Hostname: page-component-76fb5796d-x4r87 Total loading time: 0 Render date: 2024-04-26T00:42:53.696Z Has data issue: false hasContentIssue false

Experiments on the structure and dynamics of forced, quasi-two-dimensional turbulence

Published online by Cambridge University Press:  26 April 2006

S. Narimousa
Affiliation:
University of Southern California, Department of Mechanical Engineering, Los Angeles, CA 90089-1453, USA
T. Maxworthy
Affiliation:
University of Southern California, Departments of Aerospace and Mechanical Engineering, Los Angeles, CA 90089-1191, USA
G. R. Spedding
Affiliation:
University of Southern California, Department of Mechanical Engineering, Los Angeles, CA 90089-1453, USA

Abstract

Simulated upwelling fronts have been generated around the outer edge of a cylindrical tank filled with a two-layer fluid system and driven by a surface stress. Initially, an axisymmetric front was observed which subsequently became unstable to small baroclinic eddies. These eddies continued to grow until they reached an equilibrium size. Under some circumstances, cyclonic eddies pinched-off from the fully developed front and moved away from the mean position of the front into the fluid interior. Streak photographs of the fully developed flow field were digitized to generate a velocity field interpolated on to a regular grid. A direct two-dimensional Fourier transform was performed on the turbulent kinetic energy field deduced from such images and one-dimensional energy E(k) spectra were extracted. Consistent $k^{-\frac{5}{3}}$ energy spectra were found at lower wavenumber, k and approximately k−5.5 spectra at higher k. In any given experiment, the two spectral slopes meet close to a wavenumber kw = 2π/λw (where λw is the mean diameter of a frontal eddy and kw is the associated wavenumber). According to classical theories, kw would be the input wavenumber, and the range of k with a $k^{-\frac{5}{3}}$ spectrum would correspond to an inverse energy cascade range; this yielded a Kolmogorov constant (C) that varied within the limits 2.8 [les ] C [les ] 3.8. The approximately k−5.5 range, which is much steeper than that predicted by the original statistical theories, is nevertheless consistent with those found frequently in numerical experiments.

The spectral slopes inferred from particle dispersion methods and from one-dimensional Fourier transforms of the longitudinal velocity correlations were compared with the results obtained above and in previous laboratory experiments. In general, the global energy spectra are consistent with an interpretation of the fluid dynamics as being that of two-dimensional turbulence. This in turn implies that known properties of such flows may be invoked to explain the appearance of a number of naturally occurring phenomena in coastal upwelling fronts.

Type
Research Article
Copyright
© 1991 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

