Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-x4r87 Total loading time: 0 Render date: 2024-04-25T09:42:56.884Z Has data issue: false hasContentIssue false

13 - Lagrangian analysis and prediction of coastal and ocean dynamics (LAPCOD)

Published online by Cambridge University Press:  07 September 2009

Arthur J. Mariano
Affiliation:
Rosenstiel School of Marine and Atmospheric Science, University of Miami, Miami, Florida, USA
Edward H. Ryan
Affiliation:
Rosenstiel School of Marine and Atmospheric Science, University of Miami, Miami, Florida, USA
Annalisa Griffa
Affiliation:
University of Miami
A. D. Kirwan, Jr.
Affiliation:
University of Delaware
Arthur J. Mariano
Affiliation:
University of Miami
Tamay Özgökmen
Affiliation:
University of Miami
H. Thomas Rossby
Affiliation:
University of Rhode Island
Get access

Summary

Introduction

It was during the 1999 Liege Colloquium on, “Three-Dimensional Ocean Circulation: Lagrangian measurements and diagnostic analyses,” that a number of researchers started to discuss the idea of having a meeting centered on studying the ocean, the atmosphere, and marine biology from a Lagrangian viewpoint. At the time of this writing, three Lagrangian Analysis and Prediction of Coastal and Ocean Dynamics (LAPCOD) meetings have been held in (i) Ischia, Italy from October 2–6, 2000, (ii) Key Largo, FL, USA from December 12–16, 2002, and (iii) Lerici, Italy from June 13–17, 2005. The LAPCOD meetings bring together a diverse group of scientists for the purpose of exchanging ideas on the collection, analysis, modeling, and assimilation of coastal and oceanic (quasi-)Lagrangian data. The purpose of this chapter is to provide both a tutorial for readers who are not specialists, and a summary of the material presented at the LAPCOD meetings and in this book. Since this chapter summarizes the material presented at LAPCOD meetings and because of space constraints, many important Lagrangian-based studies are not detailed here and the chapter topics, listed in the next paragraph, are those topics that have been central to the LAPCOD meetings and this book. There are a number of unpublished results presented at LAPCOD 2005 discussed here and referenced by personal communication, hereafter pers. com.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2007

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

Aliani, S., Griffa, A., and Molcard, A., 2003. Floating debris in the Ligurian Sea, North-Western Mediterranean. Marine Pollution Bulletin, 46, 1142–9.CrossRefGoogle ScholarPubMed
Allen, A. and J. V. Plourde, 1999. Review of Leeway: Field Experiments and Implementation. Report CG-D-08-99, US Coast Guard Research and Development Center, 1082 Shennecossett Road, Groton, CT, USA.
Aref, H., 1984. Stirring by chaotic advection. J. Fluid Mech., 143, 1–21.CrossRefGoogle Scholar
Artale, V., Boffetta, G., Celani, A., Cencini, M., and Vulpiani, A., 1997. Dispersion of passive tracers in closed basins: beyond the diffusion coefficient. Phys. Fluids, 9, 3162.CrossRefGoogle Scholar
Ashjian, C. J., 1993. Trends in copepod species abundances across and along a Gulf Stream Meander: evidence for entrainment and detrainment of fluid parcels from the Gulf Stream. Deep-Sea Res., 40, 461–82.CrossRefGoogle Scholar
Assenbaum, M. and Reverdin, G., 2005. Near real-time analyses of the mesoscale circulation during the POMME experiment, Deep-Sea Res. I, 52, 1345–73.CrossRefGoogle Scholar
Aurell, E., Boffetta, G., Crisanti, A., Paladin, G., and Vulpiani, A., 1997. Predictability in the large: an extension of the concept of Lyapunov exponent. J. Phys. A, 30, 1–26.CrossRefGoogle Scholar
Bane, J. M. and Dewar, W. K., 1988. Gulf Stream bimodality and variability downstream of the Charleston bump. J. of Geophys. Res., 93(C6), 6695–710.CrossRefGoogle Scholar
Batchelor, G. K., 1952. Diffusion in a field of homogeneous turbulence. II. The relative motion of particles. Proc. Cambridge Philos. Soc., 48, 345–62.CrossRefGoogle Scholar
Bauer, S., Hitchcock, G. L., and Olson, D. B., 1991. Influence of monsoonally-forced Ekman dynamics upon surface layer depth and plankton biomass distribution in the Arabian Sea. Deep-Sea Res., 38(5A), 531–53.CrossRefGoogle Scholar
Bauer, S., Swenson, M. S., Griffa, A., Mariano, A. J., and Owens, K., 1998. Eddy-mean flow decomposition and eddy-diffusivity estimates in the tropical Pacific Ocean. J. Geophys. Res., 103, 30855–71.CrossRefGoogle Scholar
Bennett, A. F., 1984. Relative dispersion: local and nonlocal dynamics. J. Atmos. Sci., 41(11), 1881–6.2.0.CO;2>CrossRefGoogle Scholar
Bennett, A. F., 1987. A Lagrangian analysis of turbulent diffusion. Rev. Geophys., 25, 799–822.CrossRefGoogle Scholar
Bennett, A. F. and Chua, B. S., 1999. Open boundary conditions for Lagrangian geophysical fluid dynamics. J. Comput. Phys., 153(2), 418–36.CrossRefGoogle Scholar
Bennett, A., 2006. Lagrangian Fluid Dynamics (Cambridge Monographs on Mechanics). Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Berloff, P., McWilliams, J. C., and Bracco, A., 2002. Material transport in oceanic gyres. Part I: Phenomelogy. J. Phys. Oceanogr., 32, 764–96.2.0.CO;2>CrossRefGoogle Scholar
Berloff, P. and McWilliams, J. C., 2002. Material transport in oceanic gyres. Part II: Hierarchy of stochastic models. J. Phys. Oceanogr., 32, 797–830.2.0.CO;2>CrossRefGoogle Scholar
Berloff, P. and McWilliams, J. C., 2003. Material transport in oceanic gyres. Part III: Randomized stochastic models. J. Phys. Oceanogr., 33, 1416–45.2.0.CO;2>CrossRefGoogle Scholar
Beron-Vera, F. J., Olascoaga, M. J., and Brown, M. G., 2004. Tracer patchiness and particle trajectory stability in incompressible two-dimensional flows. Nonlin. Proc. Geophys., 11, 67–74.CrossRefGoogle Scholar
Bhatta, P., E. Fiorelli, F. Lekien, N. E. Leonard, D. A. Paley, F. Zhang, R. Bachmayer, R. E. Davis, D. Fratantoni, and R. Sepulchre, 2006. Coordination of an Underwater Glider Fleet for Adaptive Sampling, International Workshop on Underwater Robotics, (in press), 61–69.
