Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-ndmmz Total loading time: 0 Render date: 2024-06-06T16:25:26.576Z Has data issue: false hasContentIssue false

Bibliography

Published online by Cambridge University Press:  10 March 2021

Michael C. Gregg
Affiliation:
University of Washington
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Ocean Mixing , pp. 336 - 364
Publisher: Cambridge University Press
Print publication year: 2021

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

Agee, E.M. 1987. Mesoscale cellular convection over the oceans. Dyn. Atmos. Oceans, 10, 317341.CrossRefGoogle Scholar
Agrawal, Y.C., and Belting, C.J. 1988. Laser velocimetry for benthic sediment transport. Deep-Sea Res., 35, 10471067.Google Scholar
Alford, M.H. 2008. Observations of parametric subharmonic instability of the diurnal internal tide in the South China Sea. Geophys. Res. Lett., 35(L15602).Google Scholar
Alford, M.H. 2010. Sustained, full-water-column observations of internal waves and mixing near Mendocino Escarpment. J. Phys. Oceanogr., 40, 26432660.Google Scholar
Alford, M.H., Gerdt, D.W., and Adkins, C.M. 2006a. An ocean refractometer: Resolving millimeter-scale turbulent density fluctuations via the refractive index. J. Atmos. Ocean. Tech., 23, 121137.Google Scholar
Alford, M.H., Gregg, M.C., and D’Asaro, E.A. 2005. Mixing, 3-D mapping and Lagrangian evolution of a thermohaline intrusion. J. Phys. Oceanogr., 35, 16891711.CrossRefGoogle Scholar
Alford, M.H., Gregg, M.C., and Merrifield, M.A. 2006b. Structure, propagation and mixing of energetic baroclinic tides in Mamala Bay, Oahu, Hawaii. J. Phys. Oceanogr., 36, 9971018.Google Scholar
Alford, M.H., Gregg, M.C., Zervakis, V., and Kontoyiannis, H. 2012. Internal wave measurements on the Cycladic Plateau of the Aegean Sea. J. Geophys. Res., 117(C01015).Google Scholar
Alford, M.H., MacKinnon, J.A., Pinkel, R., and Klymak, J. 2017. Space-time scales of shear in the North Pacific. J. Phys. Oceanogr., 47, 24552478.Google Scholar
Alford, M.H., MacKinnon, J.A., Simmons, H.L., and Nash, J.D. 2016. Near-inertial internal gravity waves in the ocean. Annu. Rev. Mar. Sci., 8, 95123.Google Scholar
Alford, M.H., MacKinnon, J.A., Zhao, Z., Pinkel, R., Klymak, J., and Peacock, T. 2007. Internal waves across the Pacific. Geophys. Res. Lett., 34(L24601).Google Scholar
Alford, M.H., and Pinkel, R. 2000. Patterns of turbulent and double-diffusive phenomena: Rapid-profiling microconductivity probe. J. Phys. Oceanogr., 30, 833854.2.0.CO;2>CrossRefGoogle Scholar
Alford, M.H., Shcherbina, A.Y., and Gregg, M.C. 2013. Observations of near-inertial internal gravity waves radiating from a frontal jet. J. Phys. Oceanogr., 43, 12251239.Google Scholar
Alford, M.H., and Whitmont, M. 2007. Seasonal and spatial variability of near-inertial kinetic energy from historical moored velocity records. J. Phys. Oceanogr., 37, 20222037.CrossRefGoogle Scholar
Allen, K.R., and Joseph, R.I. 1989. A cannonical statistical theory of oceanic internal waves. J. Fluid Mech., 204, 185228.Google Scholar
Annis, A., and Moum, J.N. 1995. Surface wave-turbulence interactions: Scaling ϵ(z) near the sea surface. J. Phys. Oceanogr., 25, 20252045.Google Scholar
Apel, J.R. 2002. Oceanic internal waves and solitons. In: An Atlas of Oceanic Internal Solitary Waves (May 2002). Global Ocean Associates.Google Scholar
Armi, L., and Farmer, D. 2002. Stratified flow over topography: Bifurcation fronts and transition to the uncontrolled state. Proc. Roy. Soc. Lond. A, 458, 513538.Google Scholar
Armi, L., Hebert, D., Oakey, N., and Price, J.F. 1989. Two years in the life of a Mediterranean salt lens. J. Phys. Oceanogr., 19, 354370.2.0.CO;2>CrossRefGoogle Scholar
Arons, A.B. 1981. The scientific work of Henry Stommel. Pages xiv–xviii of: Warren, B.A., and Wunsch, C. (eds), Evolution of Physical Oceanography. Boston: MIT Press.Google Scholar
Aucan, J., Merrifield, M.A., and Luther, D.S. 2006. Tidal mixing events on the deep flanks of Kaena Ridge, Hawaii. J. Phys. Oceanogr., 36, 12021219.CrossRefGoogle Scholar
Baines, P.G., and Gill, A.E. 1969. On thermohaline convection with linear gradients. J. Fluid Mech., 37, 289306.Google Scholar
Batchelor, G.K. 1946. The theory of axisymmetric turbulence. Proc. Roy. Soc. London Ser. A, 186, 480502.Google Scholar
Batchelor, G.K. 1959. Small-scale variation of convected quantitites like temperature in turbulent fluid. Part 1. General discussion and the case of small conductivity. J. Fluid Mech., 5, 113–139.Google Scholar
Beaird, N., Fer, I., Rhines, P., and Eriksen, C. 2012. Dissipation of turbulent kinetic energy inferred from Seagliders: An application to the eastern Nordic Seas overflows. J. Phys. Oceanogr., 42, 22682282.CrossRefGoogle Scholar
Bebieva, Y., and Timmermans, M.-L. 2017. The relationship between double diffusive intrusions and staircases in the Arctic Ocean. J. Phys. Oceanogr., 47(4), 867878.Google Scholar
Bell, T.H., 1975a. Lee waves in stratified flow with simple harmonic time dependence. J. Fluid Mech., 67, 705722.CrossRefGoogle Scholar
Bell, T.H. 1975b. Topographically generated internal waves in the open ocean. J. Geophys. Res., 80, 320337.Google Scholar
Bender, C.M., and Orszag, S.A. 1978. Advanced Mathematical Methods for Scientists and Engineers. New York, NY: McGraw-Hill.Google Scholar
Bertuccioli, L., Roth, G.I., Katz, J., and Osborn, T.R. 1999. A submersible particle image velocimetry system for turbulence measurements in the bottom boundary layer. J. Atmos. Ocean. Tech., 16, 16351646.Google Scholar
Biescas, B., Armi, L., Sallarès, V., and Gràcia, E. 2010. Seismic imaging of staircase layers below the Mediterranean undercurrent. Deep-Sea Res. I, 57, 13451353.CrossRefGoogle Scholar
Billant, P., and Chomaz, J.-M. 2000a. Experimental evidence for a new instability of a vertical columnar vortex pair in a strongly stratified fluid. J. Fluid Mech, 418, 167188.Google Scholar
Billant, P., and Chomaz, J.-M. 2000b. Three-dimensional stability of a vertical columnar vortex pair in a stratified fluid. J. Fluid Mech., 419, 65–91B.Google Scholar
Billant, P., and Chomaz, J.-M. 2001. Self-similarity of strongly stratified inviscid flows. Phys. Fluids, 13(6), 16451651.Google Scholar
Bogucki, D., and Garrett, C. 1993. A simple model for the shear-induced decay of an internal solitary wave. J. Phys. Oceanogr., 23, 17671776.Google Scholar
Bogucki, D., Dickey, T., and Redekop, L.G. 1997. Sediment resuspension and mixing by resonantly generated internal solitary waves. J. Phys. Oceanogr., 27, 11811196.Google Scholar
Bolgiano, R. 1959. Turbulent spectra in a stably stratified atmosphere. J. Geophys. Res., 64, 22262229.Google Scholar
Bouruet-Aubertot, P., Koudella, C., Staquet, C., and Winters, K.B. 2001. Particle dispersion and mixing by breaking internal gravity waves. Dyn. Atmos. Ocean, 33, 95134.Google Scholar
Brainerd, K.E., and Gregg, M.C. 1992. Turbulent decay during diurnal restratification of the oceanic mixed layer. Pages (J8)203–(J8)206 of: AMS 10th Symposium on Turbulence and Diffusion, September 29–October 2, 1992, Portland, Oregon.Google Scholar
Brandt, P., Rubino, A., Alpers, W., and Backhaus, J.O. 1997. Internal waves in the Strait of Messina studied by a numerical model and synthetic aperature radar images from the ERS 1/2 satellites. J. Phys. Oceanogr., 27, 648663.2.0.CO;2>CrossRefGoogle Scholar
Bretherton, F.P. 1966. The propagation of groups of internal gravity waves in a shear flow. Quart. J. Roy. Meteor. Soc., 92, 466480.CrossRefGoogle Scholar
Bretherton, F.P., and Garrett, C.J.R. 1968. Wavetrains in inhomogeneous moving media. Proc. Roy. Soc. Lond. A, 302, 529554.Google Scholar
Brethouwer, G., Billiant, P., Lindborg, E., and Chomaz, J.-M. 2007. Scaling analysis and simulation of strongly stratified turbulent flows. J. Fluid Mech., 585, 343368.Google Scholar
Briscoe, M.G. 1975. Preliminary results from the trimoored Internal Waves Experiment. J. Geophys. Res., 80(27), 38723884.CrossRefGoogle Scholar
Briscoe, M.G., and Weller, R.A. 1984. Preliminary results from the Long-term Upper-Ocean Study (LOTUS). Dyn. Atmos. Oceans, 8, 243265.Google Scholar
Brown, N. 1974. A precision CTD microprofiler. Pages 270–278 of: Engr. in the Ocean Envir., Ocean’74, vol. 2.Google Scholar
Browne, L.W.B., Antonia, R.A., and Shad, D.A. 1987. Turbulent energy dissipation in a wake. J. Fluid Mech., 179, 307326.Google Scholar
Bryden, H.L., and Nurser, A.J.G. 2003. Effect of strait mixing on ocean stratification. J. Phys. Oceanogr., 33, 18701872.Google Scholar
Cacchione, D.A., Pratson, L.F., and Ogston, A.S. 2002. The shaping of continental slopes by internal tides. Science, 296, 724727.Google Scholar
Cairns, J.L., and Williams, G.O. 1976. Internal wave observations from a midwater float, 2. J. Geophys. Res., 81, 19431950.Google Scholar
Caldwell, D.R. 1973. Thermal and Fickian diffusion of sodium chloride in a solution of oceanic concentration. Deep-Sea Res., 20, 10291039.Google Scholar
Caldwell, D.R. 1974a. The effect of pressure on thermal and Fickian diffusion of sodium chloride. Deep-Sea Res., 21, 369375.Google Scholar
Caldwell, D.R. 1974b. Experimental studies on the onset of thermohaline convection. J. Fluid Mech., 64, 347367.Google Scholar
Caldwell, D.R., Wilcox, S.D., and Matsier, M. 1975. A relatively simple freely-falling probe for small-scale temperature gradients. Limnol. Oceanogr., 20, 10351047.Google Scholar
Carmack, E.C., Williams, W.J., Zimmerman, S.L., and McLaughlin, F.A. 2012. The Arctic Ocean warms from below. Geophys. Res. Lett., 39(L07604).Google Scholar
Carpenter, J.R., Sommer, T., and Wüest, A. 2012a. Simulations of a double-diffusive interface in the diffusive convection regime. J. Fluid Mech., 711, 411436.Google Scholar
Carpenter, J.R., Sommer, T., and Wüest, A. 2012b. Stability of a double-diffusive interface in the diffusive convection regime. J. Phys. Oceanogr., 42, 840854.Google Scholar
Carter, G.S., and Gregg, M.C. 2006. Persistent near-diurnal internal waves observed above a site of M2 barotropic-to-baroclinic conversion. J. Phys. Oceanogr., 36, 11361147.Google Scholar
Carter, G.S., Gregg, M.C., and Lien, R.-C. 2005. Internal waves, solitary-like waves, and mixing on the Monterey Bay shelf. Cont. Shelf Res., 25, 14991520.Google Scholar
Chang, M.-H., Lien, R.-C., Yang, Y.J., Tang, T.Y., and Wang, J. 2008. A composite view of surface signatures and interior properties of nonlinear internal waves: Observations and applications. J. Atmos. Ocean. Tech., 25, 12181227.Google Scholar
Corrsin, S. 1963. Estimates of the relations between Eulerian and Lagrangian scales in large Reynolds number turbulence. J. Atmos. Sci., 20, 115119.Google Scholar
Cox, C.S., Nagata, Y., and Osborn, T. 1969. Oceanic fine structure and internal waves. Bull. Japanese Soc. Fisheries Oceanography, Papers in Dedication to Prof. Michitaka Uda(Nov.), 67–71.Google Scholar
Crapper, P.F. 1975. Measurements across a diffusive interface. Deep-Sea Res., 22, 537545.Google Scholar
Curry, J.A., and Webster, P.J. 1999. Thermodynamics of Atmospheres and Oceans. Intl. Geophys. Ser., vol. 65. 24–28 Oval Rd., London, NW1 7DX, UK: Academic Press.Google Scholar
D’Asaro, E.A. 1989. The decay of wind-forced mixed layer inertial oscillations due to the β effect. J. Geophys. Res., 94, 20452056.Google Scholar
D’Asaro, E.A. 2001. Turbulent vertical kinetic energy in the ocean mixed layer. J. Phys. Oceanogr., 31, 35303537.Google Scholar
D’Asaro, E.A., Farmer, D.M., Osse, J.T., and Dairiki, G.T. 1996. A Lagrangian float. J. Atmos. Ocean. Tech., 13(6), 12301246.Google Scholar
D’Asaro, E.A., and Lien, R.-C. 2000a. Lagrangian measurements of waves and turbulence in stratified flows. J. Phys. Oceanogr., 30, 641655.Google Scholar
D’Asaro, E.A., and Lien, R.-C. 2000b. The wave-turblence transition for stratified flows. J. Phys. Oceanogr., 30, 16691678.Google Scholar
D’Asaro, E.A., and Lien, R.-C. 2007. Measurement of scalar variance dissipation from Lagrangian floats. J. Atmos. Ocean. Tech., 24, 10661077.Google Scholar
D’Asaro, E.A., and Morehead, M.D. 1991. Internal waves and velocity fine structure in the Arctic Ocean. J. Geophys. Res., 96, 12,725–12.738.Google Scholar
D’Asaro, E.A., and Morison, J.H. 1992. Internal waves and mixing in the Arctic Ocean. Deep-Sea Res A, 39(2), S459–S484.Google Scholar
Davis, R.E. 1994a. Diapycnal mixing in the ocean: Equations for large-scale budgets. J. Phys. Oceanogr., 24, 777800.Google Scholar
Davis, R.E. 1994b. Diapycnal mixing in the ocean: The Osborn-Cox model. J. Phys. Oceanogr., 24, 25602576.Google Scholar
Davis, R.E., Regier, L.A., Dufour, J., and Webb, D.C. 1992. The Autonomous Lagrangian Circulation Explorer (ALACE). J. Atmos. Ocean. Tech., 9, 264285.Google Scholar
Davis, R.E., Sherman, J.T., and Dufour, J. 2001. Profiling ALACEs and other advances in autonomous subsurface floats. J. Atmos. Ocean. Tech., 18, 982993.Google Scholar
de Groot, S.R., and Mazur, P. 1969. Non-Equilibrium Thermodynamics. Amsterdam: North-Holland.Google Scholar
