrossRefGoogle Scholar
Kida, S. & Takaoka, M. 1994 Vortex reconnection. Annu. Rev. Fluid Mech. 26, 169189.CrossRefGoogle Scholar
Koenderink, J. J. & VanDoorn, A. J. Doorn, A. J. 1992 Surface shape and curvature scales. Image Vision Comput. 10, 557564.CrossRefGoogle Scholar
Krauskopf, B., Osinga, H. M., Doedel, E. J., Henderson, M. E., Guckenheimer, J., Vladimirsky, A., Dellnitz, M. & Junge, O. 2005 A survey of methods for computing (un)stable manifold of vector fields. Intl J. Bifurcation Chaos 15, 763791.CrossRefGoogle Scholar
Lamb, H. 1932 Hydrodynamics, 6th edn. Cambridge University Press.Google Scholar
Lim, T. T. & Nickels, T. B. 1993 Instability and reconnection in the head-on collision of two vortex rings. Nature 357, 225227.CrossRefGoogle Scholar
Lindsay, K. & Krasny, R. 2001 A particle method and adaptive treecode for vortex sheet motion in three-dimensional flow. J. Comput. Phys. 172, 879907.Google Scholar
Lu, L. & Doering, C. R. 2008 Limits on enstrophy growth for solutions of the three-dimensional Navier–Stokes equations. Indiana Univ. Math. J. 57, 26932727.CrossRefGoogle Scholar
Majda, A. J. & Bertozzi, A. L. 2001 Vorticity and Incompressible Flow. Cambridge University Press.CrossRefGoogle Scholar
Mingyu, H., Küpper, T. & Masbaum, N. 1997 Computation of invariant tori by the Fourier methods. SIAM J. Sci. Comput. 18, 918942.CrossRefGoogle Scholar
Nore, C., Abid, M. & Brachet, M. E. 1997 Decaying Kolmogorov turbulence in a model of superflow. Phys. Fluids 9, 26442669.CrossRefGoogle Scholar
Ohkitani, K. 2008 A geometrical study of 3D incompressible Euler flows with Clebsch potentials: a long-lived Euler flow and its power-law energy spectrum. Physica D 237, 20202027.Google Scholar
Ohkitani, K. & Constantin, P. 2003 Numerical study of the Eulerian–Lagrangian formulation of the Navier–Stokes equations. Phys. Fluids 15, 32513254.Google Scholar
Pelz, R. B. 2001 Symmetry and the hydrodynamic blow-up problem. J. Fluid Mech. 444, 299320.Google Scholar
Pullin, D. I. & Saffman, P. G. 1998 Vortex dynamics in turbulence. Annu. Rev. Fluid Mech. 30, 3151.CrossRefGoogle Scholar
Pumir, A., Shraiman, B. I. & Siggia, E. D. 1992 Vortex morphology and Kelvin's theorem. Phys. Rev. A 45, R5351R5354.Google Scholar
Pumir, A. & Siggia, E. 1990 Collapsing solutions to the 3-D Euler equations. Phys. Fluids A 2, 220241.Google Scholar
Ruetsch, G. R. & Maxey, M. R. 1992 The evolution of small-scale structures in homogeneous isotropic turbulence. Phys. Fluids A 4, 27472760.Google Scholar
Russo, G. & Smereka, P. 1999 Impulse formulation of the Euler equations: general properties and numerical methods. J. Fluid Mech. 391, 189209.Google Scholar
Saffman, P. G. 1992 Vortex Dynamics. Cambridge University Press.Google Scholar
Salmon, R. 1988 Hamiltonian fluid mechanics. Annu. Rev. Fluid Mech. 20, 225256.CrossRefGoogle Scholar
Taylor, G. I. & Green, A. E. 1937 Mechanism of the production of small eddies from large ones. Proc. R. Soc. Lond. A 158, 499521.Google Scholar
Wolfram Research, Inc. 2008 Mathematica. v7.0.Google Scholar
Yang, Y., Pullin, D. I. & Bermejo-Moreno, I. 2010 Multi-scale geometric analysis of Lagrangian structures in isotropic turbulence. J. Fluid Mech. 654, 233270.Google Scholar