Bitterman, D. S. and Hansen, D. V., 1989. Direct measurements of current shear in the Tropical Pacific Ocean and its effect on drift buoy performance. J. Atmos. Ocean. Tech., 6, 274–9.2.0.CO;2>CrossRefGoogle Scholar
Bitterman, D. B. and Hansen, D. V., 1993. Evaluation of sea surface temperature measurements from drifting buoys. J. Atmos. Ocean. Tech., 10(1), 88–96.2.0.CO;2>CrossRefGoogle Scholar
Blanke, B., Speich, S., Madec, G., and Doos, K., 2001. A global diagnostic of interocean mass transfers. J. Phys. Oceanogr., 31(6), 1623–32.2.0.CO;2>CrossRefGoogle Scholar
Boebel, O., Davis, R. E., Ollitrault, M., Peterson, R. G., Richardson, P. L., Schmid, C., and Zenk, W., 1999. The Intermediate Depth Circulation of the Western South Atlantic. Geophys. Res. Lett., 26(21), 3329–32.CrossRefGoogle Scholar
Boffetta, G., Cencini, M., Espa, S., and Querzoli, G., 2000. Chaotic advection and relative dispersion in an experimental convective flow. Amer. Inst. of Physics, 12/12, 3160–7.Google Scholar
Bower, A. S., 1991. A simple kinematic mechanism for mixing fluid parcels across a meandering jet. J. Phys. Oceanogr., 21, 173–80.2.0.CO;2>CrossRefGoogle Scholar
Bower, A. S. and Lozier, M. S., 1994. A closer look at particle exchange in the Gulf Stream. J. Phys. Oceanogr., 24, 1399–418.2.0.CO;2>CrossRefGoogle Scholar
Bower, A. S. and Rossby, H. T., 1989. Evidence of cross-frontal exchange processes in the Gulf Stream based on isopycnal RAFOS float data. J. Phys. Oceanogr., 19, 1177–90.2.0.CO;2>CrossRefGoogle Scholar
Bower, A. S., Cann, B., Rossby, T., Zenk, W., Gould, J., Speer, K., Richardson, P. L., Prater, M. D., and Zhang, H.-M., 2002a. Directly measured mid-depth circulation in the northeastern North Atlantic Ocean, Nature, 419, 603–7.CrossRefGoogle Scholar
Bower, A. S., Serra, N., and Ambar, I., 2002b. Structure of the Mediterranean Undercurrent and Mediterranean Water spreading around the southwestern Iberian Peninsula. J. Geophys. Res., 107(C10), 3161, doi:10.1029/2001JC001007.CrossRefGoogle Scholar
Bracco, A., LaCasce, J. H., Pasquero, C., and Provenzale, A., 2000. The velocity distribution of barotropic turbulence. Phys. Fluids, 12, 2478–88.CrossRefGoogle Scholar
Breivik, O. and Allen, A., 2005. An operational search and rescue model for the Norwegian Sea and the North Sea. submitted to J. Mar. Syst.Google Scholar
Brink, K. H., Beardsley, R. C., Niiler, P. P., Abbott, M. R., Huyer, A., Ramp, S. R., Stanton, T. P., and Stuart, D., 1991. Statistical properties of near-surface flow in the California coastal transition zone. J. Geophys. Res., 96(C8), 14693–706.CrossRefGoogle Scholar
Brink, K. H., Beardsley, R. C., Paduan, J., Limeburner, R., Caruso, M., and Sires, J., 2000. A view of the 1993–1994 California current based on surface drifters, floats, and remotely sensed data. J. Geophys. Res., 105(C4), 8575–604.CrossRefGoogle Scholar
Brodie, J. W., 1960. Coastal surface currents around New Zealand. New Zealand Journal of Geology and Geophysics, 3(2), 235–52.CrossRefGoogle Scholar
Bryden, H. L., Candela, J. C., and Kinder, T. H., 1994. Exchange through the Strait of Gibraltar. Prog. Oceanogr., 33, 201–48.CrossRefGoogle Scholar
Bumpus, D. F. and Lauzier, L. M., 1965. Surface Circulation on the Continental Shelf off Eastern North America between Newfoundland and Florida. Serial Atlas of the Marine Environment. New York: American Geographical Society.Google Scholar
Burt, W. V. and B. Wyatt, 1964. Drift bottle observations of the Davidson Current off Oregon. In Studies on Oceanography, ed. Yoshida, K.. Seattle: Univ. Washington Press, 156–65.Google Scholar
Carter, E. F., 1989. Assimilation of Lagrangian data into a numerical model. Dyn. Atmos. Oceans, 13, 335–48.CrossRefGoogle Scholar
Carter, E. F. and Robinson, A. R., 1987. Analysis models for the estimation of oceanic fields. J. Atmos. Ocean. Tech., 4, 49–74.2.0.CO;2>CrossRefGoogle Scholar
Castronovo, E. and Kramer, P. R., 2004. Subdiffusion and superdiffusion in Lagrangian stochastic models of oceanic transport, Monte Carlo Methods and Appl., 10(34), 245–56.CrossRefGoogle Scholar
Cecconi, F., Cencini, M., Falcioni, M., and Vulpiani, A., 2005. Brownian motion and diffusion: from stochastic processes to chaos and beyond. Chaos, 15, 26–102.CrossRefGoogle ScholarPubMed
Centurioni, L. R. and Niiler, P. P., 2003. On the surface currents of the Caribbean Sea. Geophys. Res. Lett., 30(6), doi:10.1029/2002GL016231, 1279.CrossRefGoogle Scholar
Centurioni, L. R. and Gould, W. J., 2004. Winter conditions in the Irminger Sea observed with profiling floats. J. Mar. Res., 62, 313–36.CrossRefGoogle Scholar
Cheney, R. E., Gemmill, W. H., Shank, M. K., Richardson, P. L., and Webb, D., 1976. Tracking a Gulf Stream ring with SOFAR floats. J. Phys. Oceanogr., 6, 741–9.2.0.CO;2>CrossRefGoogle Scholar
Cheney, R. E., Richardson, P. L., and Nagasaka, K., 1980. Tracing a Kuroshio ring with a free-drifting surface buoy. Deep-Sea Res., 27, 641–54.CrossRefGoogle Scholar
Cowen, R. K., Lwiza, K. M. M., Sponaugle, S., Paris, C. B., and Olson, D. B., 2000. Connectivity of marine populations: Open or closed? Science, 287, 857–9.CrossRefGoogle ScholarPubMed
Cowen, R. K., Paris, C. B., Olson, D. B., and Fortuna, J. L., 2003. The role of long distance dispersal versus local retention in replenishing marine populations. J. Gulf Caribbean Res., 14(2), 129–37.Google Scholar
Cresswell, G. R., 1977. The trapping of two drifting buoys by an ocean eddy. Deep-Sea Res., 24, 1203–9.CrossRefGoogle Scholar
Da Costa, M. V. and Blanke, B., 2004. Lagrangian methods for flow climatologies and trajectory error assessment. Ocean Modelling, 6, 335–58.CrossRefGoogle Scholar
Davis, R. E., 1983. Oceanic property transport, Lagrangian particle statistics, and their prediction. J. Mar. Res., 41, 163–94.CrossRefGoogle Scholar