Desaubies, Y.J.F., and Gregg, M.C. 1981. Reversible and irreversible finestructure. J. Phys. Oceanogr., 11, 541556.Google Scholar
Desaubies, Y.J.F., and Smith, W.K. 1982. Statistics of Richardson number and instability in oceanic internal waves. J. Phys. Oceanogr., 12, 12451259.Google Scholar
Dickson, R.R., and Brown, J. 1994. The production of North Atlantic Deep Water: Sources, rates, and pathways. J. Geophys. Res., 99, 12,319–12,341.Google Scholar
Dillon, T.M. 1982. Vertical overturns: A comparison of Thorpe and Ozmidov length scales. J. Geophys. Res., 87, 96019613.Google Scholar
Dillon, T.M., and Caldwell, D.R. 1980. The Batchelor spectrum and dissipation in the upper ocean. J. Geophys. Res., 85(C4), 19101916.Google Scholar
Döös, K, and Coward, A. 1997. The Southern Ocean as the major upwelling zone of North Atlantic Deep Water. WOCE newsletter, 27, 34.Google Scholar
Doron, P., Bertuccioli, L., Katz, J., and Osborn, T.R. 2001. Turbulence characteristics and dissipation estimates in the coastal ocean bottom boundary layer using PIV data. J. Phys. Oceanogr., 31(8), 21082134.Google Scholar
Dosser, H.V., and Rainville, L. 2016. Dynamics of the changing near-inertial internal wave field in the Arctic Ocean. J. Phys. Oceanogr., 46(2), 395415.Google Scholar
Dunkerton, T.J. 1997. Shear instability of internal inertial-gravity waves. J. Atmos. Sci., 54, 16281641.Google Scholar
Dushaw, B.D., Cornuelle, B.D., Worcester, P.F., Howe, B.M., and Luther, D.S. 1995. Barotropic and baroclinic tides in the central North Pacific Ocean determined from long range reciprocal acoustic transmissions. J. Phys. Oceanogr., 25, 631647.Google Scholar
Efron, B., and Gong, G. 1983. A leisurely look at at the bootstrap, the jackknife, and cross-validation. Amer. Stat., 37, 3648.Google Scholar
Egbert, G.D., and Ray, R.D. 2000. Significant dissipation of tidal energy in the deep ocean inferred from satellite altimeter data. Nature, 405(15 June 2000), 775–778.Google Scholar
Egbert, G.D., and Ray, R.D. 2001. Estimates of M2 tidal energy dissipation from TOPEX/Poseidon altimeter data. Geophys. Res. Ltrs., 106(C10), 22,475–22,502.Google Scholar
Ekman, V.W. 1905. On the influence of the Earth’s rotation on ocean-currents. Ark. Mat. Astron. Fys., 2, 152.Google Scholar
Elliott, J.A., and Oakey, N.S. 1975. Horizontal coherence of temperature microstructure. J. Phys. Oceanogr., 5(7), 506515.Google Scholar
Elliott, J.A., and Oakey, N.S. 1976. Spectrum of small-scale oceanic temperature gradient. J. Fish. Res. Bd. Canada, 33, 22962306.CrossRefGoogle Scholar
Ellison, T.H. 1957. Turbulent transport of heat and momentum from an infinite rough plane. J. Fluid Mech, 2, 456466.Google Scholar
Eriksen, C.C. 1982. Observations of internal wave reflection off sloping bottoms. J. Geophys. Res., 87, 525538.Google Scholar
Eriksen, C.C., Osse, T.J., Light, R.D., Wen, T., Lehman, T.W., Sabin, P.L., Ballard, J.W., and Chiodi, A.M. 2001. Seaglider: A long-range autonomous underwater vehicle for oceanographic research. IEEE J. Ocean Engr., 26(4), 424436.Google Scholar
Evans, D.L., Rossby, H.T., Mork, M., and Gytre, T. 1979. YVETTE – a free-fall shear profiler. Deep-Sea Res A, 26, 703718.Google Scholar
Ewart, T.E., and Bendiner, W.P. 1981. An observation of the horizontal and vertical diffusion of a passive tracer in the deep ocean. J. Geophys. Res., 86(C11), 10,974–10,982.Google Scholar
Fabula, A.G. 1968a. The dynamic response of towed thermometers. J. Fluid Mech., 34, 449464.Google Scholar
Fabula, A.G. 1968b. The dynamic response of towed thermometers. J. Fluid Mech., 34(3), 449464.Google Scholar
Farmer, D.M., and Smith, J.D. 1980. Tidal interaction of stratified flow with a sill in Knight Inlet. Deep-Sea Res., 27A, 239254.Google Scholar
Feistel, R. 2008. A Gibbs function for seawater thermodynamics for −6o to 80o C and salinity up to 120 gkg−1. Deep-Sea Res. I, 55, 16391671.Google Scholar
Fer, I., Peterson, A.K., and Ullgren, J.E. 2014. Microstructure measurements from an underwater glider in the turbulent Faroe Bank Channel overflow. J. Atmos. Ocean. Tech., 31, 11281150.Google Scholar
Fernando, H.J.S. 1987. The formation of a layered structure when a stable salinity gradient is heated from below. J. Fluid Mech., 182, 525541.Google Scholar
Fernando, H.J.S. 1989. Oceanographic implications of laboratory experiments on diffusive interfaces. J. Phys. Oceanogr., 19, 17071715.Google Scholar
Ferrari, R., Mashayek, A., McDougall, T.J., Nukurashin, M., and Campin, J.-M. 2016. Turning ocean mixing upside down. J. Phys. Oceanogr., 46, 22392260.CrossRefGoogle Scholar
Ferrari, R., and Rudnick, D.L. 2000. Thermohaline variability in the upper ocean. J. Geophys. Res., 105, 16,857–16,883.Google Scholar
Ferron, B., Mercier, H., Speer, K., Gargett, A., and Polzin, K. 2003. Mixing in the Romanche Fracture Zone. J. Phys. Oceanogr., 28, 19291945.Google Scholar
Flanagan, J.D., Lefler, A.S., and Radko, R. 2013. Heat transport through diffusive interfaces. Geophys. Res. Lett., 40, 24662470.CrossRefGoogle Scholar
Flanagan, J.D., Radko, T., Shaw, W.J., and Stanton, T.P. 2014. Dynamic and double-diffusive instabilities in a weak pycnocline. Part II: Direct numerical simulations and flux laws. J. Phys. Oceanogr., 44, 19922012.CrossRefGoogle Scholar
Flatté, S.M., Henyey, F.S., and Wright, J.A. 1985. Eikonal calculations of short-wavelength internal wave spectra. J. Geophys. Res., 90, 72657272.Google Scholar
Fleury, M., and Lueck, R.G. 1994. Direct heat flux estimates using a towed vehicle. J. Phys. Oceanogr., 24(4), 801818.Google Scholar
Fofonoff, N.P. 1962. Physical properties of seawater. Pages 3–30 of: Hill, M.N. (ed.), The Sea, vol. 1. New York, NY: Wiley-Interscience.Google Scholar
Fofonoff, N.P. 1969. Spectral characteristics of internal waves in the ocean. Deep-Sea Res, 16 , Suppl., 5871.Google Scholar
Fofonoff, N.P. 1998. Nonlinear limits to ocean thermal structure. J. Mar. Res., 56(4), 783811.Google Scholar
Fofonoff, N.P. 2001. Thermal stability of the world ocean thermocline. J. Phys. Oceanogr., 31(8), 21692177.Google Scholar
Foster, T.D. 1972. An analysis of the cabbeling instability in sea water. J. Phys. Oceanogr., 2, 294301.Google Scholar
Foster, T.D., and Carmack, E.C. 1976. Temperature and salinity structure in the Weddell Sea. J. Phys. Oceanogr., 6(1), 3644.Google Scholar
Frants, M., Damerell, G.M., Gille, S.T., Heywood, K.J., MacKinnon, J., and Sprintall, J. 2013. An assessment of density-based finescale methods for estimating diapycnal diffusivity in the Southern Ocean. J. Atmos. Ocean. Tech., 30(11), 26472661.Google Scholar
Fringer, O.B., and Street, R.L. 2003. The dynamics of breaking progressive interfacial waves. J. Fluid Mech., 494, 319353.CrossRefGoogle Scholar
Fu, L.-L. 1981. Observations and models of inertial waves in the deep ocean. Rev. Geophys. and Space Phys., 19(1), 141170.Google Scholar
Furuichi, N., Hibiya, T., and Niwa, Y. 2005. Bispectral analysis of energy transfer within the two-dimensional oceanic internal wave field. J. Phys. Oceanogr., 35, 21042109.CrossRefGoogle Scholar
Galbraith, P.S., and Kelley, D.E. 1996. Identifying overturns in CTD profiles. J. Atmos. Ocean. Tech., 13, 688702.Google Scholar
Gargett, A.E. 1985. Evolution of scalar spectra with the decay of turbulence in a stratified fluid. J. Fluid Mech., 159, 379407.Google Scholar
Gargett, A.E. 1990. Do we really know how to scale the turbulent kinetic energy dissipation rate ϵ due to breaking of oceanic internal waves? J. Geophys. Res., 95, 15,971–15,974.Google Scholar
Gargett, A.E. 2003. Differential diffusion: an oceanographic primer. Progr. Oceanogr., 56, 559570.Google Scholar
Gargett, A.E., Hendricks, P.J., Sanford, T.B., Osborn, T.R., and Williams, III, A.J. 1981. A composite spectrum of vertical shear in the upper ocean. J. Phys. Oceanogr., 11, 12581271.Google Scholar
Gargett, A.E., Merryfield, W., and Holloway, G. 2003. Direct numerical simulation of differential scalar diffusion in three-dimensional stratified turbulence. J. Phys. Oceanogr., 33, 17581782.Google Scholar
Gargett, A.E., Osborn, T.R., and Nasmyth, P.W. 1984. Local isotropy and the decay of turbulence in a stratified fluid. J. Fluid Mech., 144, 231280.Google Scholar
Gargett, A.E., and Schmitt, R.W. 1982. Observations of salt fingers in the central waters of the eastern North Pacific. J. Geophys. Res., 87, 80178029.Google Scholar
Garrett, C. 2001. What is the ‘near-inertial’ band and why is it different from the rest of the internal wave spectrum? J. Phys. Oceanogr., 31, 962971.Google Scholar
Garrett, C., and Horne, E. 1978. Frontal circulation due to cabbeling and double diffusion. J. Geophys. Res., 83, 46514656.Google Scholar
Garrett, C.J.R, and Munk, W.H., 1972. Space-time scales of internal waves. Geophys. Fluid Dyn., 2, 225264.Google Scholar
Garrett, C.J.R., and Munk, W.H. 1975. Space-time scales of internal waves: A progress report. J. Geophys. Res., 80, 291297.Google Scholar
Gibson, C.H., and Schwarz, W.H. 1963. The universal equilibrium spectra of turbulent velocity and scalar fields. J. Fluid Mech., 16, 365384.CrossRefGoogle Scholar
Giles, A.B., and McDougall, T.J. 1986. Two methods for the reduction of salinity spiking of CTDs. Deep-Sea Res., 33, 12531274.Google Scholar
Gill, A.E. 1982. Atmosphere-Ocean Dynamics. New York, NY: Academic.Google Scholar
Gill, A.E. 1984. On the behavior of internal waves in the wake of a storm. J. Phys. Oceanogr., 14, 11291151.Google Scholar
Gille, S.T. 1997. The Southern Ocean momentum balance: Evidence for topographic effects from numerical model output and altimeter data. J. Phys. Oceanogr., 27, 22192232.Google Scholar
Gille, S.T. 2004. How nonlinearities in the equation of state of seawater can confound estimates of steric sea level change. J. Geophys. Res. Oceans, 109(C03005).CrossRefGoogle Scholar
Goodman, L. 1990. Acoustic scattering from ocean microstructure. J. Geophys. Res., 95(C7), 11,557–11,573.Google Scholar
Goodman, L., Levine, E.R., and Lueck, R.G. 2006. On measuring the terms of the turbulent kinetic energy budget from an AUV. J. Atmos. Ocean. Tech., 23(7), 977990.Google Scholar
Goodman, L., and Wang, Z. 2009. Turbulence observations in the northern bight of Monterey Bay from a small AUV. J. Mar. Syst., 77(4), 441458.Google Scholar
Gordon, A.L. 2013. Bottom water formation. In: Reference Module in Earth Systems and Environmental Sciences. Elsevier.Google Scholar
Graham, F.S., and McDougall, T.J. 2013. Quantifying the nonconservative production of conservative temperature, potential temperature, and entropy. J. Phys. Oceanogr., 43, 838862.Google Scholar
Grant, H.L., Hughes, B., Vogel, W., and Moillet, A. 1968b. The spectrum of temperature fluctuations in turbulent flow. J. Fluid Mech., 34, 423441.Google Scholar
Grant, H.L., Moilliet, A., and Stewart, R.W. 1959. A spectrum of turbulence at very high Reynolds number. Nature, 184, 808810.Google Scholar
Grant, H.L., Moilliet, A., and Vogel, W.M. 1968a. Some observations of the occurrence of turbulence in and above the thermocline. J. Fluid Mech., 33, 443448.Google Scholar
Grant, H.L., Stewart, R.W., and Moilliet, A. 1962. Turbulence spectra from a tidal channel. J. Fluid Mech., 12, 241268.Google Scholar
Gregg, M.C. 1968. Mechanical stirring and salt fingers. Pages 42–51 of: WHOI Summer Institute of Geophysical Fluid Dynamics Participant Reports, vol. II. Woods Hole, MA: Woods Hole Oceanographic Institution.Google Scholar
Gregg, M.C. 1975a. Microstructure and intrusions in the California Current. J. Phys. Oceanogr., 5, 253278.Google Scholar
Gregg, M.C. 1975b. Oceanic fine and microstructure. Rev. Geophys. and Space Phys., 13, 586–591 and 635–636.Google Scholar
Gregg, M.C. 1977a. Cruise report of the Mixed Layer Experiment (MILE) on the U.S.N.S. DE STEIGUER, 18 August – 8 September 1977. Tech. rept. APL-UW Technical Note 5–77. Applied Physics Laboratory, University of Washington, Seattle, WA.Google Scholar
Gregg, M.C. 1977b. Variations in the intensity of small-scale mixing in the main thermocline. J. Phys. Oceanogr., 7, 436454.Google Scholar
Gregg, M.C. 1980a. Microstructure patches in the thermocline. J. Phys. Oceanogr., 10, 915943.Google Scholar
Gregg, M.C. 1980b. The three-dimensional mapping of a small thermohaline intrusion. J. Phys. Oceanogr., 10, 14681492.Google Scholar
Gregg, M.C. 1984. Entropy generation in the ocean by small-scale mixing. J. Phys. Oceanogr., 14, 688711.Google Scholar
Gregg, M.C. 1987. Structures and fluxes in a deep convecting mixed layer. Pages 1–23 of: Muller, P., and Henderson, D. (eds.), Dynamics of the Oceanic Surface Mixed Layer, Proceedings,’Aha Huliko’a Hawaiian Winter Workshop, University of Hawaii at Manoa, January 14–16, 1987. Hawaii Institute of Geophysics.Google Scholar
Gregg, M.C. 1989a. Scaling turbulent dissipation in the thermocline. J. Geophys. Res., 94, 96869698.Google Scholar
