Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-28T04:01:58.967Z Has data issue: false hasContentIssue false

11 - Plankton: Lagrangian inhabitants of the sea

Published online by Cambridge University Press:  07 September 2009

Gary L. Hitchcock
Affiliation:
Division of Marine Biology and Fisheries, Rosenstiel School of Marine and Atmospheric Science, University of Miami, Miami, Florida, USA
Robert K. Cowen
Affiliation:
Division of Marine Biology and Fisheries 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

Plankton have inhabited the Earth's oceans for hundreds of millions of years as evidenced by the fossil record. The exterior covering of identifiable dinoflagellates, for example, are well preserved in Mesozoic rock strata. Pelagic diatoms possess siliceous frustules with identifiable species dating from early Cretaceous sediments (see Falkowski et al., 2004). Given the extent of fossil plankton, it is apparent that a drifting mode of life has been a successful means for survival in the sea for much of life's history.

With the importance of plankton in marine ecosystems, it is surprising that biological oceanographers have only recently begun to use drifting, or more formally Lagrangian, techniques. However, as with other aspects of biological oceanography, the Lagrangian ‘tools’ for studying plankton are relatively recent, and have often followed technique development by physical oceanographers and engineers. The main goal of this chapter is to summarize how biological oceanographers have applied Lagrangian and related methods to further our understanding of oceanic plankton distributions and dynamics, as well as biogeochemical processes. Our target audience is physical oceanographers and mathematicians who will hopefully gain some benefit from this exercise, while biological oceanographers may also be encouraged to further consider Lagrangian approaches in their field studies. We include studies on bacterio-, phyto-, zoo-, and ichthyoplankton and discuss the advances made in specific sub-disciplines of biological oceanography through the use of Lagrangian techniques. This review is timely in that new, low power sensors are now being adapted for deployments on a variety of Lagrangian platforms.