Agüí, J. C. & Jimenez, J. 1987 On the performance of particle tracking. J. Fluid Mech. 185, 447468.Google Scholar
Armi, L. & Flament, P., 1985 Cautionary remarks on the spectral interpretation of turbulent flows. J. Geophys. Res. 90, 1177911782.Google Scholar
Babiako, A., Basdevant, C. & Sadourny, R., 1985 Structure functions and dispersion laws in two-dimensional turbulence. J. Atmos. Sci. 42, 941949.Google Scholar
Basdevant, C., Legras, B., Sadourny, R. & Beland, M., 1981 A study of barotropic model flows: intermittency waves and predictability. J. Atmos. Sci. 38, 23052326.Google Scholar
Batchelor, G. K.: 1953 The Theory of Homogeneous Turbulence. Cambridge University Press.
Batchelor, G. K.: 1969 Computation of energy spectrum in homogeneous two-dimensional turbulence. Phys. Fluids Suppl. 12, II 233.Google Scholar
Bennett, A. F.: 1984 Relative dispersions: local and nonlocal dynamics. J. Atmos. Sci. 41, 18811886.Google Scholar
Bennett, A. F. & Haidvogel, D. B., 1983 Low-resolution numerical simulation of decaying two-dimensional turbulence. J. Atmos. Sci. 40, 738748.Google Scholar
Bracket, M. E., Meneguzzi, M., Politano, H. & Sulem, P. L., 1988 The dynamics of freely decaying two-dimensional turbulence. J. Fluid Mech. 194, 333349.Google Scholar
Bracket, M. E., Meneguzzi, M. & Sulem, P. L., 1986 Small-scale dynamics of the high Reynolds number two-dimensional turbulence. Phys. Rev. Lett. 57, 683686.Google Scholar
Brown, P. S. & Robinson, G. D., 1979 The variance spectrum of tropospheric winds over Eastern Europe. J. Atmos. Sci 36, 270286.Google Scholar
Charney, J. G.: 1971 Geostrophic turbulence. J. Atmos. Sci. 28, 10871095.Google Scholar
Chia, F. R., Griffiths, R. W. & Linden, P. F., 1982 Laboratory experiments on fronts. Part II. The formulation of cyclonic eddies at upwelling fronts. Geophys. Astrophys. Fluid Dyn. 19, 189206.Google Scholar
Deem, G. S. & Zabusky, N. J., 1971 Ergodic boundary in numerical simulations of two-dimensional turbulence. Phys. Rev. Lett. 27, 396399.Google Scholar
Dubois, M.: 1975 Large-scale kinetic energy spectra from Eulerian analysis of Eole wind data. J. Atmos. Sci. 32, 18381847.Google Scholar
Fornberg, B.: 1977 A numerical study of two-dimensional turbulence. J. Comput. Phys. 15, 131.Google Scholar
Fox, D. G. & Orszag, S. A., 1972 Inviscid dynamics of two-dimensional turbulence. Natl Center Atmos. Res. MS. 72–80.Google Scholar
Frisch, U. & Sulem, P. L., 1984 Numerical simulation of inverse cascade in two-dimensional turbulence. Phys. Fluids 27, 19211923.Google Scholar
Gage, K. S.: 1979 Evidence for a K–5/3 law inertial range in a mesoscale two-dimensional turbulence. J. Atmos. Sci. 36, 19501954.Google Scholar
Griffiths, R. W. & Hopfinger, E. J., 1984 The structure of mesoscale turbulence and horizontal spreading at ocean fronts. Deep-Sea Res. 31, 245269.Google Scholar
Griffiths, R. W. & Linden, P. F., 1981 The stability of vortices in a rotating stratified fluid. J. Fluid Mech. 195, 283316.Google Scholar
Griffiths, R. W. & Linden, P. F., 1982 Laboratory experiments on fronts. Part I: Density driven boundary currents. Geophys. Astrophys. Fluid Dyn. 19, 159187.Google Scholar
Haidvogel, D. B. & Keffer, T., 1984 Tracer dispersal by midocean mesoscale eddies. I: Ensemble statistics. Dyn. Atmos. Oceans 8, 140.Google Scholar
Herring, J. R.: 1980 Statistical theory of quasi-geostrophic turbulence. J. Atmos. Sci. 37, 969977.Google Scholar
Herring, J. R. & Mcwilliams, J. C., 1985 Comparison of direct numerical simulation of two-dimensional turbulence with two-joint closure: the effects of intermittency. J. Fluid Mech. 153, 229242.Google Scholar
Herring, J. R., Orszag, S. A., Kraichnan, R. H. & Fox, D. G., 1974 Decay of two-dimensional homogeneous turbulence. J. Fluid Mech. 66, 417444.Google Scholar
Heyer, J. M. & Sadourny, R., 1982 Closure modeling of fully developed baroclinic instabilities. J. Atmos. Sci. 39, 707721.Google Scholar
Imaichi, K. & Ohmi, K., 1983 Numerical processing of flow-visualization pictures - measurement of two-dimensional vortex flow. J. Fluid Mech. 129, 283311.Google Scholar
Julian, P. R., Washington, W. M., Hembre, L. & Ridley, C., 1970 On the spectral distribution of large-scale atmospheric kinetic energy. J. Atmos. Sci. 27, 376387.Google Scholar
Kao, S. K., Jenne, R. L. & Sagendolf, J. F., 1970 The kinetic energy of large-scale atmospheric motion in wavenumber-frequency space: II. Mid-Troposphere of the southern Hemisphere. J. Atmos. Sci. 27, 10081020.Google Scholar