Gregg, M.C. 1989b. Small-scale mixing: A first-order process? Pages 117–126 of: Müller, P., and Henderson, D. (eds.), Parameterization of Small-Scale Processes: Proceedings, ‘Aha Huliko’a Hawaiian Winter Workshop, University of Hawaii at Manoa, January 17–20, 1989. Hawaii Institute of Geophysics. Special Publication.Google Scholar
Gregg, M.C. 1991. The study of mixing in the ocean: A brief history. Oceanography, 4, 3945.Google Scholar
Gregg, M.C. 1999. Uncertainities and limitations in measuring ϵ and χT. J. Atmos. Ocean. Tech., 16, 14831490.Google Scholar
Gregg, M.C., Alford, M.H., Kontoyiannis, H., Zervakis, V., and Winkel, D. 2012. Mixing over the steep side of the Cycladic Plateau in the Aegean Sea. J. Mar. Syst, 89, 3047.Google Scholar
Gregg, M.C., and Cox, C.S. 1971. Measurements of the oceanic microstructure of temperature and electrical conductivity. Deep-Sea Res., 18, 925934.Google Scholar
Gregg, M.C., and Cox, C.S. 1972. The vertical microstructure of temperature and salinity. Deep-Sea Res., 19, 355376.Google Scholar
Gregg, M.C., Cox, C.S., and Hacker, P.W. 1973. Vertical microstructure measurements in the Central North Pacific. J. Phys. Oceanogr., 3, 458469.Google Scholar
Gregg, M.C., D’Asaro, E.A., Shay, T.J., and Larson, N. 1986. Observations of persistent mixing and near-inertial internal waves. J. Phys. Oceanogr., 16, 856885.Google Scholar
Gregg, M.C., D’Asaro, E.A., Riley, J.J., and Kunze, E. 2018. Mixing efficiency in the ocean. Annu. Rev. Mar. Sci., 10, 443473.Google Scholar
Gregg, M.C., and Hess, W.C. 1985. Dynamic response calibration of Sea-Bird temperature and conductivity probes. J. Atmos. Ocean. Tech., 2, 304313.2.0.CO;2>CrossRefGoogle Scholar
Gregg, M.C., and Kunze, E. 1991. Shear and strain in Santa Monica Basin. J. Geophys. Res., 96, 16,709–16,719.Google Scholar
Gregg, M.C., and Meagher, T.B. 1980. The dynamic response of glass-rod thermistors. J. Geophys. Res., 85, 27792786.Google Scholar
Gregg, M.C., Meagher, T.B., Pederson, A.M., and Aagaard, E.A. 1978. Low noise temperature microstructure measurements with thermistors. Deep-Sea Res., 25, 843856.Google Scholar
Gregg, M.C., Meagher, T.B., Aagaard, E.E., and Hess, W.C. 1981. A salt-stratified tank for measuring the dynamic response of conductivity probes. IEEE J. of Oceanic Engr., OE-6, 113118.Google Scholar
Gregg, M.C., Nodland, W.E., Aagaard, E.E., and Hirt, D.H. 1982. Use of a fiber-optic cable with a free-fall microstructure profiler. Pages 260–265 of: Oceans’82: Conference Record, Sept. 20–22, 1982. Washington, DC: Marine Technology Soc.Google Scholar
Gregg, M.C., and Özsoy, E. 2002. Flow, water mass changes and hydraulics in the Bosphorus. J. Geophys. Res., 107(C3), 2000JC000485.Google Scholar
Gregg, M.C., and Sanford, T.B. 1980. Signatures of mixing from the Bermuda Slope, the Sargasso Sea and the Gulf Stream. J. Phys. Oceanogr., 10, 105127.Google Scholar
Gregg, M.C., and Sanford, T.B. 1981. Reply to Dillon and Caldwell. J. Phys. Oceanogr., 11, 14381439.Google Scholar
Gregg, M.C., and Sanford, T.B. 1987. Shear and turbulence in thermohaline staircases. Deep-Sea Res., 34, 16891696.Google Scholar
Gregg, M.C., and Sanford, T.B. 1988. The dependence of turbulent dissipation on stratification in a diffusively stable thermocline. J. Geophys. Res., 93, 12,381–12,392.Google Scholar
Gregg, M.C., Sanford, T.B., and Winkel, D.P. 2003. Reduced mixing from the breaking of internal waves in equatorial ocean waters. Nature, 422, 513515.Google Scholar
Gregg, M.C., Seim, H.E., and Percival, D.B. 1993. Statistics of shear and turbulent dissipation profiles in random internal wave fields. J. Phys. Oceanogr., 23, 17771799.Google Scholar
Gregg, M.C., Winkel, D.P., Sanford, T.B., and Peters, H. 1996. Turbulence produced by internal waves in the oceanic thermocline at mid and low latitudes. Dyn. Atmos. Oceans, 24, 114.Google Scholar
Gross, T.F., Williams, A.J., and Terray, E.A. 1984. Bottom boundary layer spectral dissipation estimates in the presence of wave motions. Cont. Shelf Res., 14(10–11), 12391256.Google Scholar
Groves, G.W. 1959. Flow estimate for the perpetual salt fountain. Deep-Sea Res., 5, 209214.Google Scholar
Gurevich, A.V., and Pitaevskii, L.P. 1973. Nonstationary structure of a collisionless shock wave. Sov. Phys JETP, 38, 291297.Google Scholar
Gurvich, A.S., and Yaglom, A.M. 1993. Breakdown of eddies and probability distributions for small scale turbulence. Phys. Fluids, 10, 5965.Google Scholar
Guthrie, J.D., Fer, I., and Morison, J. 2015. Observational validation of the diffusive flux laws in the Amundsen Basin. J. Geophys. Res., 120, 78807896.Google Scholar
Hasselman, K. 1966. Feynman diagrams and interaction rules of wave-wave scattering processes. Rev. Geophys. and Space Phys., 4, 132.Google Scholar
Hayes, S.P., Joyce, T.M., and Millard, R.C. 1975. Measurements of vertical finestructure. J. Geophys. Res., 80, 314320.Google Scholar
Hayes, S.P., Milburn, H.B., and Ford, E.F. 1984. TOPS: A free-fall velocity and CTD profiler. J. Atmos. Ocean. Tech., 1, 220236.Google Scholar
Hazel, P. 1972. Numerical studies of the stability of inviscid stratified shear flows. J. Fluid Mech., 51, 3961.Google Scholar
Head, M.J. 1983. The Use of Miniature Four-Electrode Conductivity Probes for High Resolution Measurement of Turbulent Density or Temperature Variations in Salt-Stratified Water Flows. Ph.D. thesis, University of California, San Diego.Google Scholar
Hebert, D. 1988a. The available potential energy of an isolated feature. J. Geophys. Res., 93(C1), 556564.Google Scholar
Hebert, D. 1988b. Estimates of salt finger fluxes. Deep-Sea Res., 35(12), 18871901.Google Scholar
Henyey, F.S., and Hoering, A. 1997. Energetics of borelike internal waves. J. Geophys. Res., 102, 33233330.Google Scholar
Henyey, F.S., and Pomphrey, N. 1983. Eikonal description of internal-wave interactions. Dyn. Atmos. Oceans, 7, 189208.Google Scholar
Henyey, F.S., Wright, J., and Flatté, S.M. 1986. Energy and action flow through the internal wave field: An eikonal approach. J. Geophys. Res., 91, 84878495.Google Scholar
Hesselberg, T., and Sverdrup, H.U. 1914. Die stabilitätsverhältnisse des seewassers bei vertikalen verschiebungen. Bergens Mus. Aarb., 15, 116.Google Scholar
Hibiya, T., Niwa, Y., Nakajima, K., and Suginohara, N. 1996. Direct numerical simulation of the roll-off range of internal wave shear spectra in the ocean. J. Geophys. Res., 101(C6), 14,123–14,129.Google Scholar
Hieronymus, M., and Carpenter, J.R. 2016. Energy and variance budgets of a diffusive staircase with implications for heat flux scaling. J. Phys. Oceanogr., 46, 25532569.Google Scholar
Hill, K.D., and Woods, D.J. 1988. The dynamic response of the two-electrode conductivity cell. IEEE J. of Oceanic Engr., 13(3), 118123.Google Scholar
Hinze, J.O. 1975. Turbulence. 2nd ed. New York, NY: Mc-Graw Hill.Google Scholar
Hirst, E. 1991. Internal wave-wave resonance theory: Fundamentals and limitations. Pages 211–226 of: Müller, P., and Henderson, D. (eds.), Dynamics of Oceanic Internal Gravity Waves: Proceedings,’Aha Huliko’a Hawaiian Winter Workshop, University of Hawaii at Manoa, January 11–13, 1991. Hawaii Institute of Geophysics.Google Scholar
Ho, D.T., Ledwell, J.R., and Jr., Smethie, W.M.. 2008. Use of SF5 CF3 for ocean tracer release experiments. Geophys. Res. Lett., 35(L04702).Google Scholar
Holbrook, W.S., and Fer, I. 2005. Oceanic internal wave spectra inferred from seismic reflection transects. Geophys. Res. Lett., 32(L15604).Google Scholar
Holbrook, W.S., Páramo, P., Pearse, S., and Schmitt, R.W. 2003. Thermohaline fine structure in a oceanographic front from seismic reflection profiling. Science, 301(5634), 821824.Google Scholar
Holleman, R.C., Geyer, W.R., and Ralston, D.K. 2016. Stratified turbulence and mixing efficiency in a salt wedge estuary. J. Phys. Oceanogr., 46, 17691783.Google Scholar
Holloway, G. 1980. Oceanic internal waves are not weak waves. J. Phys. Oceanogr., 10, 906914.Google Scholar
Holloway, G. 1982. On interaction time scales of oceanic internal waves. J. Phys. Oceanogr., 12, 293296.Google Scholar
Holloway, G., and Gargett, A.E. 1987. The inference of salt fingering from towed microstructure observations. J. Geophys. Res., 92, 19631966.Google Scholar
Holloway, P.E., and Merrifield, M.A. 2003. On the spring-neap variability and age of the internal tide at the Hawaiian Ridge. J. Geophys. Res., 108(C4), 3126, doi:10.1029/2002JC001486.Google Scholar
Holtermann, P.L., Umlauf, L., Tanhua, T., Schmale, O., Rehder, G., and Waniek, J.J. 2012. The Baltic Sea tracer release experiment. 1. Mixing rates. J. Geophys. Res., 117(C01021).Google Scholar
Horne, E.P.W., and Toole, J.M. 1980. Sensor response mismatches and lag correction techniques for temperature-salinity profilers. J. Phys. Oceanogr., 10(7), 11221130.Google Scholar
Howard, L.N. 1961. Note on a paper by John W. Miles. J. Fluid Mech., 10, 509512.Google Scholar
Huang, R.X. 2005. Available potential energy in the world’s oceans. J. Mar. Res., 63(1), 141158.Google Scholar
Huppert, H.E. 1971. On the stability of a series of double-diffusivity layers. Deep-Sea Res., 18, 10051021.Google Scholar
Huppert, H.E., and Linden, P.F. 1979. On heating a stable salinity gradient from below. J. Fluid Mech., 95, 431464.Google Scholar
Ijichi, T., and Hibiya, T. 2015. Frequency-based correction of finescale parameterization of turbulent dissipation in the deep ocean. J. Atmos. Ocean. Tech., 32, 15261535.Google Scholar
Ingham, M.C. 1966 (June). The Salinity Extrema of the World Ocean. PhD thesis, Oregon State University, Corvalis, OR.Google Scholar
Inoue, R., Kunze, E., Laurent, L. St., Schmitt, R.W., and Toole, J.M. 2008. Evaluating salt-fingering theories. J. Mar. Res., 66(4), 413440.Google Scholar
IOC, SCOR, and IAPSO. 2010. The International Thermodynamic Equation of Seawater – 2010: Calculation and Use of Thermodynamic Properties. Manuals and Guides 56. Intergovernmental Oceanographic Commission, UNESCO.Google Scholar
Irish, J.D., and Nodland, W.E. 1978. Evaluation of metal-film temperature and velocity sensors and the stability of a self-propelled research vehicle for making measurements of ocean turbulence. Pages 180–187 of: IEEE/MTS Proceeding of OCEANS-78, September 1978, Washington, D.C.Google Scholar
Iselin, C. O’D. 1939. The influence of vertical and lateral turbulence on the characteristics of the waters at mid-depths. Trans., American Geophys. Union, 20(3), 414417.Google Scholar
Ivers, W.D. 1975. The Deep Circulation in the Northern North Atlantic, with Especial Reference to the Labrador Sea. PhD thesis, University of California San Diego, San Diego, CA.Google Scholar
Ivey, G.N., and Imberger, J. 1991. On the nature of turbulence in a stratified fluid. Part I: The energetics of mixing. J. Phys. Oceanogr., 21, 650658.Google Scholar
Jackett, D.R., and McDougall, T.J. 1985. An oceanographic variable for the characterization of intrusions and water masses-. Deep-Sea Res., Part A, 32, 11951207.Google Scholar
Jackett, D.R., and McDougall, T.J. 1997. A neutral density surface for the world’s oceans. J. Phys. Oceanogr., 27(2), 237263.Google Scholar
Jackson, P.R., and Rehmann, C.R. 2003. Laboratory measurements of differential diffusion in a diffusively stable, turbulent flow. J. Phys. Oceanogr., 33(8), 15921603.Google Scholar
Jones, W.P., and Musonge, P. 1988. Closure of the Reynolds stress and scalar flux equations. Phys. Fluids, 31.Google Scholar
Joyce, T.M. 1977. A note on the lateral mixing of water masses. J. Phys. Oceanogr., 7, 626629.Google Scholar
Joyce, T.M. 1980. On production and dissipation of thermal variance in the ocean. J. Phys. Oceanogr., 10, 460463.Google Scholar
Kao, T.W., and Pao, H.-P. 1978. Note on the flow of a stratified fluid over a stationary obstacle in a channel. Geophys. Astrophys. Fluid Dyn., 10(1), 109114.CrossRefGoogle Scholar
Kelley, D.E. 1984. Effective diffusivities within oceanic thermohaline staircases. J. Geophys. Res., 89, 10,484–10,488.Google Scholar
Kelley, D.E. 1989. Explaining effective diffusivities within diffusive oceanic staircases. Pages 481–502 of: Nihoul, J.C.J., and Jamart, B.M. (eds.), Small-Scale Turbulence and Mixing in the Ocean. Amsterdam: Elsevier.Google Scholar
Kelley, D.E. 1990. Fluxes through diffusive staircases, a new formulation. J. Geophys. Res., 95, 33653371.Google Scholar
Kelley, D.E., Fernando, H.J.S., Gargett, A.E., Tanny, J., and Özsoy, E. 2003. The diffusive regime of double diffusion. Prog. Oceanogr., 56(3–4), 461481.Google Scholar
Kennelly, M.A., McKeown, P.A., and Sanford, T.B. 1986. XCP Performances Near the Geomagnetic Equator. Report APL-UW Tech. Rept. 8607. University of Washington, Applied Physics Laboratory, Seattle, WA.Google Scholar
Killworth, P.D. 1977. Mixing on the Weddell Sea Continental Slope. Deep-Sea Res., 24, 427448.Google Scholar
Kimura, S., and Smyth, W. 2007. Direct numerical simulation of salt sheets and turbulence in a double-diffusive shear layer. Geophys. Res. Lett., 34(L21610).Google Scholar