Davis, R. E., 1985a. Drifter observations of coastal currents during CODE: The method and descriptive view. J. Geophys. Res., 90, 4741–55.CrossRefGoogle Scholar
Davis, R. E., 1985b. Drifter observations of coastal surface currents during CODE: The statistical and dynamical views. J. Geophys. Res., 90(C7), 4756–72.CrossRefGoogle Scholar
Davis, R. E., 1991. Observing the general circulation with floats. Deep Sea Res., 38(Suppl.1), 531–71.CrossRefGoogle Scholar
Davis, R., Webb, D. C., Regier, L. A., and Dufour, J.. 1992. The Autonomous Lagrangian Circulation Explorer (ALACE). J. Atmos. Ocean. Tech., 9, 264–85.2.0.CO;2>CrossRefGoogle Scholar
Davis, R. E., 1998. Preliminary results from directly measuring mid-depth circulation in the tropical and South Pacific. J. Geophys. Res., 103, 619–39.CrossRefGoogle Scholar
Davis, R. E. and W. Zenk, 2001. Subsurface Lagrangian Observations during the 1990s. In Ocean Circulation and Climate: Observing and Modeling the Global Ocean, ed. Church, J., Siedler, G., and Gould, J.. San Diego: Academic Press, Chapter 3.2.Google Scholar
Dewar, W. K. and Flierl, G. R., 1985. Particle trajectories and simple models of transport in coherent vortices. Dyn. Atmos. Oceans, 9, 21–52.CrossRefGoogle Scholar
Doglioli, A. M., Magaldi, M. G., Vezzulli, L., and Tucci, S., 2004. Development of a numerical model to study the dispersion of wastes coming from a marine fish farm in the Ligurian Sea (Western Mediterranean). Aquaculture, 231(1), 215–35.CrossRefGoogle Scholar
Drijfhout, S. S., Vries, P., Doos, K., and Coward, A. C., 2003. Impact of eddy-induced transport on the Lagrangian structure of the upper branch of the thermohaline circulation. J. Phys. Oceanogr., 33, 2141–55.2.0.CO;2>CrossRefGoogle Scholar
Duan, J. Q. and Wiggins, S., 1996. Fluid exchange across a meandering jet with quasi-periodic time variability. J. Phys. Oceanogr., 26, 1176–88.2.0.CO;2>CrossRefGoogle Scholar
Dvorkin, Y. and Paldor, N., 1999. Analytical considerations of Lagrangian cross-equatorial flow. J. Atmos. Sci., 56(9), 1229–37.2.0.CO;2>CrossRefGoogle Scholar
Ebbesmeyer, C. C. and Coomes, C. A., 1993. Historical shoreline recoveries of drifting objects: an aid for future shoreline utilization. Oceans 93 Proceedings, III, 159–64.Google Scholar
Eckart, C., 1948. An analysis of the stirring and mixing processes in incompressible fluids. J. Mar. Res., 7, 265–75.Google Scholar
Falco, P., Griffa, A., Poulain, P.-M., and Zambianchi, E., 2000. Transport properties in the Adriatic Sea as deduced from drifter data. J. Phys. Oceanogr., 30, 2055–71.2.0.CO;2>CrossRefGoogle Scholar
Festa, J. F. and Molinari, R. L., 1992. An evaluation of the WOCE volunteer observing ship-XBT network in the Atlantic. J. Atmos. Ocean. Tech., 9, 305–17.2.0.CO;2>CrossRefGoogle Scholar
Fiechter, J. and Mooers, C. N. K., 2003. Simulation of frontal eddies on the East Florida Shelf. Geophys. Res. Lett., 30(22), doi:10.1029/2003GL018307.CrossRefGoogle Scholar
Figueroa, H. A. and Olson, D. B., 1989. Lagrangian statistics in the South Atlantic as derived from SOS and FGGE drifters. J. Mar. Res., 47(3), 525–46.CrossRefGoogle Scholar
Flierl, G. R., 1981. Particle motions in large amplitude wave fields. Geophys. Astrophys. Fluid Dyn., 18, 39–74.CrossRefGoogle Scholar
Fratantoni, D. M., 2001. North Atlantic surface circulation during the 1990's observed with satellite-tracked drifters. J. Geophys. Res., 106(C10), 22067–93.CrossRefGoogle Scholar
Freeland, H. J., Rhines, P. B., and Rossby, T., 1975. Statistical observations of the trajectories of neutrally buoyant floats in the North Atlantic. J. Mar. Res., 33, 383–404.Google Scholar
Freeland, H. J. and Gould, W. J, 1976. Objective analysis of mesoscale ocean circulation features. Deep-Sea Res., 23, 915–24.Google Scholar
Freeland, H. J. and Cummins, P. F., 2005. Argo: A new tool for environmental monitoring and assessment of the world's oceans, an example from the NE Pacific. Prog. Oceanog., 64, 31–44.CrossRefGoogle Scholar
Garfield, N., Maltrud, M. E., Collins, C. A., Rago, T. A., and Paquette, R. G., 2001. Lagrangian flow in the California Undercurrent, an observation and model comparison. J. Mar. Syst., 29, 201–20.CrossRefGoogle Scholar
Garraffo, Z. D., Mariano, A. J., Griffa, A., Veneziani, C., and Chassignet, E. P., 2001a. Lagrangian data in a high resolution numerical simulation of the North Atlantic. I: Comparison with in-situ drifter data. J. Mar. Syst., 29/1–4, 157–76.CrossRefGoogle Scholar
Garraffo, Z. D., Griffa, A., Mariano, A. J., and Chassignet, E. P., 2001b. Lagrangian data in a high resolution numerical simulation of the North Atlantic. II: On the pseudo-Eulerian averaging of Lagrangian data. J. Mar. Syst., 29/1–4, 177–200.CrossRefGoogle Scholar
Gould, W. J., 2005. From Swallow floats to Argo – the development of neutrally buoyant floats, Deep-Sea Res. II, 52/3–4, 529–43.CrossRefGoogle Scholar
Gould, W. J., et al., 2001. Hydrographic Observations. In Observing the Oceans in the 21st Century, ed. Koblinsky, C. and Smith, N.. Victoria, Australia: CSIRO Publishing.Google Scholar
Griffa, A., 1996. Applications of stochastic particle models to oceanographic problems. In Stochastic Modeling in Physical Oceanography, ed. Adler, R., Muller, P., and Rozovoskii, B.. Cambridge, MA: Birkhäuser Boston.CrossRefGoogle Scholar
Grodsky, S. A. and Carton, J. A., 2002. Surface drifter pathways originating in the equatorial Atlantic cold tongue. Geophys. Res. Lett., 29, 23, 2147, doi:10.1029/2002GL015788.CrossRefGoogle Scholar
Haller, G., 2000. Finding finite-time invariant manifolds in two-dimensional velocity fields. Chaos, 10, 99–108.CrossRefGoogle ScholarPubMed
Halliwell, G. R., R. H. Weisberg, and D. Mayer, 2003. A synthetic float analysis of upper-limb meridional overturning circulation interior ocean pathways in the tropical/subtropical Atlantic. In Interhemisphere Water Exchange in the Atlantic Ocean, ed. Goni, G. and Malanotte-Rizzoli, P.. Amsterdam: Elsevier, 193–204.Google Scholar