Killworth, P. D., Paldor, N. & Stern, M., 1984 Wave propagation and growth on a surface front in a two-layer geostrophic current. J. Mar. Res. 42, 761785.Google Scholar
Kolmogorov, A. N.: 1941 Dispersion of energy in locally isotropic turbulence. CR Acad. Sci. URSS 32, 16.Google Scholar
Kraichnan, R. H.: 1967 Inertial ranges in two-dimensional turbulence. Phys. Fluids 10, 14171423.Google Scholar
Kraichnan, R. H.: 1975 Statistical dynamics of two-dimensional flow. J. Fluid Mech. 67, 155.Google Scholar
Leith, C. E.: 1968 Diffusion approximation for two-dimensional turbulence. Phys. Fluids 11, 671673.Google Scholar
Lesieiur, M.: 1987 Turbulence in Fluids. Martinus Nijhoff.
Lilly, D. K.: 1969 Numerical simulation of two-dimensional turbulence. High-speed computing in fluid dynamics. Phys. Fluid Suppl. 12, II 240.Google Scholar
Lilly, D. K.: 1971 Numerical simulation of developing and decaying two-dimensional turbulence. J. Fluid Mech. 45, 395415.Google Scholar
Lilly, D. K.: 1972 Numerical simulation studies of two-dimensional turbulence. I. Geophys. Fluid. Dyn. 3, 289.Google Scholar
Maxworthy, T.: 1989 The dynamics of two-dimensional turbulence. In Proc. Conf. on the Oceanography of Sea Straits, les Arcs, France.Google Scholar
Maxworthy, T., Caperan, P. & Spedding, G. B., 1987 Two-dimensional turbulence and vortex dynamics in a stratified fluid. In Proc. Third Intl. Conf. on Stratified Flows. Pasadena, California.Google Scholar
Mcwilliams, J. C.: 1984 The emergence of isolated coherent vortices in turbulent flow. J. Fluid Mech. 146, 2143.Google Scholar
Morel, P. & Larcheveque, M., 1974 Relative dispersion for constant-level balloons in te 200 mb general circulation. J. Atmos. Sci. 31, 21892196.Google Scholar
Moby, M. & Hopfinger, E. J., 1986 Structure functions in a rotationally dominated turbulent flow. Phys. Fluids 29, 21402146.Google Scholar
Narimousa, S. & Maxworthy, T., 1985 Two-layer model of shear driven coastal upwelling in the presence of bottom topography. J. Fluid Mech. 159, 503531.Google Scholar
Narimousa, S. & Maxworthy, T., 1986 Effects of a discontinuous surface stress on a model of coastal upwelling. J. Phys. Oceanogr. 16, 20712083.Google Scholar
Narimousa, S. & Maxworthy, T., 1987a Coastal upwelling on a sloping bottom: The formation of plumes, jets and pinched-off cyclones. J. Fluid Mech. 176, 169190.Google Scholar
Narimousa, S. & Maxworthy, T., 1987b Effects of coastline perturbations on coastal currents and fronts. J. Phys. Oceanogr. 17, 12961303.Google Scholar
Narimousa, S. & Maxworthy, T., 1989 Application of a laboratory model to the interpretation of satellite and field observations of coastal upwelling. Dyn. Atmos. Oceans 13, 146.Google Scholar
Narimousa, S., Maxworthy, T. & Spedding, G. R., 1987 Structure of meso-scale baroclinic turbulence. Sixth Symp. on Turbulent Shear Flows. Toulouse, France, September 7–9.Google Scholar
Nastrom, G. D. & Gage, K. S., 1983 A first look at wavenumber spectra from GASP data. Tellus 35, 383388.Google Scholar
Phillips, N. A.: 1954 Energy transformations and meridional circulations associated with simple baroclinic waves in a two-level, quasi-geostropic model. Tellus 6, 273286.Google Scholar
Rignot, E. J. M. & Spedding, G. R. 1988 Performance analysis of automated image processing and grid interpolation techniques for fluid flows. University of Southern California, Department of Aerospace Engineering Internal Rep. USCAE 143.Google Scholar
Saffman, P. G.: 1971 On the spectrum and decay of random two-dimensional vorticity distributions at large Reynolds numbers. Stud. Appl. Maths 50, 377383.Google Scholar
Salmon, R.: 1978 Two-layer quasi-geostrophic turbulence in a simple case. Geophys. Astrophys. Fluid Dyn. 10, 2551.Google Scholar
Santangelo, P., Benzi, R. & Legras, B., 1989 The generation of vortices in high-resolution, two-dimensional decaying turbulence and the influence of initial conditions on the breaking of self-similarity. Phys. Fluids A Al, 10271034.Google Scholar
Siggia, E. D. & Aref, H., 1981 Point vortex simulation of the inverse cascade in two-dimensional turbulence. Phys. Fluids 24, 171173.Google Scholar
Sommeria, J.: 1986 Experimental study of the two-dimensional inverse energy cascade in a square box. J. Fluid Mech. 170, 139168.Google Scholar
Townsend, A. A.: 1976 Structure of turbulent Shear Flow. Cambridge University Press.