Kimura, S., and Smyth, W. 2011. Turbulence in a sheared, salt-fingering-favorable environment: Anisotropy and effective diffusivities. J. Phys. Oceanogr., 41(6), 11411159.Google Scholar
Klaassen, G.P., and Peltier, W.R. 1985. The onset of turbulence in finite-amplitude Kelvin– Helmholtz billows. J. Fluid Mech., 155, 135.Google Scholar
Klocker, A., and McDougall, T.J. 2010. Influence of the nonlinear equation of state on global estimates of dianeutral advection and diffusion. J. Phys. Oceanogr., 40, 16901709.Google Scholar
Klymak, J.M., and Gregg, M.C. 2004. Tidally generated turbulence over the Knight Inlet sill. J. Phys. Oceanogr., 34, 11351151.Google Scholar
Klymak, J.M., Legg, S.M., and Pinkel, R. 2010. High-mode stationary waves in stratified flow over large obstacles. J. Fluid Mech., 664(321–336).Google Scholar
Klymak, J.M., and Moum, J.N. 2006. Oceanic isopycnal slope spectra: Part I: Internal waves. J. Phys. Oceanogr., 36(12), 116.Google Scholar
Klymak, J.M., and Moum, J.N. 2007. Oceanic isopycnal slope spectra. Part II: Turbulence. J. Phys. Oceanogr., 37, 12321245.Google Scholar
Klymak, J.M., Moum, J.N., Nash, J.D., Kunze, E., Girton, J.B., Carter, G.S., Lee, C.M., Sanford, T.B., and Gregg, M.C. 2006a. An estimate of tidal energy lost to turbulence at the Hawaiian Ridge. J. Phys. Oceanogr., 36, 11481164.Google Scholar
Klymak, J.M., Pinkel, R., Liu, C.-T., Liu, A.K., and David, L. 2006b. Prototypical solitons in the South China Sea. Geophys. Res. Lett., 33(L11607).Google Scholar
Kolmogorov, A.N. 1941. The local structure of turbulence in incompressible viscous fluid for very large Reynolds’ numbers. Dokl. Akad. Nauk SSSR, 30, 299303.Google Scholar
Kolmogorov, A.N. 1962. A refinement of previous hypotheses concerning the local structure of turbulence in a viscous incompressible fluid at high Reynolds number. J. Fluid Mech., 13, 8285.Google Scholar
Kraichnan, R.H. 1968. Small-scale structure of a scalar field convected by turbulence. Phys. Fluids, 11, 945953.Google Scholar
Kukuruznyak, D.A., Bulkey, S.A., Omland, K.A., Ohuchi, F.S., and Gregg, M.C. 2001. Preparation and properties of thermistor-thin-films by metal organic decomposition. Thin Solid Films, 385, 8995.Google Scholar
Kundu, P.K., and Cohen, I.M. 2004. Fluid Mechanics. 3rd ed. Academic Press.Google Scholar
Kunze, E. 1985. Near-inertial propagation in geostrophic shear. J. Phys. Oceanogr., 15, 544565.Google Scholar
Kunze, E. 1987. Limits on growing, finite-length salt fingers, a Richardson number constraint. J. Mar. Res., 45, 533556.Google Scholar
Kunze, E. 1993. Submesoscale dynamics near a seamount. Part II: The partition of energy between internal waves and geostrophy. J. Phys. Oceanogr., 23(12), 25892601.Google Scholar
Kunze, E. 2003. A review of oceanic salt-fingering theory. Prog. Oceanogr., 56, 399417.Google Scholar
Kunze, E. 2019. Biologically generated mixing in the ocean. Annu. Rev. Mar. Sci., 11.Google Scholar
Kunze, E., and Lien, R.-C. 2019. Energy sinks for lee waves in shear flow. J. Phys. Oceanogr., 49(3), 28512865.Google Scholar
Kunze, E., Dower, J.F., Beveridge, I., Dewey, R., and Bartlett, K.P. 2006b. Observations of biologically generated mixing in a coastal inlet. Science, 313, 17681770.Google Scholar
Kunze, E., Firing, E., Hummon, J.M., Chereskin, T.K., and Thurnherr, A.M. 2006a. Global abyssal mixing inferred from lowered ADCP shear and CTD strain profiles. J. Phys. Oceanogr., 36, 15531576.Google Scholar
Kunze, E., and Lueck, R. 1986. Velocity profiles in a warm-core ring. J. Phys. Oceanogr., 16(5), 991995.Google Scholar
Kunze, E., Rosenfeld, L.K., Carter, G.S., and Gregg, M.C. 2002. Internal waves in Monterey Submarine Canyon. J. Phys. Oceanogr., 32, 18901913.Google Scholar
Kunze, E., and Sanford, T.B. 1984. Observations of near-inertial waves in a front. J. Phys. Oceanogr., 14, 566581.Google Scholar
Kunze, E., Schmitt, R.W., and Toole, J.M. 1995. The energy balance in a warm-core ring’s near-inertial critical layer. J. Phys. Oceanogr., 25, 942957.Google Scholar
Kunze, E., Williams, III, A.J., and Briscoe, M.G. 1990. Observations of shear and vertical stabiliity from a neutrally buoyant float. J. Geophys. Res., 95, 18,127–18,142.Google Scholar
Kunze, E., Williams, III, A.J., and Schmitt, R.W. 1987. Optical microstructure in the thermocline staircase east of Barbados. Deep-Sea Res., 34, 16971704.Google Scholar
Lai, D.Y., Paka, V.T., Delisi, D.P., Arjannikov, A.V., and Khanaev, S.A. 2000. An inter-comparison study using electromagnetic three-component turbulent velocity probes. J. Atmos. Ocean. Tech., 17(7), 980994.Google Scholar
Lamb, K.G., Lien, R.-C., and Diamessis, P. 2019. Encyclopedia of Ocean Sciences. Elsevier Ltd. Chap. Internal solitary waves and mixing.Google Scholar
Landau, L.D., and Lifshitz, E.M. 1959. Fluid Mechanics. New York, NY: Addison-Wesley.Google Scholar
Larson, N.G., and Gregg, M.C. 1983. Turbulent dissipation and shear in thermohaline intrusions. Nature, 306, 2632.Google Scholar
Lazier, J.R.N. 1973. Temporal changes in some fresh water temperature structures. J. Phys. Oceanogr., 3, 226229.Google Scholar
Ledwell, J.R., and Bratkovich, A. 1995. A tracer study of mixing in the Santa Cruz Basin. J. Geophys. Res., 100(C10), 20,681–20,704.Google Scholar
Ledwell, J.R., Duda, T.F., Sundermeyer, M.A., and Seim, H.E. 2004. Mixing in a coastal environment: 1. A view from dye dispersion. J. Geophys. Res., 109(C10013).Google Scholar
Ledwell, J.R., He, R., Xue, Z., DiMarco, S.F., Spencer, L.J., and Chapman, P. 2016. Dispersion of a tracer in the deep Gulf of Mexico. J. Geophys. Res. Oceans, 121, 11101132.Google Scholar
Ledwell, J.R., and Hickey, B.M. 1995. Evidence for enhanced boundary mixing in the Santa Monica Basin. J. Geophys. Res., 100(C10), 20,665–20,679.Google Scholar
Ledwell, J.R., Laurent, L.C. St., Girton, J.B., and Toole, J.M. 2011. Diapycnal mixing in the Antarctic Circumpolar Current. J. Phys. Oceanogr., 41, 241246.Google Scholar
Ledwell, J.R., Montgomery, E.T., Polzin, K.L., Laurent, L. C. St., Schmitt, R.W., and Toole, J.M. 2000. Evidence for enhanced mixing over rough topography in the abyssal ocean. Nature, 403, 179182.Google Scholar
Ledwell, J.R., and Watson, A.J. 1991. The Santa Monica Basin Tracer Experiment: A study of diapycnal and isopycnal mixing. J. Geophys. Res., 96(C5), 86958718.Google Scholar
Ledwell, J.R., Watson, A.J., and Broecker, W.S. 1986. A deliberate tracer experiment in Santa Monica Basin. Nature, 323, 322324.Google Scholar
Ledwell, J.R., Watson, A.J., and Law, C.S. 1993. Evidence for slow mixing across the pycnocline from an open-ocen tracer-release experiment. Nature, 364(6439), 701703.Google Scholar
Ledwell, J.R., Watson, A.J., and Law, C.S. 1998. Mixing of a tracer in the pycnocline. J. Geophys. Res., 103(C10), 21,499–21,529.Google Scholar
Lee, C.M., Kunze, E., Sanford, T.B., Nash, J.D., Merrifield, M.A., and Holloway, P.E. 2006. Internal tides and turbulence along the 3000-m isobath of the Hawaiian Ridge. J. Phys. Oceanogr., 36, 11651183.Google Scholar
Lee, I-H., Lien, R.-C., Liu, J.T., and Chuang, W. 2009. Turbulence mixing and internal tides in Gaoping (Kaoping) submarine canyon, Taiwan. J. Mar. Syst., 76(4), 383396.Google Scholar
Legg, S., and Huijts, K.M.H. 2006. Preliminary simulations of internal waves and mixing generated by finite amplitude tidal flow over finite topography. Deep-Sea Res II, 53(1–2), 140156.Google Scholar
Legg, S., and Klymak, J. 2008. Internal hydraulic jumps and overturning generated by tidal flow over a tall steep ridge. J. Phys. Oceanogr., 38(9), 19491964.Google Scholar
Lelong, M.-P., and Dunkerton, T.J. 1998a. Inertia-gravity wave breaking in three dimensions. I. Convectively unstable waves. J. Atmos. Sci., 55(15), 24892501.Google Scholar
Lelong, M.-P., and Dunkerton, T.J. 1998b. Inertia-gravity wave breaking in three dimensions. II. Convectively stable waves. J. Atmos. Sci., 55(15), 24732488.Google Scholar
Lelong, M.-P., and Sundermeyer, M.A. 2005. Geostrophic adjustment of an isolated diapycnal mixing event and its implications for small scale lateral dispersion. J. Phys. Oceanogr., 35(12), 23522367.Google Scholar
Levine, E.R., and Lueck, R.G. 1999. Turbulence measurements from an autonomous underwater vehicle. J. Atmos. Ocean. Tech., 16, 15331544.Google Scholar
Levine, M.D. 1990. Internal waves under the Arctic ice pack during the Arctic Internal Waves Experiment: The coherence structure. J. Geophys. Res., 95, 73477357.Google Scholar
Levine, M.D. 2002. A modification of the Garrett-Munk internal wave spectrum. J. Phys. Oceanogr., 32, 31663181.Google Scholar
Levine, M.D., Paulson, C.A., and Morison, J.H. 1985. Internal waves in the Arctic Ocean: Comparison with lower-latitude observations. J. Phys. Oceanogr., 15, 800809.Google Scholar
Levine, M.D., Paulson, C.A., and Morison, J.H. 1987. Observations of internal gravity waves under the Arctic ice pack. J. Geophys. Res., 92(C1), 779782.Google Scholar
Libby, P.A. 1996. Introduction to Turbulence. Washington, D.C.: Taylor & Francis.Google Scholar
Lien, R.-C., and D’Asaro, E.A. 2002. The Kolmogorov constant for the Lagrangian velocity spectrum and structure function. Phys. Fluids, 14, 44564459.Google Scholar
Lien, R.-C., and D’Asaro, E.A. 2006. Measurement of turbulent kinetic energy dissipation rate with a Lagrangian float. J. Atmos. Ocean. Tech., 23, 964976.Google Scholar
Lien, R.-C., and Müller, P. 1992. Consistency relations for gravity and vortical modes in the ocean. Deep-Sea Res A, 39(9), 15951612.Google Scholar
Lien, R.-C., and Sanford, T.B. 2019. Small-scale potential vorticity in the upper ocean thermocline. J. Phys. Oceanogr., 49, 18451872.Google Scholar
Lighthill, J. 1979. Waves in Fluids. Cambridge University Press.Google Scholar
Lincoln, B.J., Rippeth, R.P., Lenn, Y.-D., Timmermans, M.L., Williams, W.J., and Bacon, S. 2016. Wind-driven mixing at intermediate depths in an ice-free Arctic. Geophys. Res. Lett., 43, 97499756.Google Scholar
Lindborg, E. 2006. The energy cascade in a strongly stratified fluid. J. Fluid Mech., 550, 207242.Google Scholar
Linden, P.F. 1971. Salt fingers in the presence of grid-generated turbulence. J. Fluid Mech., 49, 611624.Google Scholar
Linden, P.F. 1973. On the structure of salt fingers. Deep-Sea Res., 20, 325340.Google Scholar
Linden, P.F. 1974. Salt fingers in a steady shear flow. Geophys. Fluid Dyn., 6, 127.Google Scholar
Liu, A.K. 1988. Analysis of nonlinear internal waves in the New York Bight. J. Geophys. Res., 93, 12,317–12,329.Google Scholar
Liu, A.K., Chang, Y.S., Hsu, M.-K., and Liang, N.K. 1998. Evolution of nonlinear internal waves in the East and South China Seas. J. Geophys. Res., 103, 79958008.Google Scholar
Liu, H.T. 1995. Energetics of grid turbulence in a stably stratified fluid. J. Fluid Mech., 296, 127157.Google Scholar
Llewellyn Smith, S., and Young, W. 2002. Conversion of the barotropic tide. J. Phys. Oceanogr., 32, 15541566.Google Scholar
Lombardo, C.P., and Gregg, M.C. 1989. Similarity scaling of viscous and thermal dissipation in a convecting surface boundary layer. J. Geophys. Res., 94, 62736284.Google Scholar
Lorenz, E. 1955. Available potential energy and maintenance of the general circulation. Tellus, 7, 157167.Google Scholar
Lorke, A., and Wuest, A. 2005. Application of coherent ADCP for turbulence measurements on the bottom boundary layer. J. Atmos. Ocean. Tech., 22, 18211828.Google Scholar
Lueck, R.G. 1987. Microstructure measurements in a thermohaline staircase. Deep-Sea Res., 34, 16771688.Google Scholar
Lueck, R.G., Huang, D., Newman, D., and Box, J. 1997. Turbulence measurement with a moored instrument. J. Atmos. Ocean. Tech., 14(2), 143161.2.0.CO;2>CrossRefGoogle Scholar
Lueck, R.G., and Osborn, T.R. 1986. The dissipation of kinetic energy in a warm-core ring. J. Geophys. Res., 4, 681698.Google Scholar
Lumley, J.L. 1964. The spectrum of nearly inertial turbulence in a stably stratified fluid. J. Atmos. Sci., 21, 99102.Google Scholar
Lynn, R.J., and Reid, J.L. 1968. Characteristics and circulation of deep and abyssal waters. Deep-Sea Res, 15, 577598.Google Scholar
Mack, S.A. 1985. Two-dimensional measurements of ocean microstructure: The role of double diffusion. J. Phys. Oceanogr., 15, 15811604.Google Scholar
Mack, S.A. 1989. Towed-chain measurements of ocean microstructure. J. Phys. Oceanogr., 19, 11081129.Google Scholar
Mack, S.A., and Schoerberlein, H.C. 1993. Discriminating salt fingering from turbulence-induced microstructure: Analysis of towed temperature-conductivity chain data. J. Phys. Oceanogr., 23(9), 20732106.Google Scholar
MacKinnon, J.A., and Gregg, M.C. 2003. Mixing on the late-summer New England shelf – solibores, shear and stratification. J. Phys. Oceanogr., 33, 14761492.Google Scholar
MacKinnon, J.A., Johnston, T.M.S., and Pinkel, R. 2008. Strong transport and mixing of deep water through the Southwest Indian Ridge. Nat. Geosci., 38, 19431950.Google Scholar
MacKinnon, J.A., Laurent, L. St., and Garbato, A.C.N. 2013. Diapycnal mixing processes in the ocean interior. Chap. 7, pages 159–183 of: Siedler, G., Griffies, S.M., Gould, J., and Church, J.A. (eds.), Ocean Circulation and Climate: A 21st Century Perspective. International Geophysics, vol. 103. Amsterdam: Elsevier.Google Scholar
