Hostname: page-component-7c8c6479df-8mjnm Total loading time: 0 Render date: 2024-03-28T19:05:40.567Z Has data issue: false hasContentIssue false

Waves and turbulence on a beta-plane

Published online by Cambridge University Press:  29 March 2006

Peter B. Rhines
Affiliation:
Woods Hole Oceanographic Institution, Massachusetts 02543

Abstract

Two-dimensional eddies in a homogeneous fluid at large Reynolds number, if closely packed, are known to evolve towards larger scales. In the presence of a restoring force, the geophysical beta-effect, this cascade produces a field of waves without loss of energy, and the turbulent migration of the dominant scale nearly ceases at a wavenumber kβ = (β/2U)½ independent of the initial conditions other than U, the r.m.s. particle speed, and β, the northward gradient of the Coriolis frequency.

The conversion of turbulence into waves yields, in addition, more narrowly peaked wavenumber spectra and less fine-structure in the spatial maps, while smoothly distributing the energy about physical space.

The theory is discussed, using known integral constraints and similarity solutions, model equations, weak-interaction wave theory (which provides the terminus for the cascade) and other linearized instability theory. Computer experiments with both finite-difference and spectral codes are reported. The central quantity is the cascade rate, defined as \[ T = 2\int_0^{\infty} kF(k)dk/U^3\langle k\rangle , \] where F is the nonlinear transfer spectrum and 〈k〉 the mean wavenumber of the energy spectrum. (In unforced inviscid flow T is simply U−1dk−1/dt, or the rate at which the dominant scale expands in time t.) T is shown to have a mean value of 3·0 × 10−2 for pure two-dimensional turbulence, but this decreases by a factor of five at the transition to wave motion. We infer from weak-interaction theory even smaller values for k [Lt ] kβ.

After passing through a state of propagating waves, the homogeneous cascade tends towards a flow of alternating zonal jets which, we suggest, are almost perfectly steady. When the energy is intermittent in space, however, model equations show that the cascade is halted simply by the spreading of energy about space, and then the end state of a zonal flow is probably not achieved.

The geophysical application is that the cascade of pure turbulence to large scales is defeated by wave propagation, helping to explain why the energy-containing eddies in the ocean and atmosphere, though significantly nonlinear, fail to reach the size of their respective domains, and are much smaller. For typical ocean flows, $k_{\beta}^{-1} = 70\,{\rm km} $, while for the atmosphere, $k_{\beta}^{-1} = 1000\,{\rm km}$. In addition the cascade generates, by itself, zonal flow (or more generally, flow along geostrophic contours).