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

Abbott, M. R., Brink, K. H., Booth, C. R., Blasco, D., Codispoti, L. A., Niiler, P. P., and Ramp, S. R., 1990. Observations of phytoplankton and nutrients from a Lagrangian drifter off northern California. J. Geophys. Res., 95, 9393–409.CrossRefGoogle Scholar
Abbott, M. R., Brink, K. H., Booth, C. R., Blasco, D., Swenson, M. S., Davis, C. O., and Codispoti, L. A., 1995. Scales of variability of bio-optical properties as observed from near-surface drifters. J. Geophys. Res., 100, 13,345–67.CrossRefGoogle Scholar
Abbott, M. R. and R. M. Letelier, 1997. Going with the flow – The use of optical drifters to study phytoplankton dynamics. In Monitoring Algal Blooms: New Techniques for Detecting Large-Scale Environmental Change, ed. Kahru, M. and Brown, C.. New York: R. G. Landes Co., 145–70.Google Scholar
Abbott, M. R. and Letelier, R. M., 1998. Decorrelation scales of chlorophyll as observed from bio-optical drifters in the California Current. Deep-Sea Res. II. Top. Stud. Oceanogr., 45 (8–9), 1639–67.CrossRefGoogle Scholar
Abbott, M. R., Richman, J. G., Letelier, R. M., and Bartlett, J. S., 2000. The spring bloom in the Antarctic Polar Frontal Zone as observed from a mesoscale array of bio-optical sensors. Deep-Sea Res. II. Top. Stud. Oceanogr., 47 (15–16), 3285–314.CrossRefGoogle Scholar
Alexander, J. E. and Corcoran, E. F., 1963. Distribution of chlorophyll in the Straits of Florida. Limnol. Oceanogr., 8(2), 294–6.CrossRefGoogle Scholar
Arnold, W. S., Hitchcock, G. L., Frischer, M. E., Wanninkhof, R., and Sheng, Y. P., 2005. Dispersal of an introduced larval cohort in a coastal lagoon. Limnol. Oceanogr., 50(2), 587–97.CrossRefGoogle Scholar
Batchelder, H. P., Edwards, C. A., and Powell, T. M., 2002. Individual-based models of copepod populations in coastal upwelling regions: implications of physiologically and environmentally influenced diel vertical migration on demographic success and nearshore retention. Deep-sea Res., 53 (2–4), 307–33.Google Scholar
Becker, D. S., 1978. Evaluation of a hard clam spawner transplant site using a dye tracer technique. Marine Sciences Research Center. State Univ. of New York, Stony Brook (USA).Google Scholar
Bidigare, R. R., B. B. Prézelin, and R. C. Smith, 1992. Bio-optical models and the problems of scaling. In Primary Productivity and Biogeochemical Scales in the Sea, ed. Falkowski, P. G. and Woodhead, A. D.. New York: Plenum Press, 175–212.CrossRefGoogle Scholar
Bishop, J. K. B., Davis, R. E., and Sherman, J. T., 2002. Robotic observations of dust storm enhancement of carbon biomass in the North Pacific. Science, 298, 817–21.CrossRefGoogle ScholarPubMed
Bishop, J. K. B., Wood, T. J., Davis, R. E., and Sherman, J. T., 2004. Robotic observations of enhanced carbon biomass and export at 55 degree S during SOFeX. Science, 304, 417–20.CrossRefGoogle Scholar
Bjoerke, H., 1978. Food and feeding of young herring larvae of Norwegian spring spawners. Fiskeridir. Skr. (Havunders.), 16(11), 405–22.Google Scholar
Boicourt, W. C., Chao, S.-Y., Ducklow, H. W., Glibert, P. M., Malone, T. C., Roman, M. R., Sanford, L. P., Fuhrman, J. A., Garside, C., and Garvine, R. W., 1987. Physics and microbial ecology of a buoyant estuarine plume on the continental shelf. Eos., 68 (31), 666–8.CrossRefGoogle Scholar
Bone, Q., 1998. The Biology of Pelagic Tunicates. Oxford: Oxford University Press.Google Scholar
Botsford, L. W., Hastings, A., and Gaines, S. D., 2001. Dependence of sustainability on the configuration of marine reserves and larval dispersal distance. Ecol. Lett., 4, 144–50.CrossRefGoogle Scholar
Bowden, K. F., 1965. Horizontal mixing in the sea due to a shearing current. J. Fluid. Mech., 21, 83–95.CrossRefGoogle Scholar
Boyd, P. W., Watson, A. J., Law, C. S.et al., 2000. A mesoscale phytoplankton bloom in the polar Southern Ocean stimulated by iron fertilization. Nature, 407, 695–702.CrossRefGoogle ScholarPubMed
Boyd, P. W. and Law, C. S., 2001. The Southern Ocean Iron RElease Experiment (SOIREE) – introduction and summary. Deep-Sea Res. II Top. Stud. Oceanogr., 48 (11–12), 2425–38.CrossRefGoogle Scholar
Boyd, P. W.et al., 2004. The decline and fate of an iron-induced subarctic phytoplankton bloom. Nature, 428, 549–53.CrossRefGoogle ScholarPubMed
Brink, K. H., B. H. Jones, J. C. Van Leer, C. N. K. Mooers, D. W. Stuart, M. R. Stevenson, R. C. Dugdale, and G. W. Heburn, 1981. Physical and biological structure and variability in an upwelling center off Peru near 15° S during March 1977. In Coastal Upwelling, ed. Richards, F. A.. Washington DC: American Geophysical Union, 473–95.CrossRefGoogle Scholar