MacKinnon, J.A., and Winters, K.B. 2005. Subtropical catastrophe: Significant loss of low-mode tidal energy at 28.9◦. Geophys. Res. Lett., 32, doi:10.1029/2005GL023376.Google Scholar
MacKinnon, J.A., Zhao, Z., Whalen, C.B., et al. 2017. Climate process team on internal wave-driven ocean mixing. Bull. Am. Met. Soc., Nov., 2429–2454.Google Scholar
Macoun, P., and Lueck, R. 2004. Modelling the spatial response of the airfoil shear probe using different sized probes. J. Atmos. Ocean. Tech., 21, 284297.Google Scholar
Marmorino, G.O. 1987a. Observations of small-scale mixing in the thermocline. Part I: Salt fingering. J. Phys. Oceanogr., 17(9), 13391347.Google Scholar
Marmorino, G.O. 1987b. Observations of small-scale mixing processes in the seasonal thermocline: Part II: Wave breaking. J. Phys. Oceanogr., 17, 13481355.Google Scholar
Marmorino, G.O. 1991. Intrusions and diffusive interfaces in a salt-finger staircase. Deep-Sea Res. A, 38(11), 14311454.Google Scholar
Marmorino, G.O., and Caldwell, D. 1976. Heat and salt transport through a diffusive thermohaline interface. Deep-Sea Res., 23, 5967.Google Scholar
Marmorino, G.O., and Greenewalt, D. 1988. Inferring the nature of microstructure signals. J. Geophys. Res., 93, 12191225.Google Scholar
Marmorino, G.O., Rosenblum, L.J., and Trump, C.L. 1987. Finescale temperature variability: The influence of near-inertial waves. J. Geophys. Res., 92, 13,049–13,062.Google Scholar
Marshall, J., and Speer, K. 2012. Closure of the meridional overturning circulation through Southern Ocean upwelling. Nature-Geo., 5(March).Google Scholar
Marshall, J., Jamous, D., and Nilsson, J. 1999. Reconciling thermodynamic and dynamic methods of computation of water-mass transformation rates. Deep-Sea Res. I, 46, 545572.Google Scholar
Mashayek, A., and Peltier, W.R. 2011. Three-dimensionalization of the stratified mixing layer at high Reynolds number. Phys. Fluids, 23(111701).Google Scholar
Mater, B.D., and Venayagamoorthy, S.K. 2014. A unifying framework for parameterizing stably stratified shear-flow turbulence. Phys. Fluids, 26(036601).Google Scholar
Mauritzen, C., Polzin, K.L., McCartney, M.S., Millard, R.C., and West-Mack, D.E. 2002. Evidence in hydrography and density fine structure for enhanced vertical mixing over the Mid-Atlantic Ridge in the western Atlantic. J. Geophys. Res., 107(C10).Google Scholar
May, B.D., and Kelley, D.E. 1997. Effect of baroclinicity on double-diffusive interleaving. J. Phys. Oceanogr., 27, 19972008.Google Scholar
McComas, C.H., and Bretherton, F.P. 1977. Resonant interaction of oceanic internal waves. J. Geophys. Res., 82, 13971412.Google Scholar
McComas, C.H., and Müller, P. 1981a. The dynamic balance of internal waves. J. Phys. Oceanogr., 11(July), 970986.Google Scholar
McComas, C.H., and Müller, P. 1981b. Time scales of resonant interactions among oceanic internal waves. J. Phys. Oceanogr., 11(Feb.), 139147.Google Scholar
McDougall, T.J. 1981. Fluxes of properties through a series of double-diffusive interfaces with a non-linear equation of state. J. Phys. Oceanogr., 11, 12941299.Google Scholar
McDougall, T.J. 1984. The relative roles of diapycnal and isopycnal mixing on subsurface water mass conversion. J. Phys. Oceanogr., 14, 15771589.Google Scholar
McDougall, T.J. 1987a. Neutral surfaces. J. Phys. Oceanogr., 17(12), 19501964.Google Scholar
McDougall, T.J. 1987b. Thermobaricity, cabbeling and water-mass conversion. J. Geophys. Res., 92(C5), 54485464.Google Scholar
McDougall, T.J. 1988. Some implications of ocean mixing for ocean modelling. Pages 21–36 of: Nihoul, J.C.J., and Jamart, B.M. (eds.), Small-Scale Turbulence and Mixing in the Ocean. Amsterdam: Elsevier.Google Scholar
McDougall, T.J. 1991. Interfacial advection in the thermocline staircase east of Barbados. Deep-Sea Res., 38(3), 357370.Google Scholar
McDougall, T.J. 2003. Potential enthalpy: A conservative oceanic variable for evaluating heat content and heat fluxes. J. Phys. Oceanogr., 33, 945963.Google Scholar
McDougall, T.J., and Barker, P.M. 2012. Comment on ‘Buoyancy frequency profiles and internal semidiurnal tide turning depths in the oceans’ by B. King et al. J. Geophys. Res. Oceans, 119, 17.Google Scholar
McDougall, T.J., and Feistel, R. 2003. What causes the adiabatic lapse rate? Deep-Sea Res I, 50, 15231535.Google Scholar
McDougall, T.J., Feistel, R., and Pawlowicz, R. 2013. Thermodynamics of seawater. Chap. 6, pages 141–158 of: Siedler, G., Griffies, S.M., Gould, J., and Church, J.A. (eds.), Ocean Circulation and Climate. International Geophysics, vol. 103. Amsterdam: Elsevier.Google Scholar
McDougall, T.J., and Ferrari, R. 2017. Abyssal upwelling and downwelling driven by near-boundary mixing. J. Phys. Oceanogr., 47(2), 261283.Google Scholar
McDougall, T.J., and Jackett, D.R. 1988. On the helical nature of neutral surfaces. Progr. Oceanogr., 20, 153183.Google Scholar
McDougall, T.J., and Krzysik, O.A. 2015. Spiciness. J. Mar. Res., 73, 141152.Google Scholar
McDougall, T.J., and Taylor, J.R. 1984. Flux measurements across a finger interface at low values of the stability ratio. J. Mar. Res., 42, 114.Google Scholar
McKean, R.S. 1974. Interpretation of internal wave measurements in the presence of finestructure. J. Phys. Oceanogr., 4, 200213.Google Scholar
McPhee, M.G. 1992. Turbulent heat flux in the upper ocean under sea ice. J. Geophys. Res., 97, 53655379.Google Scholar
Meagher, T.B., Pederson, A.M., and Gregg, M.C. 1982. A low-noise conductivity microstructure instrument. Pages 283–290 of: Oceans’82: Conference Record, Sept. 20–22, 1982. Washington, D.C.: Marine Technology Society.Google Scholar
Merrifield, M.A., Holloway, P.E., and Johnston, T.M.S. 2001. The generation of internal tides at the Hawaiian Ridge. Geophys. Res. Lett., 28(4), 559562.Google Scholar
Merryfield, W.J. 2000. Origin of thermohaline staircases. J. Phys. Oceanogr., 30, 10461068.Google Scholar
Merryfield, W.J., and Grinder, M. 1999. Salt fingering fluxes from numerical simulations. Unpublished manuscript.Google Scholar
Meyer, A., Polzin, K.L., Sloyan, B.M., and Phillips, H.E. 2015a. Internal waves and mixing near the Kerguelen Plateau. J. Phys. Oceanogr., 46(2), 417437.Google Scholar
Meyer, A., Sloyan, B.M., Polzin, K.L., Phillips, H.E., and Bindoff, N.L. 2015b. Mixing variability in the Southern Ocean. J. Phys. Oceanogr., 45(4), 966987.Google Scholar
Miles, J. 1961. On the stability of heterogenous shear flows. J. Fluid Mech., 10, 496508.Google Scholar
Miller, J.B., Gregg, M.C., Miller, V.W., and Welsh, G.L. 1989. Vibration of tethered microstructure profilers. J. Atmos. Ocean. Technol., 6, 980984.Google Scholar
Millero, F., Chen, C.T., Bradshaw, A., and Schleicher, K. 1980. A new high pressure equation of state for seawater. Deep-Sea Res., 27A, 255264.Google Scholar
Miyake, Y., and Koizumi, M. 1948. The measurement of the viscosity coefficient of sea water. J. Mar. Res., 7, 6366.Google Scholar
Monin, A.S., and Yaglom, A.M. 1971. Statistical Fluid Mechanics: Mechanics of Turbulence. Vol. 1. Cambridge, MA: The MIT Press.Google Scholar
Monin, A.S., and Yaglom, A.M. 1975. Statistical Fluid Mechanics: Mechanics of Turbulence. Vol. 2. Cambridge, MA: The MIT Press.Google Scholar
Montgomery, R.B. 1938. Circulation in the upper layers of the southern North Atlantic, deduced with the use of isentropic analysis. Pap. Phys. Oceanogr. Meteor, 6(2), 55.Google Scholar
Montroll, E.W., and Shlesinger, M.F. 1982. On 1/f and other distributions with long tails. Proc. Natl. Acad. Sci. USA, 79, 33803383.Google Scholar
Morison, J.H., Long, C.E., and Levine, M.D. 1985. Internal wave dissipation under sea ice. J. Geophys. Res., 90, 11,959–11,966.Google Scholar
Morrison, A.T., Billings, J.D., and Doherty, K.W. 2000. The McLane moored profiler: An autonomous platform for oceanographic measurements. In: Oceans 2000 MTS/IEEE Conf. & Exhib.Google Scholar
Moum, J.N. 1990. Profiler measurements of vertical velocity microstructure in the ocean. J. Atmos. Ocean. Tech., 7, 323333.Google Scholar
Moum, J.N. 1996. Energy-containing scales of turbulence in the ocean thermocline. J. Geophys. Res., 101(C6), 14,095–14,109.Google Scholar
Moum, J.N. 2015. Ocean speed and turbulence measurements using pitot–static tubes on moorings. J. Atmos. Ocean. Tech., 32, 14001413.Google Scholar
Moum, J.N., and Nash, J.D. 2009. Mixing measurements on an equatorial ocean mooring. J. Atmos. Ocean. Tech., 26, 317336.Google Scholar
Moum, J.N., Gregg, M.C., Lien, R.C., and Carr, M.E. 1995. Comparison of turbulent kinetic energy dissipation rate estimates from two ocean microstructure profilers. J. Atmos. Ocean. Tech., 12, 346366.Google Scholar
Moum, J.N., Caldwell, D.R., Nash, J.D., and Gunderson, G.D. 2002. Observations of boundary mixing over the continental slope. J. Phys. Oceanogr., 32, 21132130.Google Scholar
Moum, J.N., Farmer, D.M., Smyth, W.D., Armi, L., and Vagle, S. 2003. Structure and generation of turbulence at interfaces strained by internal solitary waves propagating shoreward over the continental shelf. J. Phys. Oceanogr., 33(10), 20932112.Google Scholar
Moum, J.N., Farmer, D.M., Shroyer, E.L., Smyth, W.D., and Armi, L. 2007a. Dissipative losses in nonlinear internal waves propagating across the continental shelf. J. Phys. Oceanogr., 37(7), 19891995.Google Scholar
Moum, J.N., Klymak, J.M., Nash, J.D., Perlin, A., and Smyth, W.D. 2007b. Energy transport by nonlinear internal waves. J. Phys. Oceanogr., 37(7), 19681988.Google Scholar
Müller, P., and Siedler, G. 1976. Consistency relations for internal waves. Deep-Sea Res, 23, 613628.Google Scholar
Müller, P., Olbers, D.J., and Willebrand, J. 1978. The IWEX spectrum. J. Geophys. Res., 83, 479500.Google Scholar
Munk, W.H. 1966. Abyssal recipes. Deep-Sea Res., 13, 707730.Google Scholar
Munk, W.H. 1981. Internal waves and small-scale processes. Pages 264–291 of: Warren, B.A., and Wunsch, C. (eds.), Evolution of Physical Oceanography: Scientific Surveys in Honor of Henry Stommel. Cambridge, MA: MIT Press.Google Scholar
Munk, W.H., and Wunsch, C. 1998. Abyssal recipes II: Energetics of tidal and wind mixing. Deep-Sea Res. I, 45(12), 19772010.Google Scholar
Nagai, T., Tandon, A., Kunze, E., and Mahadevan, A. 2015. Spontaneous generation of near-inertial waves by the Kuroshio Front. J. Phys. Oceanogr., 45, 23812406.Google Scholar
Nagasawa, M., Hibiya, T., Niwa, Y., Watanabe, M., Isoda, Y., Takagi, S., and Kamei, Y. 2002. Distribution of fine-scale shear in the deep waters of the North Pacific obtained using expendable current profilers. J. Geophys. Res., 107(C2).Google Scholar
Nandi, P., Holbrook, W.S., Pearse, S., Páramo, P., and Schmitt, R.W. 2004. Seismic reflection profiling of water mass boundaries in the Norwegian Sea. Geophys. Res. Lett., 31(L23311).Google Scholar
Nash, J.D., Alford, M.H., Kunze, E., Martini, K., and Kelly, S. 2007. Hotspots of deep ocean mixing on the Oregon continental slope. Geophys. Res. Lett., 34(L01605).Google Scholar
Nash, J.D., and Moum, J.N. 1999. Estimating salinity variance dissipation rate from conductivity microstructure measurements. J. Atmos. Ocean. Tech., 16, 263274.Google Scholar
Nash, J.D., and Moum, J.N. 2002. Microstructure estimates of turbulent salinity flux and the dissipation spectrum of salinity. J. Phys. Oceanogr., 32(8), 23122334.Google Scholar
Nash, J.D., Caldwell, D.R., Zeiman, M.J., and Moum, J.N. 1999. A thermocouple probe for high speed temperature measurement in the ocean. J. Atmos. Ocean. Tech., 16, 14741482.Google Scholar
Nash, J.D., Kunze, E., Lee, C.M., and Sanford, T.B. 2006. Structure of the baroclinic tide generated at Kaena Ridge, Hawaii. J. Phys. Oceanogr., 36, 11231135.Google Scholar
Nasmyth, P.W. 1970. Oceanic Turbulence. Ph.D. thesis, University of British Columbia, Vancouver, Canada.Google Scholar
Naveira Garabato, A.C., Oliver, K.I.C., Watson, A.J., and Messias, M.-J. 2004. Turbulent diapycnal mixing in the Nordic seas. J. Geophys. Res., 109(C12010).Google Scholar
Neal, V.T., Neshbya, S., and Denner, W. 1969. Thermal stratification in the Arctic. Science, 166(3903), 373374.Google Scholar
Newman, F.C. 1976. Temperature steps in Lake Kivu: A bottom heated saline lake. J. Phys. Oceanogr., 6, 157163.Google Scholar
Nikurashin, M., and Ferrari, R. 2010a. Radiation and dissipation of internal waves generated by geostrophic flows impinging on small-scale topography: Application to the Southern Ocean. J. Phys. Oceanogr., 40(5), 20252042.Google Scholar
Nikurashin, M., and Ferrari, R. 2010b. Radiation and dissipation of internal waves generated by geostrophic motions impinging on small-scale topography: Theory. J. Phys. Oceanogr., 40(5), 10551074.Google Scholar
Nikurashin, M., and Ferrari, R. 2011. Global energy conversion rate from geostrophic flows into internal lee waves in the deep ocean. Geophys. Res. Lett., 38(L08610).Google Scholar
