Hostname: page-component-848d4c4894-wzw2p Total loading time: 0 Render date: 2024-05-30T06:54:45.041Z Has data issue: false hasContentIssue false

Submesoscale surface fronts and filaments: secondary circulation, buoyancy flux, and frontogenesis

Published online by Cambridge University Press:  20 June 2017

James C. McWilliams*
Affiliation:
Department of Atmospheric and Oceanic Sciences, University of California, Los Angeles, CA 90095-1565, USA
*
Email address for correspondence: jcm@atmos.ucla.edu

Abstract

Problems are posed and solved for upper-ocean submesoscale density fronts and filaments in the presence of surface wind stress and the associated boundary-layer turbulent mixing, their associated geostrophic and secondary circulations and their instantaneous buoyancy fluxes and frontogenetic evolutionary tendencies in both velocity and buoyancy gradients. The analysis is diagnostic rather than prognostic, and it is based on a momentum-balanced approximation that assumes the ageostrophic acceleration is negligible, although the Rossby number is finite and ageostrophic advection is included, justified by the quasi-steady, coherent-structure flow configurations of fronts and filaments. Across a wide range of wind and buoyancy-gradient parameters, the ageostrophic secondary circulation for a front is a single overturning cell with downwelling on the dense side, hence with a positive (restratifying) vertical buoyancy flux. For a dense filament the circulation is a double cell with central downwelling and again positive vertical buoyancy flux. The primary explanation for these secondary-circulation cells is a ‘turbulent thermal wind’ linear momentum balance. These circulation patterns, and their associated frontogenetic tendencies in both the velocity and buoyancy gradients, are qualitatively similar to those due to the ‘classical’ mechanism of strain-induced frontogenesis. For linear solutions, the secondary circulation and frontogenesis are essentially independent of wind direction, but in nonlinear solutions ageostrophic advection provides a strong intensification of the peak vertical velocity, while generally preserving the ageostrophic circulation pattern, when the Rossby number is order one and the wind orientation relative to the frontal axis is favourable. At large Rossby number the solution procedure fails to converge, with an implication of a failure of existence of wholly balanced circulations.

Type
Papers
Copyright
© 2017 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