Harcourt, R. R., Elizabeth, S. L., Garwood, R. W., and D'Asaro, E. A., 2002. Fully Lagrangian floats in Labrador Sea Deep Convection: comparison of numerical and experimental results. J. Phys. Oceanogr., 32, 493–510.2.0.CO;2>CrossRefGoogle Scholar
Hitchcock, G. L., Wiseman, W. L. Jr., Boicourt, W. C., Mariano, A. J., Walker, N., Nelsen, T., and Ryan, E. H., 1997. Property fields in the effluent plume of the Mississippi River. J. Mar. Syst., 12, 109–26.CrossRefGoogle Scholar
Hunt, J. R. C., 1998. Lewis Fry Richardson and his contributions to mathematics, meteorology, and models of conflict. Annu. Rev. Fluid Mech., 30, xiii–xxvi.CrossRefGoogle Scholar
Hurlburt, H. E. and Metzger, E. J., 1998. Bifurcation of the Kuroshio Extension at the Shatsky Rise. J. Geophys. Res., 103, 7549–66.CrossRefGoogle Scholar
Ide, K. and Ghil, M., 1998a. The extended Kalman filtering for vortex Systems, Part I. Methodology and point vortices. Dyn. Atmos. Oceans, 27(1–4), 301–32.CrossRefGoogle Scholar
Ide, K. and Ghil, M., 1998b. The extended Kalman filtering for vortex Systems, Part II. Rankine vortex and observing-system design. Dyn. Atmos. Oceans, 27(1–4), 333–50.CrossRefGoogle Scholar
Ide, K., Kuznetsov, L., and Jones, C. K. R. T., 2002a. Lagrangian data assimilation for point-vortex system. J. Turbulence, 3, 053.CrossRefGoogle Scholar
Ide, K., Small, D., and Wiggins, S., 2002b. Distinguished hyperbolic trajectories in time-dependent fluid flows: analytical and computational approach for velocity fields defined as data sets. Nonlinear Proc. Geoph., 9, 237–63.CrossRefGoogle Scholar
Idrisi, N., Olascoaga, M. J., Garraffo, Z. D., Olson, D. B., and Smith, S. L., 2004. Mechanisms for emergence from diapause of Calanoides carinatus in the Somali Current. Limnol. Oceanogr., (Biocomplexity Special Issue) 49(4, Part 2), 1262–8.CrossRefGoogle Scholar
Isern-Fontanet, J., Garcia-Ladona, E., and Font, J., 2006. The vortices of the Mediterranean sea: an altimetric perspective. J. Phys. Oceanogr., 36(1), 87–103.CrossRefGoogle Scholar
Kamachi, M. and O'Brien, J. J., 1995. Continuous assimilation of drifting buoy trajectories into an equatorial Pacific Ocean model. J. Mar. Sys., 6, 159–78.CrossRefGoogle Scholar
Kelly, K. A., Singh, S., and Huang, R. X., 1998. Seasonal variations of sea surface height in the Gulf Stream region. J. Phys. Oceanogr., 29(3), 313–27.2.0.CO;2>CrossRefGoogle Scholar
Kirwan, A. D. Jr., McNally, G., and Coehlo, J., 1976. Gulf Stream kinematics inferred from a satellite tracked drifter. J. Phys. Oceanogr., 6(5), 750–5.2.0.CO;2>CrossRefGoogle Scholar
Kirwan, A. D. Jr., McNally, G., Reyna, E., and Merrell, W. J. Jr., 1978. The near-surface circulation of the eastern North Pacific. J. Phys. Oceanogr., 8(6), 937–45.2.0.CO;2>CrossRefGoogle Scholar
Kirwan, A. D. Jr., Merrell, W. J. Jr., Lewis, J. K., and Whitaker, R. E., 1984. Lagrangian observations of an anticyclonic ring in the western Gulf of Mexico. J. Geophys. Res., 89(NC3), 3417–24.CrossRefGoogle Scholar
Kirwan, A. D. Jr., Toner, M., and Kantha, L., 2003. Predictability, uncertainty, and hyperbolicity in the ocean. Int. J. Engin. Sci., 41, 249–58.CrossRefGoogle Scholar
Kolmogorov, A. N., 1941. The local structure of turbulence in incompressible viscous fluid for very large Reynolds numbers. C. R. Acad. Sci. URSS, 30, 301.Google Scholar
Korotenko, K. A., Mamedov, R. M., and Mooers, C. N. K., 2002. Prediction of the transport and dispersal of oil in the South Caspian Sea resulting from blowouts. Environmental Fluid Mechanics, 1, 383–414.CrossRefGoogle Scholar
Kraichnan, R. H., 1966. Dispersion of particle pairs in homogeneous turbulence. Phys. Fluids, 9, 1937–43.CrossRefGoogle Scholar
Kraichnan, R. H., 1967. Inertial ranges in two-dimensional turbulence. Phys. Fluids, 10(7), 1417–23.CrossRefGoogle Scholar
Kuznetsov, L., Toner, M., Kirwan, A. D., Jones, C. K. R. T., Kantha, L. H., and Choi, J., 2002. The loop current and adjacent rings delineated by Lagrangian analysis of the near-surface flow. J. Mar. Res., 60(3), 405–29.CrossRefGoogle Scholar
Kuznetsov, L., Ide, K., and Jones, C. K. R. T., 2003. A method for assimilation for Lagrangian data. Mon. Wea. Rev., 131, 2247–60.2.0.CO;2>CrossRefGoogle Scholar
LaCasce, J., 2000. Floats and f/H. J. Mar. Res., 58, 61–95.CrossRefGoogle Scholar
LaCasce, J. H. and Bower, A., 2000. Relative dispersion in the subsurface North Atlantic. J. Mar. Res., 58(6), 863–94.CrossRefGoogle Scholar
LaCasce, J. H. and Ohlman, C., 2003. Relative dispersion at the surface of the Gulf of Mexico. J. Mar. Res., 61, 285–312.CrossRefGoogle Scholar
Lacorata, G., Aurell, E., and Vulpiani, A., 2001. Drifter dispersion in the Adriatic Sea: Lagrangian data and chaotic model. Ann. Geophys., 19, 121–9.CrossRefGoogle Scholar
Lacorata, G., E.Aurell, B.Legras, A.Vulpiani, , 2004. Evidence for a k− 5/3 spectrum from the EOLE Lagrangian balloons in the low stratosphere. J. Atmos. Sci., 61(23), 2936–42.CrossRefGoogle Scholar
Lavender, K. L., Owens, W. B., and Davis, R. E., 2005. The mid-depth circulation of the subpolar North Atlantic Ocean as measured by subsurface floats. Deep Sea Res. I, 52(5), 767–85.CrossRefGoogle Scholar
Leaman, K. D., Molinari, R., and Vertes, P., 1987. Structure and variability of the Florida Current at 27N: April 1982–July 1984. J. Phys. Oceanogr., 17, 565–83.2.0.CO;2>CrossRefGoogle Scholar
Leaman, K. D. and Vertes, P., 1996. Topographic influences on recirculation in the Deep Western Boundary Current: Results from RAFOS float trajectories between the Blake-Bahama Outer Ridge and the San Salvador “Gate”. J. Phys. Oceanogr., 26, 941–61.2.0.CO;2>CrossRefGoogle Scholar