Nimmo Smith, W.A.M. 2008. A submersible three-dimensional particle tracking velocimetry system for flow visualization in the coastal ocean. Limnol. Oceanogr., 6, 96104.Google Scholar
Nimmo Smith, W.A.M., Katz, J., and Osborn, T.R. 2005. On the structure of turbulence in the bottom boundary layer of the coastal ocean. J. Phys. Oceanogr., 35(1), 7293.Google Scholar
Niwa, Y., and Hibiya, T. 2014. Generation of baroclinic tide energy in a global three-dimensional numerical model with different spatial grid resolutions. Ocean Modelling, 80, 5973.Google Scholar
Oakey, N.S. 1977. OCTUPROBE III: An Instrument to Measure Oceanic Turbulence and Microstructure. Rept. Series BI-R-77-3. Bedford Inst. Oceanogr., Dartmouth, N.S., Canada.Google Scholar
Oakey, N.S. 1982. Determination of the rate of dissipation of turbulent energy from simultaneous temperature and velocity shear microstructure measurements. J. Phys. Oceanogr., 12, 256271.Google Scholar
Oakey, N.S., and Greenan, B.J.W. 2004. Mixing in a coastal environment: 1. A view from microstructure measurements. J. Geophys. Res., 109(C10014).Google Scholar
Oort, A.H., Anderson, L.A., and Peixoto, J.P. 1994. Estimates of the energy cycle of the oceans. J. Geophys. Res., 99(C4), 76657688.Google Scholar
Orlanski, I., and Bryan, K. 1969. Formation of the thermocline step structure of large-amplitude internal gravity waves. J. Geophys. Res., 74, 69756983.Google Scholar
Orr, M.H. 1981. Remote, acoustic detection of zooplankton response to fluid processes, oceanographic instrumentation, and predators. Can. J. Fish. Aquat. Sci., 38, 10961105.Google Scholar
Orr, M.H., and Mignerey, P.C. 2003. Nonlinear internal waves in the South China Sea: Observation of the conversion of depression internal waves to elevation internal waves. J. Geophys. Res., 108(3064).Google Scholar
Osborn, T.R. 1974. Vertical profiling of velocity microstructure. J. Phys. Oceanogr., 4, 109115.Google Scholar
Osborn, T.R. 1980. Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr., 10, 8389.Google Scholar
Osborn, T.R., and Cox, C.S. 1972. Oceanic fine structure. Geophys. Fluid Dyn., 3, 321345.Google Scholar
Osborn, T.R., and Crawford, W.R. 1980. An airfoil probe for measuring turbulent velocity fluctuations in water. Pages 369–386 of: Dobson, F., Hasse, L., and Davis, R. (eds.), Air-Sea Interactions: Instruments and Methods. New York: Plenum.Google Scholar
Osborn, T.R., and Lueck, R.G. 1985. Turbulence measurement with a submarine. J. Phys. Oceanogr., 15, 15021520.Google Scholar
Ozmidov, R.V. 1965. On the turbulent exchange in a stably stratified ocean. Izv., Atmos. Oceanic Phys, 1, 853860.Google Scholar
Padman, L., and Dillon, T.M. 1987. Vertical heat fluxes through the Beaufort Sea thermohaline staircases. J. Geophys. Res., 92(C10), 10,799–10,806.Google Scholar
Padman, L., and Dillon, T.M. 1991. Turbulent mixing near the Yermak Plateau during the Coordinated Eastern Arctic Experiment. J. Geophys. Res., 96, 47694782.Google Scholar
Palmer, M.R., Stephenson, G.R., Inall, M.E., Balfour, C., Düsterhus, A., and Green, J.A.M. 2015. Turbulence and mixing by internal waves in the Celtic Sea determined from ocean glider microstructure measurements. J. Mar. Syst, 144, 5769.Google Scholar
Panchev, S., and Kesich, D. 1969. Energy spectrum of isotropic turbulence at large wavenumbers. Comptes rendus de l’Academie bulgare des Sciences, 22, 627630.Google Scholar
Pawlowicz, R.T., McDougall, T., Feistel, R., and Tailleux, R. 2012. An historical perspective on the development of the development of the Thermodynamic Equation of Seawater–2010. Ocean Sci., 8, 161174.Google Scholar
Pederson, A.M. 1973. A small in-situ conductivity instrument. Pages 68–75 of: IEEE Int. Conf. Eng. in the Ocean.Google Scholar
Pederson, A.M., and Gregg, M.C. 1979. Development of a small in-situ conductivity instrument. IEEE J. Ocean Engr., OE-4, 6975.Google Scholar
Perlin, A., and Moum, J.N. 2012. Comparison of thermal variance dissipation rates from moored and profiling instruments at the equator. J. Atmos. Ocean. Tech., 29, 13471362.Google Scholar
Peters, H., Gregg, M.C., and Sanford, T.B. 1995. Detail and scaling of turbulent overturns in the Pacific Equatorial Undercurrent. J. Geophys. Res., 100, 18,349–18,368.Google Scholar
Phillips, O.M. 1966. Dynamics of the Upper Ocean. Cambridge University Press.Google Scholar
Piera, J., Roget, E., and Catalan, J. 2002. Turbulent patch identification in microstructure profiles: A method based on wavelet denoising and Thorpe displacement analysis. J. Atmos. Ocean. Tech., 19, 13901402.Google Scholar
Pingree, R.D. 1972. Mixing in the deep stratified ocean. Deep-Sea Res, 19, 549561.Google Scholar
Pinkel, R. 1981. On the use of Doppler sonar for internal wave measurements. Deep-Sea Res, 28A(3), 269289.Google Scholar
Pinkel, R. 2000. Internal solitary waves in the warm pool of the western equatorial Pacific. J. Phys. Oceanogr., 30, 29062926.Google Scholar
Pinkel, R. 2008a. Advection, phase distortion, and the frequency spectrum of finescale fields in the sea. J. Phys. Oceanogr., 38, 291313.Google Scholar
Pinkel, R. 2008b. The wavenumber-frequency spectrum of vortical and internal-wave shear in the western Arctic Ocean. J. Phys. Oceanogr., 38(Feb.), 277290.Google Scholar
Pinkel, R. 2014. Vortical and internal wave shear and strain. J. Phys. Oceanogr., 44, 20702092.Google Scholar
Pinkel, R. 2020. The Poisson link between internal wave and dissipation scales in the thermocline. Part I. Probability density functions and the Poisson modeling of vertical strain. J. Phys. Oceanogr., submitted.Google Scholar
Pinkel, R., and Anderson, S. 1992. Toward a statistical description of finescale strain in the thermocline. J. Phys. Oceanogr., 22, 773795.Google Scholar
Pinkel, R., and Anderson, S. 1997. Shear, strain, and Richardson number variations in the thermocline. Part I: Statistical description. J. Phys. Oceanogr., 27(2), 264281.Google Scholar
Pinkel, R., Golden, M.A., Smith, J.A., Sun, O.M., Aja, A.A., Bui, M.N., and Hughen, T. 2011. The wirewalker: A vertically profiling instrument carrier powered by ocean waves. J. Atmos. Ocean. Tech., 28(3), 426435.Google Scholar
Pinkel, R., Rainville, L., and Klymak, J. 2012. Semidiurnal baroclinic wave momentum fluxes at Kaena Ridge, Hawaii. J. Phys. Oceanogr., 42(8), 12491269.Google Scholar
Pollard, R.T., and Millard, R.C. 1970. Comparison between observed and simulated wind-generated intertial currents. Deep-Sea Res., 17, 813821.Google Scholar
Pollman, F., Eden, C., and Olbers, D. 2017. Evaluating the global internal wave model IDEMIX using finestructure methods. J. Phys. Oceanogr., 47(9), 22672289.Google Scholar
Polyakov, I.V., Padman, L., Lenn, Y.-D., Pnyushkov, A., Rember, R., and Ivanov, V.V. 2019. Eastern Arctic Ocean diapycnal heat fluxes through large double-diffusive steps. J. Phys. Oceanogr., 49(1), 227246.Google Scholar
Polzin, K.L. 1996. Statistics of Richardson number: Mixing models and finestructure. J. Phys. Oceanogr., 26, 14091425.Google Scholar
Polzin, K.L., and Ferrari, R. 2004. Isopycnal dispersion in NATRE. J. Phys. Oceanogr., 34, 247257.Google Scholar
Polzin, K.L., Garabato, A.C.N., Hussen, T.N., Sloyan, B.M., and Waterman, S. 2014. Finescale parameterizations of turbulent dissipation. JGR, 119, 129.Google Scholar
Polzin, K.L., Kunze, E., Hummon, J., and Firing, E. 2002. The finescale response of lowered ADCP velocity profiles. J. Atmos. Ocean. Tech., 19(2), 205224.Google Scholar
Polzin, K.L., Kunze, E., Toole, J.M., and Schmitt, R.W. 2003. The partition of finescale energy into internal waves and subinertial motions. J. Phys. Oceanogr., 33, 234248.Google Scholar
Polzin, K.L., Toole, J.M., Ledwell, J.R., and Schmitt, R.W. 1997. Spatial variability of turbulent mixing in the abyssal ocean. Science, 276(5309), 9396.Google Scholar
Polzin, K.L., and Lvov, Y.L. 2011. Toward regional characterizations of the oceanic internal wavefield. Rev. Geophys., 49(RG4003).Google Scholar
Polzin, K.L., Speer, K.G., Toole, J.M., and Schmitt, R.W. 1996. Intense mixing of Antarctic Bottom Water in the equatorial Atlantic Ocean. Nature, 380, 5457.Google Scholar
Polzin, K.L., Toole, J.M., and Schmidt, R.W. 1995. Finescale parameterization of turbulent dissipation. J. Phys. Oceanogr., 25, 306328.Google Scholar
Pope, S.B. 2000. Turbulent Flows. Cambridge, UK: Cambridge University Press.Google Scholar
Proni, J.R., and Apel, J.R. 1975. On the use of high-frequency acoustics for the study of internal waves and microstructure. J. Geophys. Res., 80, 11471151.Google Scholar
Pytkowicz, R.M. 1963. Gravity and the properties of sea water. Pages 286–287 of: The Sea, vol. 8. Wiley-Interscience.Google Scholar
Radko, T. 2014. Applicability and failure of the flux-gradient laws in double-diffusive convection. J. Fluid Mech., 750, 3372.Google Scholar
Rainville, L., and Pinkel, R. 2004. Observations of energetic high-wavenumber internal waves in the Kuroshio. J. Phys. Oceanogr., 34, 14951505.Google Scholar
Rainville, L., and Winsor, P. 2008. Mixing across the Arctic Ocean: Microstructure observations during the Beringia 2005 Expedition. Geophys. Res. Lett., 35(8).Google Scholar
Rainville, L., Johnston, T.M.S., Carter, G.S., Merrifield, M.A., Pinkel, R., Worcester, P.F., and Dushaw, B.D. 2010. Interference pattern and propagation of the M2 internal tide south of the Hawaiian Ridge. J. Phys. Oceanogr., 40, 311325.Google Scholar
Ray, G.T, and Mitchum, R.D. 1996. Surface generation of internal tides generated near Hawaii. Geophys. Res. Lett., 23, 21012104.Google Scholar
Reynolds, O. 1895. On the dynamical theory of incompressible viscous fluids and the determination of the criterion. Proc. Roy. Soc. Lond. A, A186, 123164.Google Scholar
Rice, J.A. 1988. Mathematical Statistics and Data Analysis. Wadsworth and Brooks/Cole.Google Scholar
Riley, J.J., and Lelong, M.-P. 2000. Fluid motions in the presence of strong stable stratification. Annu. Rev. Fluid Mech., 32, 613657.Google Scholar
Riley, J.J., and Lindborg, E. 2008. Stratified turbulence: A possible interpretation of some geophysical turbulence measurements. J. Atmos. Sci., 65(7), 24162424.Google Scholar
Riley, J.J., Metcalfe, R.W., and Weissman, M.A. 1981. Direct numerical simulations of homogenous turbulence in density-stratified fluids. Pages 79–112 of: West, B.J. (ed.), Proc. AIP Conf. Nonlinear Properties of Internal Waves. American Inst. Physics.Google Scholar
Rimac, A., and von Storch, J.-S. 2016. The total energy flux leaving the ocean’s mixed layer. J. Phys. Oceanogr., 46(6), 18851900.Google Scholar
Roden, G.I. 1964. Shallow temperature inversions in the Pacific Ocean. J. Geophys. Res., 61, 255263.Google Scholar
Roemmich, D., Hautala, S., and Rudnick, D. 1996. Northward abyssal transport through the Samoan Passage and adjacent regions. J. Geophys. Res., 101(C6), 14,039–14,066.Google Scholar
Rohr, J.J., Helland, K.N., Itsweire, E.C., and Atta, C.W. Van. 1987. Turbulence in a stably stratified shear flow: A progress report. In: Turbulent Shear Flows, vol. 5. New York, NY: Springer-Verlag.Google Scholar
Rosenblum, L.J., and Marmorino, G. 1990. Statistics of mixing patches observed in the Sargasso Sea. J. Geophys. Res., 95, 53495357.Google Scholar
Ross, C.K. 1984. Temperature-salinity characteristics of the ‘overflow’ water in Denmark Strait during ‘Overflow 73’. Rapp. P.-V. Reun. Cons. Int. Explor. Mer, 185, 111119.Google Scholar
Rossby, H.T. 1969. A vertical profile of currents near Plantagenet Bank. Deep-Sea Res., 16, 377385.Google Scholar
Ruddick, B. 1983. A practical indicator of the stability of the water column to double-diffusive activity. Deep-Sea Res., 3, 11051107.Google Scholar
Ruddick, B., Walsh, D., and Oakey, N. 1997. Variations in apparent mixing efficiency in the North Atlantic central water. J. Phys. Oceanogr., 27(12), 25892605.Google Scholar
Rudnick, D. 1997. Direct velocity measurements in the Samoan Passage. J. Geophys. Res., 102(C2), 32933302.Google Scholar
Rudnick, D.L., Boyd, T.J., Brainerd, R.E., Carter, G.S., Egbert, G.D., Gregg, M.C., Holloway, P.E., Klymak, J.M., Kunze, E., Lee, C.M., Levine, M.D., Luther, D.S., Martin, J.P., Merrifield, M.A., Moum, J.N., Nash, J.D., Pinkel, R., Rainville, L., and Sanford, T.B. 2003. From tides to mixing along the Hawaiian Ridge. Science, 301, 355357.Google Scholar
Salehipour, H., Peltier, W.R., and Mashayek, A. 2015. Turbulent diapycnal mixing in stratified shear flows: The influence of Prandtl number on mixing efficiency and transition at high Reynolds number. J. Fluid Mech., 773, 178223.Google Scholar
Sanchez, X., Roget, E., Planella, J., and Forcat, F. 2011. Small-scale spectrum of a scalar field in water: The Batchelor and Kraichnan models. J. Phys. Oceanogr., 41, 21552167.Google Scholar
Sanford, T.B. 1975. Observations of the vertical structure of internal waves. J. Geophys. Res., 80(27), 38613871.Google Scholar