Type
Research Article
Copyright
© 1975 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. 1972 An alternate scale representation of atmospheric energy spectra J. Atmos. Sci. 29, 649.Google Scholar
Batchelor, G. K. 1953 Homogeneous Turbulence. Cambridge University Press.
Batchelor, G. K. 1969 Computation of the energy spectrum in homogeneous two-dimensional turbulence. Phys. Fluids Suppl. 12, II 233.Google Scholar
Crease, J. 1962 Velocity measurements in the deep water of the western north Atlantic J. Geophys. Res. 68, 3173.Google Scholar
Elliasen, E. 1958 A study of long atmospheric waves on the basis of zonal harmonic analysis Tellus, 10, 206.Google Scholar
Elliasen, E. & Machenauer, R. 1965 A study of the fluctuations of the atmospheric flow patterns Tellus, 17, 220.Google Scholar
Fofonoff, N. P. 1954 Steady flow in a frictionless, homogeneous ocean J. Mar. Res. 13, 257.Google Scholar
Fox, D. G. & Orszag, S. A. 1972 Inviscid dynamics of two-dimensional turbulence. Nat. Center Atmos. Res. MS. no. 72–80.Google Scholar
Gavrlin, B. L., Mirabel, A. P. & Monin, A. S. 1972 On the energy spectrum of synoptic processes. Isv. Atmos. Ocean. Phys. 5, 483 (in Russian).Google Scholar
Gill, A. E. 1974 The stability of planetary waves on an infinite beta-plane Geophys. Fluid Dyn. 5, 2947.Google Scholar
Hasselmann, K. 1967 A criterion for nonlinear wave stability J. Fluid Mech. 30, 737.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, 417.Google Scholar
Kao, S. K. & Wendell, L. 1970 The kinetic energy of the large-scale atmospheric motion in wavenumber—frequency space J. Atmos. Sci. 27, 359.Google Scholar
Kenyon, K. 1964 Nonlinear energy transfer in a Rossby-wave spectrum. Geophys. Fluid Dynamics Program, Woods Hole, Massachusetts.Google Scholar
Kenyon, K. 1967 Discussion. Proc. Roy. Soc. A 299, 141.Google Scholar
Kraichnan, R. H. 1967 Inertial ranges in two-dimensional turbulence Phys. Fluids, 10, 1417.Google Scholar
Kraichnan, R. H. 1971 Inertial-range transfer in two- and three-dimensional turbulence J. Fluid Mech. 47, 525.Google Scholar
Leith, C. E. 1972 Atmospheric predictability and two-dimensional turbulence J. Atmos. Sci. 28, 145.Google Scholar
Lilly, D. K. 1969 Numerical simulation of two-dimensional turbulence. High-speed computing in fluid dynamics. Phys. Fluids Suppl. 12, II 240.Google Scholar
Lilly, D. K. 1971 Numerical simulation of developing and decaying two-dimensional turbulence J. Fluid Mech. 45, 395.Google Scholar
Lilly, D. K. 1972 Numerical simulation studies of two-dimensional turbulence. I Geophys. Fluid Dyn. 3, 289.Google Scholar
LONGUET-HIGGINS, M. S. & Gill, A. E. 1967 Resonant interactions between planetary waves. Proc. Roy. Soc. A 299, 120.Google Scholar
Lorenz, E. N. 1972 Barotropic instability of Rossby wave motion J. Atmos. Sci. 29, 258.Google Scholar
Mcewan, A. D. 1972 Resonant degeneration of resonantly-excited standing internal gravity waves J. Fluid Mech. 50, 431.Google Scholar
Martin, S., Simmons, W. & Wunsch, C. 1972 The excitation of resonant triads by single internal waves J. Fluid Mech. 53, 17.Google Scholar
Newell, A. C. 1970 Rossby wave packet interactions J. Fluid. Mech. 35, 255.Google Scholar
Onsager, L. 1949 Statistical hydrodynamics Nuovo Cimento, 6, 279.Google Scholar
Orszag, S. A. 1971 Numerical simulation of incompressible flows with simple boundaries: accuracy. J. Fluid Mech. 49, 75.Google Scholar
Pedlosky, J. 1965 A note on the western intensification of the oceanic circulation J. Mar. Res. 23, 207.Google Scholar
Phillips, N. A. 1966 Large-scale eddy motion in the western Atlantic J. Geophys. Res. 71, 3883.Google Scholar
Phillips, O. M. 1966 The Dynamics of the Upper Ocean, p. 109. Cambridge University Press.
Rhines, P. B. 1973 Observations of the energy-containing oceanic eddies, and theoretical models of waves and turbulence Boundary Layer Met. 4, 345.Google Scholar
Rhines, P. B. & Bretherton, F. B. 1973 Topographic Rossby waves in a rough-bottomed ocean J. Fluid Mech. 61, 583.Google Scholar
Schmitz, W. J. 1974 Observations of low-frequency current fluctuations on the continental slope and rise near Site D J. Mar. Res. 32, 233.Google Scholar
Simmons, W. F. 1969 A variational method for weak resonant wave interactions. Proc. Roy. Soc. A 309, 551.Google Scholar
Starr, V. P. 1968 Physics of Negative Viscosity Phenomena. McGraw-Hill.
Taylor, G. I. 1917 Observations and speculations on the nature of turbulent motion. Scientific Papers, vol. 2 (ed. G. K. Batchelor), p. 69. Cambridge University Press.