Bachman, S. D. & Taylor, J. R. 2016 Numerical simulations of the equilibrium between eddy-induced restratification and vertical mixing. J. Phys. Oceanogr. 46, 919935.Google Scholar
Bergeron, T. 1928 Über die dreidimensional verknüpfende Wetteranalyse I. Geophys. Publ. 5, 1111.Google Scholar
Craik, A. D. D. & Leibovich, S. 1976 A rational model for Langmuir circulations. J. Fluid Mech. 73, 401426.Google Scholar
Cronin, M. F. & Kessler, W. S. 2009 Near-surface shear flow in the Tropical Pacific cold tongue front. J. Phys. Oceanogr. 39, 12001215.Google Scholar
Eliassen, A. 1962 On the vertical circulation in frontal zones. Geophys. Publ. 24, 147160.Google Scholar
Flierl, G. & Mied, R. 1985 Frictionally induced circulations and spin down of a warm-core ring. J. Geophys. Res. Oceans 90, 89178927.CrossRefGoogle Scholar
Fox-Kemper, B., Ferrari, R. & Hallberg, R. W. 2008 Parameterization of mixed layer eddies. Part I: theory and diagnosis. J. Phys. Oceanogr. 38, 11451165.CrossRefGoogle Scholar
Garrett, C. J. R. & Loder, J. W. 1981 Dynamical aspects of shallow sea fronts. Phil. Trans. R. Soc. Lond. 302, 563581.Google Scholar
Gula, J., Molemaker, M. J. & McWilliams, J. C. 2014 Submesoscale cold filaments in the Gulf Stream. J. Phys. Oceanogr. 44, 26172643.Google Scholar
Hoskins, B. J. 1982 The mathematical theory of frontogenesis. Annu. Rev. Fluid Mech. 14, 131151.Google Scholar
Hoskins, B. J. & Bretherton, F. P. 1972 Atmospheric frontogenesis models: mathematical formulation and solution. J. Atmos. Sci. 29, 1137.Google Scholar
Large, W. G., McWilliams, J. C. & Doney, S. C. 1994 Oceanic vertical mixing: a review and a model with a nonlocal boundary layer parameterization. Rev. Geophys. 32, 363403.Google Scholar
Mahadevan, A. & Tandon, A. 2006 An analysis of mechanisms for submesoscale vertical motion at ocean fronts. Ocean Model. 14, 241256.Google Scholar
McWilliams, J. C. 2003 Diagnostic force balance and its limits. In Nonlinear Processes in Geophysical Fluid Dynamics (ed. Velasco Fuentes, O. U., Sheinbaum, J. & Ochoa, J.), pp. 287304. Kluwer Academic Publishers.CrossRefGoogle Scholar
McWilliams, J. C. 2016 Submesoscale currents in the ocean. Proc. R. Soc. Lond. A 472, 20160117,1–32.Google Scholar
McWilliams, J. C., Colas, F. & Molemaker, M. J. 2009 Cold filamentary intensification and oceanic surface convergence lines. Geophys. Res. Lett. 36, L18602.Google Scholar
McWilliams, J. C., Gula, J., Molemaker, M. J., Renault, L. & Shchepetkin, A. F. 2015 Filament frontogenesis by boundary layer turbulence. J. Phys. Oceanogr. 45, 19882005.Google Scholar
McWilliams, J. C., Huckle, E. & Shchepetkin, A. F. 2009a Buoyancy effects in a stratified Ekman layer. J. Phys. Oceanogr. 39, 25812599.Google Scholar
McWilliams, J. C., Molemaker, M. J. & Olafsdottir, E. I. 2009b Linear fluctuation growth during frontogenesis. J. Phys. Oceanogr. 39, 31113129.Google Scholar
McWilliams, J. C., Restrepo, J. M. & Lane, E. M. 2004 An asymptotic theory for the interaction of waves and currents in coastal waters. J. Fluid Mech. 511, 135178.Google Scholar
McWilliams, J. C., Yavneh, I., Cullen, M. J. P. & Gent, P. R. 1998 The breakdown of large-scale flows in rotating, stratified fluids. Phys. Fluids 10, 31783184.Google Scholar
Nagai, T., Tandon, A. & Rudnick, D. L. 2006 Two-dimensional ageostrophic secondary circulation at ocean fronts due to vertical mixing and large-scale deformation. J. Geophys. Res. 111, C09038.Google Scholar
Niiler, P. P. 1969 On the Ekman divergence in an oceanic jet. J. Geophys. Res. 74, 70487052.CrossRefGoogle Scholar
Ponte, A. L., Klein, P., Capet, X., Le Traon, P.-Y., Chapron, B. & Lherminier, P. 2013 Dagnosing surface mixed layer dynamics from high-resolution satellite observations: numerical insights. J. Phys. Oceanogr. 43, 13451355.Google Scholar
Stern, M. E. 1965 Interaction of a uniform wind stress with a geostrophic vortex. Deep-Sea Res. 12, 355367.Google Scholar
Sullivan, P. P. & McWilliams, J. C. 2016 Frontogenesis and frontal arrest for a dense filament in the surface boundary layer. J. Fluid Mech.; (submitted).Google Scholar
Suzuki, N., Fox-Kemper, B., Hamlington, P. E. & Van Roekel, L. P. 2016 Surface waves affect frontogenesis. J. Geophys. Res. - Oceans 121, 128.CrossRefGoogle Scholar
Taylor, J. R. & Ferrari, R. 2009 On the equilibration of a symmetrically unstable front via a secondary shear instability. J. Fluid Mech. 622, 103113.Google Scholar
Thomas, L. N. 2005 Destruction of potential vorticity by winds. J. Phys. Oceanogr. 35, 24572466.CrossRefGoogle Scholar
Thompson, L. 2000 Ekman laers and two-dimensional frontogenesis in the upper ocean. J. Geophys. Res. 105, 64376451.Google Scholar
Wenegrat, J. O. & McPhadden, M. J. 2016a A simple analytical model of the diurnal Ekman layer. J. Phys. Oceanogr. 46, 28772894.CrossRefGoogle Scholar
Wenegrat, J. O. & McPhadden, M. J. 2016b Wind, waves, and fronts: frictional effects in a generalized Ekman model. J. Phys. Oceanogr. 46, 711747.Google Scholar