Sanford, T.B., Carlson, J.A., Dunlap, J.H., Prater, M.D., and Lien, R.-C. 1999. An electromagnetic vorticity and velocity sensor for observing finescale kinetic fluctuations in the ocean. J. Atmos. Ocean. Tech., 16, 16471667.Google Scholar
Sanford, T.B., Drever, R.G., and Dunlap, J.H. 1978. A velocity profiler based on the principles of geomagnetic induction. Deep-Sea Res, 25, 183210.Google Scholar
Sanford, T.B., Drever, R.G., and Dunlap, J.H. 1985. An acoustic Doppler and electromagnetic profiler. J. Atmos. Ocean. Tech., 2(6), 110124.Google Scholar
Sanford, T.B., Drever, R.G., Dunlap, J.H., and D’Asaro, E.A. 1982. Design, Operation and Performance of an Expendable Temperature and Velocity Profiler. Tech. Report 8110. Applied Physics Laboratory, University of Washington, 1013 NE 40th St., Seattle, WA 98105-6698.Google Scholar
Sanford, T.B., Dunlap, J.H., Carlson, J.A., Webb, D.C., and Girton, J.B. 2005. Autonomous velocity and density profiler: EM-APEX. Pages 152–156 of: White, J.R., and Anderson, S. (eds.), Proceedings of the IEEE/OES/CMTC Eighth Working Conference on Current Measurement Technology. The Printing House, 445 Hoes Lane, Piscataway, NJ, 08854: IEEE.Google Scholar
Sanford, T.B., and Lien, R.-C. 1999. Turbulent properties in a homogenous tidal bottom boundary layer. J. Geophys. Res., 104(C1), 12451257.Google Scholar
Sanford, T.B., Price, J.F., and Girton, J.B. 2011. Upper-ocean response to Hurricane Frances (2004) observed by profiling EM-APEX floats. J. Phys. Oceanogr., 41, 10411055.Google Scholar
Schanze, J.J., and Schmitt, R.W. 2013. Estimates of cabbeling in the global ocean. J. Phys. Oceanogr., 43(4), 698705.Google Scholar
Schmitt, R.W. 1979. Flux measurements on salt fingers at an interface. J. Mar. Res., 37, 419436.Google Scholar
Schmitt, R.W. 1981. Form of the temperature-salinity relationship in the central water: Evidence for double-diffusive mixing. J. Phys. Oceanogr., 11(7), 10151026.Google Scholar
Schmitt, R.W., Ledwell, J.R., Montromery, E.T., Polzin, K.L., and Toole, J.M. 2005. Enhanced diapycnal mixing by salt fingers in the thermocline of the tropical Atlantic. Science, 308(5722), 685688.Google Scholar
Schmitt, R.W., Perkins, H., Boyd, J.D., and Stalcup, M.C. 1987. C-SALT: An investigation of the thermohaline staircase in the western tropical North Atlantic. Deep-Sea Res., 34(10), 16551665.Google Scholar
Schmitt, R.W., Toole, J.M., Koehler, R.L., Mellinger, E.C., and Doherty, K.W. 1988. The development of a fine- and microstructure profiler. J. Atmos. Ocean. Tech., 5, 484500.Google Scholar
Seim, H.E. 1999. Acoustic backscatter from salinity microstructure. J. Atmos. Ocean. Tech., 16, 14911498.Google Scholar
Seim, H.E., and Gregg, M.C. 1994. Detailed observations of a naturally occurring shear instability. J. Geophys. Res., 99, 10,049–10,073.Google Scholar
Seim, H.E., Gregg, M.C., and Miyamoto, R.T. 1995. Acoustic backscatter from turbulent microstructure. J. Atmos. Ocean. Tech., 12, 367380.Google Scholar
Shaw, W.J., and Stanton, T.P. 2014. Dynamic and double-diffusive instabilities in a weak pycnocline. Part I: Observations of heat flux and diffusivity in the vicinity of Maud Rise, Weddell Sea. J. Phys. Oceanogr., 44, 19731991.Google Scholar
Shay, T.J., and Gregg, M.C. 1986. Convectively driven turbulent mixing in the upper ocean. J. Phys. Oceanogr., 16, 17771798.Google Scholar
Shcherbina, A.Y., Gregg, M.C., Alford, M.H., and Harcourt, R.R. 2009. Characterizing thermohaline intrusions in the North Pacific subtropical frontal zone. J. Phys. Oceanogr., 39(11), 27352756.Google Scholar
Shcherbina, A.Y., Gregg, M.C., Alford, M.H., and Harcourt, R.R. 2010. Three-dimensional structure and temporal evolution of submesoscale thermohaline intrusions in the north Pacific subtropical frontal zone. J. Phys. Oceanogr., 40(8), 16691689.Google Scholar
Shen, C.Y. 1993. Heat-salt finger fluxes across a density interface. Phys. Fluids A, 5, 2633– 2643.Google Scholar
Shen, C.Y. 1995. Equilibrium salt-fingering convection. Phys. Fluids, 7, 706717.Google Scholar
Sheen, K.L., Brearley, J.A., Garabato, A.C.N., Smeed, D.A., Waterman, S., Ledwell, J.R., Meredith, M.P., Laurent, L. St., Thurnherr, A.M., Toole, J.M., and Watson, A.J. 2013. Rates and mechanisms of turbulent dissipation and mixing in the Southern Ocean: Results from the Diapycnal and Isopycnal Mixing Experiment in the Southern Ocean (DIMES). J. Geophys. Res., 118, 27742792.Google Scholar
Sherman, J.T., and Davis, R.E. 1995. Observations of temperature microstructure in NATRE. J. Phys. Oceanogr., 25, 19131929.Google Scholar
Sherman, J.T., and Pinkel, R. 1991. Estimates of the vertical wavenumber-frequency spectra of vertical shear and strain. J. Phys. Oceanogr., 21, 292303.Google Scholar
Shirtcliffe, T.G.L. 1967. Thermosolutal convection: Observation of an overstable mode. J. Fluid Mech., 213, 480490.Google Scholar
Shroyer, E.L., Moum, J.N., and Nash, J.D. 2010. Energy transformations and dissipation of nonlinear internal waves over New Jersey’s continental shelf. Nonlin. Process Geophys., 17, 345360.Google Scholar
Simmons, H.L. 2008. Spectral modification and geographic redistribution of the semidiurnal internal tide. Ocean Modell., 21, 126138.Google Scholar
Simmons, H.L., and Alford, M.H. 2012. Simulating the long range swell of internal waves generated by ocean storms. Oceanography, 25(2), 126138.Google Scholar
Simpson, J.H., Howe, M.R., Morris, N.C.G., and Stratford, J. 1979. Velocity shear in the steps below the Mediterranean outflow. Deep-Sea Res., 26A, 13811386.Google Scholar
Sloyan, B.M. 2005. Spatial variability of mixing in the Southern Ocean. Geophys. Res. Lett., 32(L18603).Google Scholar
Smith, P.C. 1976. Baroclinic instability in the Denmark Strait overflow. J. Phys. Oceanogr., 6, 355371.Google Scholar
Smith, S.A., Fritts, D.C., and VanZandt, T.E. 1987. Evidence for a saturated spectrum of atmospheric gravity waves. J. Atmos. Sci., 44(10), 14041410.Google Scholar
Smyth, W.D. 1999. Dissipation-range geometry and scalar spectra in sheared stratified turbulence. J. Fluid Mech., 401, 209242.Google Scholar
Smyth, W.D., and Kimura, S. 2011. Mixing in a moderately sheared salt-fingering layer. J. Phys. Oceanogr., 41(7), 13641384.Google Scholar
Smyth, W.D., and Moum, J.N. 2000. Anisotropy of turbulence in stably stratified layers. Phys. Fluids, 12(6), 13431362.Google Scholar
Smyth, W.D., Moum, J.N., and Caldwell, D.R. 2001. The efficiency of mixing in turbulent patches: Inferences from direct simulations and microstructure observations. J. Phys. Oceanogr., 31, 19691992.Google Scholar
Smyth, W.D., and Moum, J.N. 2012. Ocean mixing by Kelvin–Helmholtz instability. Oceanography, 25(2), 140149.Google Scholar
Smyth, W.D., Nash, J.D., and Moum, J.N. 2005. Differential diffusion in breaking Kelvin– Helmholtz billows. J. Phys. Oceanogr., 35(6), 10041020.Google Scholar
Smyth, W.D., and Thorpe, S.A. 2012. Glider measurements of overturning in a Kelvin– Helmholtz billow train. J. Mar. Res., 70, 119140.Google Scholar
Smyth, W.D., and Winters, K.B. 2003. Turbulence and mixing in Holmboe waves. J. Phys. Oceanogr., 33, 694711.Google Scholar
Smyth, W.D., Carpenter, J.R., and Lawrence, G. 2007. Mixing in symmetric Holmboe waves. J. Phys. Oceanogr., 37, 15661583.Google Scholar
Solomon, H. 1971. On the representation of isentropic mixing in ocean circulation models. J. Phys. Oceanogr., 1, 233234.Google Scholar
Soloviev, A., Lukas, R., Hacker, P., Schoerberlein, H., Baker, M., and Arjannikov, A. 1999. A near-surface microstructure sensor system used during TOGA COARE. Part II: Turbulence measurements. J. Atmos. Ocean. Tech., 16(11), 15981618.Google Scholar
Sommer, T., Carpenter, J.R., Schmid, M., Lueck, R.G., Schurter, M., and Wüest, A. 2013a. Interface structure and flux laws in a natural double-diffusive layering. J. Geophys. Res., 118, 60926106.Google Scholar
Sommer, T., Carpenter, J.R., Schmid, M., Lueck, R.G., and Wüest, A. 2013b. Revisiting microstructure sensor responses with implications for double-diffusive fluxes. J. Atmos. Ocean. Tech., 30, 19071923.Google Scholar
Sommer, T., Schmid, M., and Wüest, A. 2019. The role of double diffusion for heat and salt balances in Lake Kivu. Limnol. Oceanogr., 64(2019), 650660.Google Scholar
Spilhaus, A.F. 1939. A detailed study of the surface layers of the ocean in the neighborhood of the Gulf Stream with the aid of rapid measuring hydrographic instruments. J. Mar. Res., 3, 5175.Google Scholar
Sreenivasan, K. 1995. On the universality of the Kolmogorov constant. Phys. Fluids, 7, 27782784.Google Scholar
Sreenivasan, K. 1996. The passive scalar spectrum and the Obukhov-Corrsin constant. Phys. Fluids, 8(189), 189196.Google Scholar
St. Laurent, L. 2008. Turbulent dissipation on the margins of the South China Sea. Geophys. Res. Lett., 35(L23615).Google Scholar
St. Laurent, L., Garabato, A.C.N., Ledwell, J.R., Thurnherr, A.M., Toole, J.M., and Watson, A.J. 2012. Turbulence and diapycnal mixing in Drake Passage. J. Phys. Oceanogr., 42(12), 21432152.Google Scholar
St. Laurent, L., and Garrett, C. 2002. The role of internal tides in mixing the deep ocean. J. Phys. Oceanogr., 32(10), 28822899.Google Scholar
St. Laurent, L., and Schmitt, R.W. 1999. The contribution of salt fingers to vertical mixing in the North Atlantic Tracer Release Experiment. J. Phys. Oceanogr., 29, 14041424.Google Scholar
St. Laurent, L., Stringer, S., Garrett, C., and Perrault-Joncas, D. 2003. The generation of internal tides at abrupt topography. Deep-Sea Res. I, 50, 9871003.Google Scholar
St. Laurent, L., Toole, J.M., and Schmitt, R.W. 2001. Buoyancy forcing by turbulence above rough topography in the abyssal Brazil Basin. JPO, 31, 34763485.Google Scholar
Stacey, M.T., Monismith, S.G., and Burau, J.R. 1999. Measurement of Reynolds stress profiles in unstratified profiles. J. Geophys. Res., 104, 10,933–10,949.Google Scholar
Steele, E., Nimmo-Smith, A., Vlasenko, A., Vlasenko, V., and Hosegood, P. 2013. Examination of turbulence structures in the bottom boundary layer of the coastal ocean by submersible 3D-PTV. In: 10th International Symposium of Particle Image Velocimetry – PIV13, Delft, The Netherlands, July 1–3, 2013.Google Scholar
Steffen, E.L., and D’Asaro, E.A. 2002. Deep-convection in the Labrador Sea observed by Lagrangian floats. J. Phys. Oceanogr., 32(2), 475492.Google Scholar
Stern, M.E. 1960. The ‘salt-fountain’ and thermohaline convection. Tellus, 12, 172175.Google Scholar
Stern, M.E. 1967. Lateral mixing of water masses. Deep-Sea Res., 14, 747753.Google Scholar
Stern, M.E. 1968. T-S gradients on the micro scale. Deep-Sea Res., 15, 245250.Google Scholar
Stern, M.E. 1969. Collective instability of salt fingers. J. Fluid Mech., 35, 209218.Google Scholar
Stern, M.E. 1975. Ocean Circulation Physics. New York: Academic.Google Scholar
Stern, M.E., Radko, T., and Simeonov, J. 2001. Salt fingers in an unbounded thermocline. J. Mar. Res., 59, 355390.Google Scholar
Stewart, R.W. 1959. The problem of diffusion in a stratified fluid. Annu. Rev. Geophys., 6, 303311.Google Scholar
Stillinger, D.C., Helland, K.N., and Atta, C.W. Van. 1983. Experiments on the transition of homogeneous turbulence to internal waves in a stratified fluid. J. Fl., 131, 91122.Google Scholar
Stommel, H. 1962. On the cause of the temperature-salinity curve in the ocean. Nat. Acad. Sci., 48, 764766.Google Scholar
Stommel, H., Arons, A. B., and Blanchard, D. 1956. An oceanographical curiosity, the perpetual salt fountain. Deep-Sea Res., 3, 152153.Google Scholar
Stommel, H., and Federov, K.N. 1967. Small scale structure in temperature and salinity near Timor and Mindanao. Tellus, 19, 306326.Google Scholar
Sundermeyer, M.A., and Lelong, M.-P. 2005. Numerical simulations of lateral dispersion by the relaxation of diapycnal mixing events. J. Phys. Oceanogr., 35(12), 23682386.Google Scholar
Sundermeyer, M.A., and Price, J.F. 1998. Lateral mixing and the North Atlantic Tracer Release Eexperiment: Observations and numerical simulations of Lagrangian particles and a passive tracer. J. Geophys. Res., 103, 481497.Google Scholar
Sutherland, B.R. 2010. Internal Gravity Waves. Cambridge, UK: Cambridge University Press.Google Scholar
Swift, S.A., Bower, A.S., and Schmitt, R.W. 2012. Vertical, horizontal, and temporal changes in temperature in the Atlantis II and Discovery hot brine pools, Red Sea. Deep-Sea Res. I, 64, 118128.Google Scholar
Tailleux, R. 2009. On the energetics of stratified turbulent mixing, irreversible thermodynamics, Boussinesq models and the ocean heat engine controversy. J. Fluid Mech., 638, 339382.Google Scholar
Tailleux, R. 2013. Available potential energy and energy in stratified fluids. Ann. Rev. Fluid Mech., 45, 3558.Google Scholar
Tait, R.I., and Howe, M.R. 1968. Some observations of thermohaline stratification in the deep ocean. Deep-Sea Res., 15, 275280.Google Scholar
Takahashi, A., and Hibiya, T. 2019. Assessment of finescale parameterizations of deep ocean mixing in the presence of geostrophic current shear: Results of microstructure measurements in the Antarctic Circumpolar Current region. J. Geophys. Res: Oceans, 124, 135153.Google Scholar