Cann, B., Assenbaum, M., Gascard, J.-C., and Reverdin, G., 2005. Observed mean and mesoscale upper ocean circulation in the midlatitude northeast Atlantic. J. Geophys. Res., 110, C07S05, doi:10.1029/2004JC002768.Google Scholar
Lekien, F., Coulliette, C., Mariano, A. J., Ryan, E. H., Shay, L., Haller, G., and Marsden, J., 2005. Lagrangian structures in very high-frequency radar data along the coast of Florida and automated optimal pollution timing, Physica D, 210(1–2), 1–20.CrossRefGoogle Scholar
Lin, J. T., 1972. Relative dispersion in the enstrophy-cascading inertial range of homogeneous two-dimensional turbulence. J. Atmos. Sci., 29, 394–6.2.0.CO;2>CrossRefGoogle Scholar
Lipphardt, B. L. Jr., Kirwan, A. D. Jr., Grosch, C. E., Paduan, J. D., and Lewis, J. K., 2000. Blending HF radar and model velocities in Monterey Bay through normal mode analysis. J. Geophys. Res., 105(C2), 3425–50.CrossRefGoogle Scholar
Lobel, P. S. and Robinson, A. R., 1986. Transport and entrapment of fish larvae by ocean mesoscale eddies and currents in Hawaiian waters. Deep-Sea Res., 33(4), 483–500.CrossRefGoogle Scholar
Lozier, M. S., Bold, T. J., and Bower, A. S., 1996. The influence of propagating waves on cross-stream excursions. J. Phys. Oceanogr., 26(9), 1915–23.2.0.CO;2>CrossRefGoogle Scholar
Lumpkin, R., 2003. Decomposition of surface drifter observations in the Atlantic Ocean. Geophys. Res. Lett., 30(14), 1753, doi:10,1029/2003GL017519.CrossRefGoogle Scholar
Malhotra, N., and Wiggins, S., 1998. Geometric structures, lobe dynamics, and Lagrangian transport in flows with aperiodic time-dependence with applications to Rossby wave flow. J. Nonlin. Sci., 8, 401–56.CrossRefGoogle Scholar
Mancho, A., Small, D., and Wiggins, S., 2004. Computation of hyperbolic trajectories and their stable and unstable manifolds for oceanographic flows represented as data sets. Nonlinear Process. Geophys., 11(1), 17–33.CrossRefGoogle Scholar
Mariano, A. J. and Rossby, H. T., 1989. The Lagrangian Potential Vorticity Balance during POLYMODE, J. Phys. Oceanogr., 19(7), 927–39.2.0.CO;2>CrossRefGoogle Scholar
Mariano, A. J. and T. M. Chin, 1996. Feature and contour based data analysis and assimilation in physical oceanography. In Stochastic Modeling in Physical Oceanography, ed. Adler, R., Muller, P., and Rozovskii, B.. Cambridge, MA: Birkhäuser Boston, 311–42.CrossRefGoogle Scholar
Mariano, A. J., Griffa, A., Ozgokmen, T., and Zambianchi, E., 2002. Lagrangian analysis and predictability of coastal and ocean dynamics. J. Atmos. Ocean. Tech., 19(7), 1114–26.2.0.CO;2>CrossRefGoogle Scholar
Mazzino, A., Musacchio, S., and Vulpiani, A., 2005. Multiple-scale analysis and renormalization for pre-asymptotic scalar transport. Phys. Rev. E, 71, 111–13.CrossRefGoogle Scholar
McClean, J. L., Poulain, P.-M., Pelton, J. W., and Maltrud, M. E., 2002. Eulerian and Lagrangian statistics from surface drifters and a high-resolution POP simulation in the North Atlantic. J. Phys. Oceanogr., 32(9), 2472–91.CrossRefGoogle Scholar
McGillicuddy, D. J. and Robinson, A. R., 1997. Eddy-induced nutrient supply and new production in the Sargasso Sea. Deep-Sea Res. I, 44(8), 1427–50.CrossRefGoogle Scholar
McGillicuddy, D. J., Robinson, A. R., Siegel, D. A., Jannasch, H. W., Johnson, R., Dickey, T. D., McNeil, J., Michaels, A. F., and Knapp, A. H., 1998. Influence of mesoscale eddies on new production in the Sargasso Sea. Nature, 394, 263–6.CrossRefGoogle Scholar
McNally, G. J., Patzert, W. C., Kirwan, A. D. Jr., and Vastano, A. C., 1983. The near-surface circulation of the north Pacific using satellite tracked drifting buoys. J. Geophys. Res., 88(C9), 7507–18.CrossRefGoogle Scholar
Mead, J. L., 2004. The shallow water equations in Lagrangian coordinates. J. Comput. Phys., 200, 654–69.CrossRefGoogle Scholar
Mead, J. L. and Bennett, A. F., 2001. Towards regional assimilation of Lagrangian data: the Lagrangian form of the shallow water model and its inverse. J. Mar. Syst., 29, 365–84.CrossRefGoogle Scholar
Molcard, A., Piterbarg, L., Griffa, A., Özgökmen, T. M., and Mariano, A. J., 2003. Assimilation of drifter positions for the reconstruction of the Eulerian circulation field. J. Geophys. Res., 108(C3), doi:10.1029/200IJC001240.CrossRefGoogle Scholar
Molcard, A., Griffa, A., and Özgökmen, T. M., 2005. Lagrangian data assimilation in multi-layer primitive equation ocean models. J. Atmos. Ocean. Tech., 22, 70–83.CrossRefGoogle Scholar
Molinari, R. L. and Kirwan, A. D., 1975. Calculations of differential kinematic properties from Lagrangian observations in the western Caribbean Sea. J. Phys. Oceanogr., 5, 483–91.2.0.CO;2>CrossRefGoogle Scholar
Mooers, C. N. K. and Fiechter, J., 2003. Numerical simulations of mesoscale variability in the Straits of Florida. Ocean Dynamics, 55(3–4), doi:10.1007/s10236-005-0019-0, 309–325.CrossRefGoogle Scholar
Mooers, C. N. K., Bang, I., and Sandoval, F. J., 2005. Comparisons between observations and numerical simulations of Japan (East) Sea flow and mass fields in 1999 through 2001. Deep-Sea Res. II, 52(11–13), 1639–61.CrossRefGoogle Scholar
Neumann, G., 1968. Ocean Currents. Amsterdam: Elsevier.Google Scholar
Niiler, P., 2001. The world ocean surface circulation. In Ocean Circulation and Climate: Observing and Modeling the Global Ocean, ed. Church, J., Siedler, G., and Gould, J.. San Diego: Academic Press, 193–204.Google Scholar
Niiler, P. P., Davis, R. E., and White, H. J., 1987. Water-following characteristics of a mixed-layer drifter. Deep-Sea Res., 34, 1867–81.CrossRefGoogle Scholar
Niiler, P. P., Sybrandy, A. S., Bi, K., Poulain, P. M., and Bitterman, D., 1995. Measurements of the water-following capability of holey-sock and TRISTAR drifters. Deep-Sea Res., 42, 1951–64.CrossRefGoogle Scholar
Nodet, M., 2006. Variational assimilation of Lagrangian data in oceanography. Inverse Probl., 22, 245–63 doi:10.1088/0266-5611/22/1/014.CrossRefGoogle Scholar