Brown, P. C. and Hutchings, L., 1987. The development and decline of phytoplankton blooms in the southern Benguela upwelling system. 1. Drogue movements, hydrography and bloom development. S. Afr. J. Mar. Sci./S.-Afr. Tydskr. Seewet., 5, 357–91.CrossRefGoogle Scholar
Burkill, P. H., Archer, S. D., Robinson, C., Nightingale, P. D., Groom, S. B., Tarran, G. A., and Zubkov, M. V., 2002. Dimethyl sulphide biogeochemistry within a coccolithophore bloom (DISCO): an overview. Deep-Sea Res. II., 49, 2863–85.CrossRefGoogle Scholar
Campana, S. E. and Thorrold, S. R., 2001. Otoliths, increments, and elements: keys to a comprehensive understanding of fish populations?Can. J. Fish. Aquat. Sci., 58, 30–38.CrossRefGoogle Scholar
Carter, H. H. and A. Okubo, 1978. A study of turbulent diffusion by dye tracers: a review. In Estuarine Transport Processes, ed. Kjerfve, B.. Columbia: Univ. South Carolina Press, 95–111.Google Scholar
Chisholm, S. W. and Morel, F. M. M., 1991. What controls phytoplankton production in nutrient-rich areas of the open ocean?Limnol. Oceanogr., 36, U1501–17.Google Scholar
Chisholm, S. W., Olson, R. J., Zettler, E. R., Waterbury, J., Goericke, R., and Welschmeyer, N., 1988. A novel free-living prochlorophyte occurs at high cell concentrations in the oceanic euphotic zone. Nature, 334, 340–3.CrossRefGoogle Scholar
Churchill, J. H., Hench, J. L., Luettich, R. A., Blanton, J. O., and Werner, F. E., 1999a. Flood tide circulation near Beaufort Inlet, North Carolina: Implications for larval recruitment. Estuaries, 22 (4), 1057–70.CrossRefGoogle Scholar
Churchill, J. H., Forward, R. B., Luettich, R. A., Hench, J. L., Hettler, W. F., Crowder, L. B., and Blanton, J. O., 1999b. Circulation and larval fish transport within a tidally dominated estuary. Fish. Oceanogr., 8 (suppl. 2), 173–89.CrossRefGoogle Scholar
Clark, C. W. and Levy, D. A., 1988. Diel vertical migrations by juvenile sockeye salmon and the anti-predation window. Am. Nat., 131, 271–90.CrossRefGoogle Scholar
Cleve, P. T., 1896. Microscopic marine diatoms in the service of hydrography. Nature, 55, no. 1413, 89–90.CrossRefGoogle Scholar
Coale, K. H., Johnson, K. S., Fitzwater, S. E., Gordon, R. M., Tanner, S., Chavez, F. P., Ferioli, L., Sakamoto, C. P., Rogers, P., Millero, F., Steinberg, P., Nightingale, P., Cooper, D., Cochlan, W. P., and Kudela, R., 1996. A massive phytoplankton bloom induced by an ecosystem-scale iron fertilization experiment in the Equatorial Pacific Ocean. Nature, 383, 495–501.CrossRefGoogle ScholarPubMed
Coale, K. H.et al., 2004. Southern Ocean Iron Enrichment Experiment: Carbon Cycling in High-and Low-Si Waters. Science, 304, 408–14.CrossRefGoogle ScholarPubMed
Cowen, R. and Castro, L. R., 1994. Relation of coral reef fish larval distributions to island scale circulation around Barbados, West Indies. Bull. Mar. Sci., 54 (1), 228–44.Google 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 (5454), 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. Gulf and Caribbean Res., 14, 129–37.CrossRefGoogle Scholar
Cronin, T. W. and R. B. Forward, 1982. Tidally timed behavior: effects on larval distributions in estuaries. In Estuarine Comparisons, ed. Kennedy, V. S.. New York: Academic Press, 505–20.Google Scholar
Dahlgren, C. P., Sobel, J. A., and Harper, D. E., 2001. Assessment of the reef fish community, habitat, and potential for larval dispersal from the proposed Tortugas South Ecological Reserve. Proc. Gulf Carib. Fish. Inst., 52, 700–12.Google Scholar
D'Asaro, E. A., Farmer, D. M., Osse, J. T., and Dairiki, G. F., 1996. A Lagrangian float. J. Atmosph. Ocean. Tech., 13, 1230–46.2.0.CO;2>CrossRefGoogle Scholar
Davis, A. R. and Butler, A. J., 1989. Direct observations of larval dispersal in the colonial ascidian Podoclavella moluccensis Sluiter: Evidence for closed populations. J. Exper. Mar. Biol. Ecol., 127(2), 189–203.CrossRefGoogle Scholar
Davis, R. E., 1985. Drifter observations of coastal surface currents during CODE: The statistical and descriptive view. J. Geophys. Res., 90, 4741–55.CrossRefGoogle Scholar
Davis, R. E., 1991. Lagrangian ocean studies. Annu. Rev. Fluid Mech., 23, 43–64.CrossRefGoogle Scholar
Davis, R. E., 2003. Neutrally buoyant floats. In ALPS: Autonomous and Lagrangian Platforms and Sensors, Workshop Report, ed. D. L. Rudnick and M. J. Perry.
Denman, K. L. and Gargett, A. E., 1983. Time and space scales of vertical mixing and advection of phytoplankton in the upper ocean. Limnol. Oceanogr., 28 (5), 801–15.CrossRefGoogle Scholar
Denman, K. L. and Powell, T. M., 1984. Effects of physical processes on planktonic ecosystems in the coastal ocean. Oceanogr. Mar. Biol. Ann. Rev., 22, 125–68.Google Scholar