Talley, L.D. 2013. Closure of the global overturning circulation through the Indian, Pacific, and Southern Oceans: Schematics and transport. Oceanography, 26, 8097.Google Scholar
Talley, L.D., Pickard, G.L., Emery, W.J., and Swift, J.H. 2011. Descriptive Physical Oceanography: An Introduction. 6th ed. Elsevier Ltd.Google Scholar
Talley, L.D., and Yun, J.-Y. 2001. The role of cabbeling and double diffusion in setting the density of the North Pacific Intermediate Water salinity minimum. J. Phys. Oceanogr., 31(6), 15381549.Google Scholar
Taylor, G.I. 1935. Statistical theory of turbulence. Proc. R. Soc. Lond. A, 151, 421444.Google Scholar
Taylor, J., and Bucens, P. 1989. Laboratory experiments on the structure of salt fingers. Deep-Sea Res, 36, 16751704.Google Scholar
Tennekes, H., and Lumley, J.L. 1972. A First Course in Turbulence. Cambridge, MA: MIT Press.Google Scholar
Thorpe, S.A. 1971. Experiments on the instability of stratified shear flows: Miscible fluids. J. Fluid Mech., 46, 299319.Google Scholar
Thorpe, S.A. 1977. Turbulence and mixing in a Scottish loch. Proc. Roy. Soc. Lond. A, 286, 125181.Google Scholar
Thorpe, S.A. 1987. On the reflection of a strain of finite-amplitude internal waves from a uniform slope. J. Fluid Mech., 178, 299302.Google Scholar
Thorpe, S.A. 2005. The Turbulent Ocean. Cambridge University Press.Google Scholar
Thorpe, S.A., and Brubaker, J.M. 1983. Observation of sound reflection by temperature microstructure. Limnol. Oceanogr., 28, 601613.Google Scholar
Thorpe, S.A., Osborn, T.R., Jackson, J.F.E., Hall, A.J., and Lueck, R.G. 2002. Measurements of turbulence in the upper ocean mixing layer using Autosub. J. Phys. Oceanogr., 32(1), 122145.Google Scholar
Thurnherr, A.M., St. Laurent, L.C., Speer, K.G., Toole, J.M., and Ledwell, J.R. 2005. Mixing associated with sills in a canyon on the Midocean Ridge flank. J. Phys. Oceanogr., 35, 13701381.Google Scholar
Thwaites, F.T., Krishfield, R., Timmermans, M.-L., Toole, J.M., and III, Williams, A.J.. 2011. Noise in ice-tethered profiler and McLane Moored Profiler velocity measurements. In: Proc. IEEE/OES/CWTH Tenth Working Conf. on Current Measurement Technology.Google Scholar
Thwaites, F.T., Williams, III, A.J., Terray, E.A., and Trowbridge, J.H. 1995. A family of acoustic vorticity meters to measure ocean boundary layer shear. Pages 193–198 of: Proc. IEEE Fifth Working Conf. of Current Measurement.Google Scholar
Timmermans, M.-L., Garrett, C., and Carmack, E. 2003. The thermohaline structure and evolution of deep waters in the Canada Basin, Arctic Ocean. Deep-Sea Res I, 50, 13051321.Google Scholar
Timmermans, M.-L., Toole, J., Krishfield, R., and Winsor, P. 2008. Ice-tethered profiler observations of the double-diffusive staircases in the Canada Basin thermocline. J. Geophys. Res., 113(C00A02).Google Scholar
Toggweiler, J.R., and Samuels, B.J. 1998. On the ocean’s large-scale circulation near the limit of no vertical mixing. J. Phys. Oceanogr., 28, 18321852.Google Scholar
Toole, J.M., Polzin, K.L., and Schmitt, R.W. 1994. Estimates of diapycnal mixing in the abyssal ocean. Science, 264, 11201123.Google Scholar
Turner, J.S. 1965. The coupled transports of salt and heat across a sharp density interface. Intl. J. Heat Mass Transfer, 8, 759767.Google Scholar
Turner, J.S. 1968a. The behavior of a stable salinity gradient heated from below. J. Fluid Mech., 33, 183200.Google Scholar
Turner, J.S. 1968b. The influence of molecular diffusivity on turbulent entrainment across a density interface. J. Fluid Mech., 33, 639656.Google Scholar
Turner, J.S. 1973. Buoyancy Effects in Fluids. Cambridge University Press.Google Scholar
Turner, J.S. 1978. Double-diffusive intrusions into a density gradient. J. Geophys. Res., 83(C6), 28872901.Google Scholar
Turner, J.S., and Stommel, H. 1964. A new case of convection in the presence of combined vertical salinity and temperature gradients. Proc. Natl. Acad. Sci. USA, 52, 4953.Google Scholar
UNESCO. 1980. The Practical Salinity Scale 1978 and the International Equation of State of Seawater 1980. Technical Papers 36. UNESCO.Google Scholar
Vanneste, J. 2012. Balance and spontaneous wave generation in geophysical flows. Annu. Rev. Fluid Mech., 45, 147172.Google Scholar
Veronis, G. 1965. On finite amplitude instability in thermohaline convection. J. Mar. Res., 13, 117.Google Scholar
Veronis, G. 1968. Effect of a stabilizing gradient of solute on thermal convection. Tellus, 34, 315336.Google Scholar
Veronis, G. 1972. On properties of seawater defined by temperature, salinity, and pressure. J. Mar. Res., 30, 227255.Google Scholar
Visbeck, M., and Rhein, M. 2000. Is boundary mixing slowly ventilating Greenland Sea Deep Water? J. Phys. Oceanogr., 30, 215224.Google Scholar
Voulgaris, G., and Trowbridge, J.H. 1998. Evaluation of the acoustic Doppler velocimeter (ADV) for turbulence measurements. J. Atmos. Ocean. Tech., 15, 272289.Google Scholar
Waite, M.L., and Bartello, P. 2004. Stratified turbulence dominated by vortical motion. J. Fluid Mech., 517, 281308.Google Scholar
Wang, W., and Huang, R.X. 2004. Wind energy input to the surface waves. J. Phys. Oceanogr., 34, 12761280.Google Scholar
Washburn, L., and Käse, R.H. 1987. Double diffusion and the distribution of the density ratio in the Mediterranean waterfront south of the Azores. J. Phys. Oceanogr., 17(1), 1225.Google Scholar
Washburn, L., Duda, T.F., and Jacobs, D.C. 1996. Interpreting conductivity microstructure: Estimating the temperature variance dissipation rate. J. Atmos. Ocean. Tech., 13, 11661188.Google Scholar
Waterhouse, A.F., MacKinnon, J.A., Nash, J.D., Alford, M.H., Kunze, E., Simmons, H.I., Polzin, K.L., Laurent, L.C. St., Sun, O.M., Pinkel, R., Talley, L.D., Whalen, C.B., Huussen, T.N., Carter, G.S., Fer, I., Waterman, S., Garabato, A.C.N., Sanford, T.B., and Lee, C.M. 2014. Global patterns of diapycnal mixing from measurements of the turbulent dissipation rate. J. Phys. Oceanogr., 44(7), 18541872.Google Scholar
Waterman, S., Garabato, A.C. Naveira, and Polzin, K. 2013. Internal waves and turbulence in the Antarctic Circumpolar Current. J. Phys. Oceanogr., 43(2), 259282.Google Scholar
Waterman, S., Polzin, K.L., Garabato, A.C.N., Sheen, K.L., and Forryan, A. 2014. Suppression of internal wave breaking in the Antarctic Circumpolar Current near topography. J. Phys. Oceanogr., 44, 14661492.Google Scholar
Watson, A.J., and Ledwell, J.R. 2000. Oceanographic tracer release experiments using sulphur hexafluoride. J. Geophys. Res. Oceans, 105(C6), 14,325–14,337.Google Scholar
Watson, A.J., Messias, M.-J., Fogelqvist, E., Scoy, K.A. Van, Johannessen, T., Oliver, K.I.C., Stevens, D.P., Rey, F., Tanhua, T., Olsson, K.A., Carse, F., Simonsen, K., Ledwell, J.R., Jansen, E., Cooper, D.J., Kreupke, J.A., and Guilyardi, E. 1999. Mixing and convection in the Greenland Sea from a tracer-release experiment. Nature, 401, 902904.Google Scholar
Watson, A.J., Ledwell, J.R., Messias, M.-J., King, B.A., Mackay, N., Meredith, M.P., Mills, B., and Garabatoo, A.C. Naveira. 2013. Rapid cross-density ocean mixing at mid-depths in the Drake Passage measured by tracer release. Nature, 501, 408411.Google Scholar
Webb, D.J., and Suginohara, N. 2001. Vertical mixing in the ocean. Nature, 406, 37.Google Scholar
Weinstock, J. 1985. On the theory of temperature spectra in a stably stratified fluid. J. Phys. Oceanogr., 15(4), 475477.Google Scholar
Welander, P. 1959. An advective model of the ocean thermocline. Tellus, 11, 309318.Google Scholar
Wesson, J.C., and Gregg, M.C. 1988. Turbulent dissipation in the Strait of Gibraltar and associated mixing. Pages 201–212 of: Nihoul, J.C.J., and Jamart, B.M. (eds.), Small-Scale Turbulence and Mixing in the Ocean, Proceedings of the 19th International Liege Colloquium on Ocean Hydrodynamics. Amsterdam: Elsevier.Google Scholar
Wesson, J.C., and Gregg, M.C. 1994. Mixing at Camarinal Sill in the Strait of Gibraltar. J. Geophys. Res., 99, 98479878.Google Scholar
Whalen, C.B., Talley, L.D., and MacKinnon, J.A. 2012. Spatial and temporal variability of global ocean mixing inferred from Argo profiles. Geophys. Res. Lett., 39(L18612).Google Scholar
Whalen, C.B., MacKinnon, J.A., Talley, L.D., and Waterhouse, A.F. 2015. Estimating the mean diapycnal mixing using a finescale strain parameterization. J. Phys. Oceanogr., 45(4), 11741188.Google Scholar
White, W., and Bernstein, R. 1981. Large-scale vertical eddy diffusion in the main pycnocline of the central North Pacific. J. Phys. Oceanogr., 11(4), 434441.Google Scholar
Wijesekera, H., Padman, L., Dillon, T., Levine, M., Paulson, C., and Pinkel, R. 1993. The application of internal-wave dissipation models to a region of strong mixing. J. Phys. Oceanogr., 23, 269286.Google Scholar
Willebrand, J., Müller, P., and Olbers, D.J. 1977. Inverse Analysis of the Trimoored Internal Wave Experiment (IWEX). Berichte 20a. Inst. für Meereskunde, Christian-Albrechts-Universität, Kiel, Kiel, Germany.Google Scholar
Williams, A.J. 2014. Current measurement by differential acoustic travel-time reviewed. Pages 1–5 of: Baltic International Symposium (BALTIC). Xplore, vol. IEEE/OES. IEEE.Google Scholar
Winkel, D.P., Gregg, M.C., and Sanford, T.B. 1996. Resolving oceanic shear and velocity with the Multi-Scale Profiler. J. Atmos. Ocean. Tech., 13, 10461072.Google Scholar
Winkel, D.P., Gregg, M.C., and Sanford, T.B. 2002. Patterns of shear and turbulence across the Florida Current. J. Phys. Oceanogr., 32, 32693285.Google Scholar
Winters, K.B., and D’Asaro, E.A. 1994. Three-dimensional wave instability near a critical level. J. Fluid Mech., 272(August), 255284.Google Scholar
Winters, K.B., and D’Asaro, E.A. 1996. Diascalar flux and the rate of fluid mixing. J. Fluid Mech., 317(June), 179193.Google Scholar
Winters, K.D., and D’Asaro, E.A. 1997. Direct simulation of internal wave energy transfer. J. Phys. Oceanogr., 27(9), 19371945.Google Scholar
Woods, J.D. 1968. Wave-induced shear instability in the summer thermocline. J. Fluid Mech., 32(4), 791800.Google Scholar
Woods, J.D., Onken, R., and Fischer, J. 1986. Thermohaline intrusions created isopycnically at oceanic fronts are inclined to isopycnals. Nature, 322(31 July 1986), 446–449.Google Scholar
Wu, L.X., Jing, Z., Riser, S., and Visbeck, M. 2011. Seasonal and spatial variations of Southern Ocean diapycnal mixing from Argo profiling floats. Nat. Geosci., 4, 363366.Google Scholar
Wunsch, C. 2006. Discrete Inverse and State Estimation Problems: With Geophysical Fluid Applications. Cambridge University Press.Google Scholar
Wunsch, C., and Ferrari, R. 2004. Vertical mixing, energy, and the general circulation of the oceans. Annu. Rev. Fluid Mech., 36, 281314.Google Scholar
Wunsch, C., and Webb, S. 1979. The climatology of deep ocean internal waves. J. Phys. Oceanogr., 9, 235243.Google Scholar
Würger, A. 2010. Thermal non-equilibrium transport in colloids. Rep. Progr. Phys., 73.Google Scholar
Wust, G. 1933. Das bodenwasser und die Gliederung der Atlantischen Tiefsee. Wiss. Ergebn. Dtsch. Atlant. Exped. ‘Meteor’, 6(1), 1–107.Google Scholar
Yamazaki, H., and Lueck, R.G. 1990. Why oceanic dissipation rates are not lognormal. J. Phys. Oceanogr., 20, 19071908.Google Scholar
Yamazaki, H., and Osborn, T. 1990. Dissipation estimates for stratified turbulence. J. Geophys. Res., 95(C6), 97399744.Google Scholar
You, Y. 2002. A global ocean climatological atlas of the Turner angle: Implications for double-diffusion and water-mass structure. Deep-Sea Res. I, 49, 20752093.Google Scholar
Zhao, Z. 2016. Internal tide oceanic tomography. Geophys. Res. Lett., 43, 91579164.Google Scholar
Zhao, Z. 2018. The global mode-2 M2 internal tide. J. Geophys. Res.: Oceans, 123, 77257746.Google Scholar
Zhao, Z., and Alford, M.H. 2009. New altimetric estimates of Mode-1 M2 internal tides in the central North Pacific Ocean. J. Phys. Oceanogr., 39(7), 16691684.Google Scholar
Zhao, Z., Alford, M.H., Girton, J.B., Rainville, L., and Simmons, H.L. 2016. Global observations of open-ocean mode-1 M2 internal tides. J. Phys. Oceanogr., 46(6), 16571684.Google Scholar
Zodiatis, G., and Gasparini, G.P. 1996. Thermohaline staircase formations in the Tyrrhenian Sea. Deep-Sea Res. I, 43(5), 655678.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • Bibliography
  • Michael C. Gregg, University of Washington
  • Book: Ocean Mixing
  • Online publication: 10 March 2021
  • Chapter DOI: https://doi.org/10.1017/9781316795439.012
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • Bibliography
  • Michael C. Gregg, University of Washington
  • Book: Ocean Mixing
  • Online publication: 10 March 2021
  • Chapter DOI: https://doi.org/10.1017/9781316795439.012
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • Bibliography
  • Michael C. Gregg, University of Washington
  • Book: Ocean Mixing
  • Online publication: 10 March 2021
  • Chapter DOI: https://doi.org/10.1017/9781316795439.012
Available formats
×