Nùñez-Riboni, I., Boebel, O., Ollitrault, M., You, Y., Richardson, P., and Davis, R., 2005. Lagrangian circulation of Antarctic Intermediate Water in the subtropical South Atlantic. Deep-Sea Res. II, 52, 545–64.CrossRefGoogle Scholar
Oboukhov, A. M., 1941. Spectral energy distribution in a turbulent flow. Izv. Acad. Nauk SSR Geogr. Geofiez., 5, 453–66.Google Scholar
Ohlmann, J. C., White, P. F, Sybrandy, A. L., and Niller, P. P., 2005a. GPS-cellular drifter technology for coastal ocean observing systems, J. Atmos. Ocean. Tech., in press.CrossRefGoogle Scholar
Ohlmann, J. C., White, P. F., Washburn, L., Terrill, E., Emery, B., and Otero, M., 2005b. Interpretation of coastal HF radar derived surface currents with high resolution drifter data, J. Atmos. Ocean. Tech., in preparation.Google Scholar
Okubo, A., 1970. Horizontal dispersion of floatable particles in the vicinity of velocity singularities such as convergences. Deep-Sea Res., 17, 445–54.Google Scholar
Okubo, A., 1971. Oceanic diffusion diagrams. Deep-Sea Res., 18, 789–802.Google Scholar
Okubo, A. and Ebbesmeyer, C. C., 1976. Determination of vorticity, divergence and deformation rates from analysis of drogue observations. Deep-Sea Res., 23, 345–52.Google Scholar
Ollitrault, M., Gabillet, C., and Verdiere, A. C., 2005. Open ocean regimes of relative dispersion. J. Fluid Mech., 533, 381–407.CrossRefGoogle Scholar
Olson, D. B., Kourafalou, V. H., Johns, W. H., Samuels, G., and Veneziani, M., 2005. Aegean surface circulation from a satellite-tracked drifter array. submitted to J. Phys. Oceanogr.Google Scholar
d'Ovidio, F., Fernàndez, V., Herǹandez-Garcìa, E., and Lòpez, C., 2004. Mixing structures in the Mediterranean Sea from finite-size Lyapunov exponents. Geophys. Res. Lett., 31, doi:10.1029/2004GL020328.CrossRefGoogle Scholar
Owens, W. B., 1984. A synoptic and statistical description of the Gulf Stream and Subtropical Gyre using SOFAR floats. J. Phys. Oceanogr., 14(1), 104–13.2.0.CO;2>CrossRefGoogle Scholar
Owens, W. B., 1991. A statistical description of the mean circulation and eddy variability in the Northwest Atlantic using SOFAR floats. Prog. Oceanog., 28(3), 257–303.CrossRefGoogle Scholar
Özgökmen, T. M., Molcard, A., Chin, T. M., Piterbarg, L. I., and Griffa, A., 2003. Assimilation of drifter positions in primitive equation models of midlatitude ocean circulation. J. Geophys. Res., 108(C7), 32–8.CrossRefGoogle Scholar
Paduan, J. D. and Cook, M. S., 1997. Mapping surface currents in Monterey Bay with CODAR-type HF radar. Oceanography, 10, 49–52.CrossRefGoogle Scholar
Paduan, J. D. and Shulman, I., 2004. HF radar data assimilation in the Monterey Bay area. J. Geophys. Res., 109, C07S09, doi:10.1029/2003JC001949.CrossRefGoogle Scholar
Paldor, N., 2000. The transport in the Ekman surface layer on the spherical Earth. J. Mar. Res., 60, 47–72.CrossRefGoogle Scholar
Paldor, N., 2001. The zonal drift associated with time-dependent particle motion on the earth. Q. J. Roy. Meteor. Soc., 127A(577), 2435–50.CrossRefGoogle Scholar
Paldor, N., Sigalov, A., and Nof, D., 2003. The mechanics of eddy transport from one hemisphere to the other. Q. J. Roy. Meteor. Soc., 129B(591), 2011–25.CrossRefGoogle Scholar
Paldor, N., Dvorkin, Y., Mariano, A. J., Özgökmen, T. M., and Ryan, E., 2004. A practical, hybrid model for predicting the trajectories of near-surface ocean drifters. J. Atmos. Ocean. Tech., 21(8), 1246–58.2.0.CO;2>CrossRefGoogle Scholar
Paris, C. B. and Cowen, R. K., 2004. Direct evidence of a biophysical retention mechanism for coral reef fish larvae. Limnol. Oceanogr., 49, 1964–79.CrossRefGoogle Scholar
Pasquero, C., Provenzale, A., and Babiano, A., 2001. Parameterization of dispersion in two-dimensional turbulence. (Under consideration for publication in J. Fluid Mech., 1–30.)Google Scholar
Pasquero, C., Provenzale, A., and Weiss, J. B., 2002. Vortex statistics from Eulerian and Lagrangian time series. Phys. Rev. Lett., 89, 284–501.CrossRefGoogle ScholarPubMed
Pasquero, C., A. Bracco, and A. Provenzale, 2004. Coherent vortices, Lagrangian particles and the marine ecosystem. In Shallow Flows, ed. Jirka, G. H. and Uijttewaal, W. S. J.. Leiden, NL: Balkema Publishers, 399–412.Google Scholar
Perez-Brunius, P., Rossby, T., and Watts, D. R., 2004. A method for obtaining the mean transports of ocean currents by combining isopycnal float data with historical hydrography. J. Atmos. Ocean. Tech., 21(2), 298–316.2.0.CO;2>CrossRefGoogle Scholar
Peters, H., Shay, L. K., Mariano, A. J., and Cook, T. M., 2002. Current variability on a narrow shelf with large ambient vorticity. J. Geophys. Res., 107(C8), doi:10.1029/2001JC000813.CrossRefGoogle Scholar
Petrissans, A., Tanire, A., and Oesterl, B., 2002. Effects of nonlinear drag and negative loop correlations on heavy particle motion in isotropic stationary turbulence using a new Lagrangian stochastic model. Aerosol Sci. Tech., 36(9), 963–71.CrossRefGoogle Scholar
Piterbarg, L. I., 2005. On predictability of particle clusters in a stochastic flow. Stochastics and Dynamics, 5(1), 111–31.CrossRefGoogle Scholar
Pochapsky, T. E., 1961. Exploring subsurface waves with neutrally buoyant floats, Instrument Society of America Journal, 8, 34–7.Google Scholar
Pochapsky, T. E., 1963. Measurements of small scale oceanic motions with neutrally buoyant floats. Tellus, 4, 5352–62.Google Scholar
Pochapsky, T. E., 1966. Measurements of deep water movements with instrumented neutrally buoyant floats. J. Geophys. Res., 71, 2491–504.CrossRefGoogle Scholar
Poje, A. and Haller, G., 1999. Geometry of cross-stream Lagrangian mixing in a double gyre ocean model. J. Phys. Oceanogr., 29, 1649–65.2.0.CO;2>CrossRefGoogle Scholar
Poje, A. C., Toner, M., Kirwan, A. D. Jr., and Jones, C. K. R. T., 2002. Drifter launch strategies based on Lagrangian templates. J. Phys. Oceanogr., 32, 1855–69.2.0.CO;2>CrossRefGoogle Scholar
Poulain, P.-M., 2001. Adriatic Sea surface circulation as derived from drifter data between 1990 and 1999. J. Mar. Syst., 29, 332.CrossRefGoogle Scholar
Prater, M. D. and Rossby, T., 2005. Observations of the Faroe Bank Channel Overflow using bottom-following RAFOS floats. Deep Sea Res. II, 52(3–4), 481–494. doi:10.1016/j.dsr2.2004.12.009.CrossRefGoogle Scholar
Provenzale, A., 1999. Transport by coherent barotropic vortices. Annu. Rev. Fluid Mech., 31, 55–93.CrossRefGoogle Scholar
Reid, J., Schwarzlose, R. A., and Brown, D. M., 1963. Direct measurements of a small surface eddy off northern Baja California. J. Mar. Res., 21, 205–18.Google Scholar
Reverdin, G., Niiler, P. P., and Valdimarsson, H., 2003. North Atlantic Ocean surface currents. J. Geophys. Res., 108(C1), doi:10.1029/2001JC001020.CrossRefGoogle Scholar
Reynolds, A. M., 2002a. On Lagrangian stochastic modelling of material transport in oceanic gyres. Physica D., 172, 124–38.CrossRefGoogle Scholar
Reynolds, A. M., 2002b. Lagrangian stochastic modeling of anomalous diffusion in two-dimensional turbulence. Phys. Fluids., 14, 1442–9.CrossRefGoogle Scholar
Reynolds, A. M., 2003a. Third-order Lagrangian stochastic modeling. Phys. Fluids., 15, 2773–7.CrossRefGoogle Scholar
Reynolds, A. M., 2003b. On the application of nonextensive statistics to Lagrangian turbulence. Phys. Fluids., 115, L1–4.CrossRefGoogle Scholar
Richardson, L. F., 1926. Atmospheric diffusion shown on a distance-neighbour graph. Proc. R. Soc., A 110, 709–37.CrossRefGoogle Scholar
Richardson, P. L., 1976. Tracking a Gulf Stream Ring with SOFAR Floats. J. Phys. Oceanogr., 6(5), 741–9.Google Scholar
Richardson, P. L., 1980. Gulf Steam ring trajectories. J. Phys. Oceangr., 10, 90–104.2.0.CO;2>CrossRefGoogle Scholar
Richardson, P. L., 1981. North Atlantic subtropical gyre: SOFAR floats tracked by moored listening stations. Science, 213(4506), 435–7.CrossRefGoogle ScholarPubMed
Richardson, P. L., 1989. Worldwide ship drift distributions identify missing data. J. Geophys. Res., 94(C5), 6169–76.CrossRefGoogle Scholar
Richardson, P. L., 1993. A census of eddies observed in North Atlantic SOFAR float data. Prog. Oceanog., 31, 1–50.CrossRefGoogle Scholar
Richardson, P. L. and McKee, T. K., 1984. Average seasonal variation of the Atlantic equatorial currents from historical ship drifts. J. Phys. Oceanogr., 14(7), 1226–38.2.0.CO;2>CrossRefGoogle Scholar
Richardson, P. L. and Reverdin, G., 1987. Seasonal cycle of velocity in the Atlantic North Equatorial Countercurrent as measured by surface drifters, current meters, and ship drifts. J. Geophys. Res., 92(C4), 3691–708.CrossRefGoogle Scholar
Richardson, P. L., Walsh, D., Armi, L., Schroeder, M., and Price, J. F., 1989. Tracking three meddies with SOFAR floats. J. Phys. Oceanogr., 19, 371–83.2.0.CO;2>CrossRefGoogle Scholar
Richardson, P. L., Bower, A. S., and Zenk, W., 2000. A census of meddies tracked by floats. Prog. Oceanog., 45, 209–50.CrossRefGoogle Scholar
Richardson, R. L. and Stommel, H., 1948. Note on eddy diffusion in the sea. J. Meteorol., 5, 238–40.2.0.CO;2>CrossRefGoogle Scholar
Riser, S. C., Freeland, H. J., and Rossby, H. T., 1978. Mesoscale motions near the deep western boundary of the North Atlantic. Deep-Sea Res., 25, 1179–91.CrossRefGoogle Scholar
Roemmich, D., O. Boebel, Y. Desaubies, H. Freeland, B. King, P.-Y. LeTraon, R. Molinari, W. B. Owens, S. Riser, U. Send, K. Takeuchi, and S. Wijffels, 2001. Argo: the global array of profiling floats. In Observing the Oceans in the 21st Century, ed. Koblinsky, C. and Smith, N.. Victoria, Australia: CSIRO Publishing.Google Scholar
Rogerson, A., Miller, P. D., Pratt, L. J., and Jones, C. K. R. T. J., 1999. Lagrangian motion and fluid exchange in a barotropic meandering jet. J. Phys. Oceanogr., 29(10), 2635–55.2.0.CO;2>CrossRefGoogle Scholar
Rossby, H. T., S. C. Riser, and A. J. Mariano, 1983. The western North Atlantic – a Lagrangian viewpoint. In Eddies in Marine Science, ed. Robinson, A. R.. Heidelberg: Springer-Verlag, 66–91.CrossRefGoogle Scholar
Rudnick, D. L., Davis, R. E., Eriksen, C. C., Fratantoni, D. M., and , M. J.Perry, , 2004. Underwater gliders for ocean research. Mar. Technol. Soc. J., 38(1), 48–59.CrossRefGoogle Scholar
Rupolo, V., Marullo, S., and Iudicone, D., 2003. Eastern Mediterranean Transient studied with Lagrangian diagnostics applied to a Mediterranean OGCM forced by satellite SST and ECMWF wind stress for the years 1988–1993. J. Geophys. Res., 108(C9), doi:10.1029/2002JC001403.CrossRefGoogle Scholar
Salman, H., Kuznetsov, L., Jones, C. K. R. T., and Ide, K., 2006. A method for assimilating Lagrangian data into a shallow-water equation ocean model. submitted to Mon. Weather Rev.CrossRefGoogle Scholar
Samelson, R. M., 1992. Fluid exchange across a meandering jet. J. Phys. Oceanogr., 22, 431–40.2.0.CO;2>CrossRefGoogle Scholar
Sawford, B. L., 1999. Rotation of trajectories in Lagrangian stochastic models of turbulent dispersion. Bound-Lay. Meteorol., 93, 411–24.CrossRefGoogle Scholar
Sawford, B. L., 2001. Turbulent relative dispersion. Annu. Rev. Fluid Mech., 33, 289–317.CrossRefGoogle Scholar
Schmid, C., Z. Garraffo, E. Johns, and S. L. Garzoli, 2003. Pathways and variability at intermediate depths in the tropical Atlantic. In Interhemispheric Water Exchange in the Atlantic Ocean, ed. Goni, G. J. and Malarotte-Rizzoli, P.. Amsterdam: Elsevier, 233–68.
Schmitz, W. J. Jr., 1985. SOFAR float trajectories associated with the Newfoundland Basin. J. Mar. Res., 43, 761–78.CrossRefGoogle Scholar
Schneider, T., 1998. Lagrangian drifter models as search and rescue tools. M. S. Thesis, Dept. of Meteorology and Physical Oceanography, University of Miami.
Shay, L. K., Cook, T. M., Haus, B. K., Martinez, J., Peters, H., Mariano, A. J., VanLeer, J., Smith, S. M., An, P. E., Soloview, A., Weisberg, R., and Luther, M., 2000. A submesoscale vortex detected by very high resolution radar. Eos Trans. Amer. Geophys. Union, 81, 209–13.CrossRefGoogle Scholar
Shay, L. K., Cook, T. M., Peters, H., Mariano, A. J., Weisberg, R., An, P. E., Soloview, A., and Luther, M., 2002. Very high frequency radar mapping of surface currents, IEEE JOE, 27(2), 155–69.Google Scholar
Sheinbaum, J., Candela, J., Badan, A., and Ochoa, J., 2002. Flow structure and transport in the Yucatan Channel. Geophys. Res. Lett., 29(3), doi:10.1029/2001GL013990.CrossRefGoogle Scholar
Spaulding, M. L., 1988. A state-of-the-art review of oil spill trajectory and fate modeling. Oil Chem. Poll., 3, 455–69.Google Scholar
Stalcup, M. C. and Parker, C. E., 1965. Drogue measurements of shallow currents on the equator in the Western Atlantic Ocean. Deep-Sea Res., 12, 535–6.Google Scholar
Stommel, H., 1949. Horizontal diffusion due to oceanic turbulence. J. Mar. Res., 8, 199–225.Google Scholar
Summers, D. M., 2005. Eddy diffusion in the sea: reinterpreting an early experiment. Proc. R. Soc. Mat., 461(2058), 1811–18.CrossRefGoogle Scholar
Swallow, J. C., 1955. A neutral-buoyancy float for measuring deep currents. Deep-Sea Res., 3(1), 74–81.Google Scholar
Swallow, J. C., 1957. Some further deep current measurements using neutrally buoyant floats. Deep-Sea Res., 4, 93–104.Google Scholar
Swallow, J. C., 1971. The Aries current measurements in the Western North Atlantic. Philos. Trans. Roy. Soc. Lond. A, 270, 451–60.CrossRefGoogle Scholar
Taillandier, V., Griffa, A., and Molcard, A., 2006a. A variational approach for the reconstruction of regional scale Eulerian velocity fields from Lagrangian data. Ocean Modelling, 13, 1–24.CrossRefGoogle Scholar
Taillandier, V., A. Griffa, P.-M. Poulain, and K. Branger, 2006b. Assimilation of Argo float positions in the North Western Mediterranean Sea and impact on ocean circulation simulations. (Submitted to Geophys. Res. Lett.)
Talley, L. D., 2003. Shallow, intermediate, and deep overturning components of the global heat budget. J. Phys. Oceanogr., 33, 530–60.2.0.CO;2>CrossRefGoogle Scholar
Taylor, G. I., 1921. Diffusion by continuous movements. Proc. Lond. Math. Soc., Ser. 2, 20, 196–211.Google Scholar
Testor, P. and Gascard, J. C., 2005. Large scale flow separation and mesoscale eddy formation in the Algerian Basin. Prog. Oceanog., 66, 211–30.CrossRefGoogle Scholar
Thomson, D. J., 1987. Criteria for the selection of stochastic models of particle trajectories in turbulent flows. J. Fluid Mech., 180, 529–56.CrossRefGoogle Scholar
Thomson, D. J., 1990. A stochastic model for the motion of particle pairs in isotropic high Reynolds number, and its application to the problem of concentration variance. J. Fluid Mech., 210, 113–53.CrossRefGoogle Scholar
Toner, M. and Poje, A. C., 2004. Lagrangian velocity statistics of directed launch strategies in a Gulf of Mexico model. Nonlinear Proc. Geoph., 11, 35–46.CrossRefGoogle Scholar
Toner, M., Kirwan, A., Kantha, L., and Choi, J., 2001a. Can general circulation models be assessed and enhanced with drifter data?J. Geophys. Res., 106, 1366–83.CrossRefGoogle Scholar
Toner, M., Poje, A., Kirwan, A., Jones, C., Liphardt, B., and Grosch, C., 2001b. Reconstructing basin-scale Eulerian velocity fields from simulated drifter data. J. Phys. Oceanogr., 31, 1361–76.2.0.CO;2>CrossRefGoogle Scholar
Toner, M., Kirwan, A., Poje, A., Kantha, L., Muller-Karger, F. E., and Jones, C., 2003. Chlorophyll dispersal by eddy-eddy interactions in the Gulf of Mexico. J. Geophys. Res., 108, 3105, doi:10.1029/2002JC001499.CrossRefGoogle Scholar
Uchida, H. and Imawaki, S., 2003. Eulerian mean surface velocity field derived by combining drifter and satellite altimeter data. Geophys. Res. Lett., 30, 1229, doi:10.1029/2002GL016445.CrossRefGoogle Scholar
Veneziani, M., Griffa, A., Reynolds, A. M., and Mariano, A. J., 2004. Oceanic turbulence and stochastic models from subsurface Lagrangian data for the North-West Atlantic Ocean, J. Phys. Oceanogr., 34, 1884–906.2.0.CO;2>CrossRefGoogle Scholar
Veneziani, M., Griffa, A., Garraffo, Z. D., and Chassignet, E. P., 2005. Lagrangian spin parameter and coherent structures from trajectories released in a high-resolution ocean model. J. Mar. Res., 63(4), 753–88.CrossRefGoogle Scholar
Volkman, G., Knauss, J., and Vine, A., 1956. The use of parachute drogues in the measurement of subsurface ocean currents. Eos Trans. Amer. Geophys. Union, 37, 573–7.CrossRefGoogle Scholar
Volkman, G., 1963. Deep-current measurements using neutrally buoyant floats. In The Sea, ed. Hill, M. N.. New York: Interscience, 297–302.
Vries, P., and Doos, K., 2001. Calculating Lagrangian trajectories using time-dependent velocity fieldsJ. Atmos. Ocean. Tech., 18, 1092–101.2.0.CO;2>CrossRefGoogle Scholar
Webb, D. C., Simonetti, P. J, and Jones, C. P., 2001. SLOCUM: An underwater glider propelled by environmental energy. IEEE J. Oceanic Eng., 26(4), 447–52.CrossRefGoogle Scholar
Weiss, J., 1991. The dynamics of enstrophy transfer in two-dimensional hydrodynamics. Physica D, 48, 273–94.CrossRefGoogle Scholar
Welander, P., 1955. Studies on the general development of motion in a two-dimensional, ideal fluid. Tellus, 7, 141–56.CrossRefGoogle Scholar
Wyrtki, K., Magaard, L., and Hager, J., 1976. Eddy energy in the oceans. J. Geophys. Res., 81(15), 2641–6.CrossRefGoogle Scholar
Yang, Q., Parvin, B., Mariano, A. J., Ryan, E. H., Evans, R., and Brown, O. B., 2004. Seasonal and interannual studies of vortices in sea surface temperature data. An updated version of a paper originally presented at Oceans from Space ‘Venice 2000’ Symposium, Venice, Italy, 9–13 October 2000. Int. J. Remote Sens., 25(7–8), 1371–6.CrossRefGoogle Scholar
Zhang, H.-M., Prater, M. D., and Rossby, T., 2001. Isopycnal Lagrangian statistics from the North Atlantic Current RAFOS floats observations. J. Geophys. Res., 106, 13817–36.CrossRefGoogle 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.

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.

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.

Available formats
×