Dennis, D. M., Pitcher, C. R., and Skewes, T. D., 2001. Distribution and transport pathways of Panulirus ornatus (Fabricius, 1776) and Panulirus spp. larvae in the Coral Sea, Australia. Mar. Freshwat. Res., 52 (8), 1175–85.CrossRefGoogle Scholar
Robertis, A. and Ohman, M. D., 1999. A free-drifting mimic of vertically migrating zooplankton. J. Plankton Res., 21(10), 1865–75.CrossRefGoogle Scholar
DiBacco, C. and Levin, L. A., 2000. Development and application of elemental fingerprinting to track the dispersal of marine invertebrate larvae. Limnol. Oceanogr., 45 (4), 871–80.CrossRefGoogle Scholar
Dickey, T. D., 1991. The emergence of concurrent high-resolution physical and bio-optical measurements in the upper ocean and their application. Rev. Geophys., 29, 383–413.CrossRefGoogle Scholar
Domeier, M., 2004. A potential larval recruitment pathway originating from a Florida marine protected area. Fisheries Oceanogr., 13, 287–94.CrossRefGoogle Scholar
Dooley, H. D., 1974. A comparison of drogue and current meter measurements in shallow waters. Rapports et Procès-verbaux des Réunions Conseil International pour l'Eploration de la Mer, 167, 225–30.Google Scholar
Epifanio, C. E., 1988. Transport of invertebrate larvae between estuaries and the continental shelf. American Fisheries Society Symposium, 3, 104–14.Google Scholar
Epifanio, C. E., and Garvine, R. W., 2001. Larval transport on the Atlantic Continental Shelf of North America: a Review. Estuar., Coast. and Shelf Sci., 52(1), 51–77.CrossRefGoogle Scholar
Falkowski, P. G., Katz, M. E., Knoll, A. H., Quigg, A., Raven, J. A., Schofield, O., and Taylor, F. J. R., 2004. The evolution of modern eukaryotic phytoplankton. Science, 305, 354–60.CrossRefGoogle ScholarPubMed
Fortier, L. and Leggett, W. C., 1982. Fickian transport and the dispersal of fish larvae in estuaries. Can. J. Fish. Aquat. Sci., 39 (8), 1150–63.CrossRefGoogle Scholar
French, D. P., Furnas, M. J., and Smayda, T. J., 1983. Diel changes in nitrite concentration in the chlorophyll maximum in the Gulf of Mexico. Deep-Sea Res., 30 (7A), 707–21.CrossRefGoogle Scholar
Furnas, M. J. and Smayda, T. J., 1987. Inputs of subthermocline waters and nitrate onto the Campeche bank. Cont. Shelf Res., 7 (2), 161–75.CrossRefGoogle Scholar
Gagnon, M. and Lacroix, G., 1981. The effects of tidal advection and mixing on the statistical dispersion of zooplankton. J. Exp. Mar. Biol. Ecol., 56(1), 9–22.CrossRefGoogle Scholar
Garland, E. D. and C. A. Zimmer, 2002. Hourly variations in planktonic larval concentrations on the inner shelf: Emerging patterns and processes. J. Mar. Res., 60(2), 311–25.CrossRefGoogle Scholar
Garrett, C., 1983. On the initial streakiness of a dispersing tracer in two- and three-dimensional turbulence. Dyn. Atmos. Oceans, 7 (4), 265–77.CrossRefGoogle Scholar
Garvine, R. W., Epifanio, C. E., Epifanio, C. C., and Wong, K.-C., 1997. Transport and recruitment of blue crab larvae: A model with advection and mortality. Estuar. Coast. Shelf Sci., 45 (1), 99–111.CrossRefGoogle Scholar
Geyer, W., 1989. Field calibration of mixed layer drifters. J. Atmosph. Ocean. Tech., 6, 333–42.2.0.CO;2>CrossRefGoogle Scholar
Gillanders, B. M. and Kingsford, M. J., 2003. Spatial variation in elemental composition of otoliths of three species of fish (family Sparidae). Estuar. Coast. Shelf Sci., 57 (5–6), 1049–64.CrossRefGoogle Scholar
Glibert, P. M., Garside, C., Fuhrman, J. A., and Roman, M. R., 1991. Time-dependent coupling of inorganic and organic nitrogen uptake and regeneration in the plume of the Chesapeake Bay estuary and its regulation by large heterotrophs. Limnol. Oceanogr., 36 (3), 895–909.CrossRefGoogle Scholar
Glibert, P. M. and Garside, C., 1992. Diel variability in nitrogenous nutrient uptake by phytoplankton in the Chesapeake Bay plume. J. Plankton Res., 14 (2), 271–88.CrossRefGoogle Scholar
Gran, H. H., 1912. Pelagic plant life. In The Depths of the Ocean, ed. Murray, J. and Hjort, J.. London: Macmillan, Chapter VI.Google Scholar
Griffin, D. A., Wilkin, J. L., Chubb, C. F., Pearce, A. F., and Caputi, N., 2001. Ocean currents and the larval phase of Australian western rock lobster, Panulirus cygnus. Mar. Freshwater Res., 52, 1187–99.CrossRefGoogle Scholar
Gustafson, D. E. Jr., Stoecker, D. K., Johnson, M. D., Hukelem, W. F., and Sneider, K., 2000. Cryptophyte algae are robbed of their organelles by the marine ciliate Mesodinium rubrum. Nature, 405, 1049–52.CrossRefGoogle ScholarPubMed
Haeckel, E. 1890. Planktonstuden. Jena.
Hare, J. A., Churchill, J. H., Cowen, R. K., Berger, T. J., Cornillon, P. C., Dragos, P., Glenn, S. M., Govoni, J. J., and Lee, T. N., 2001. Routes and rates of larval fish transport from the southeast to the northeast United States continental shelf. Limnol. Oceanogr., 47(6), 1774–89.CrossRefGoogle Scholar
Hare, J., Quinlan, J., Werner, F., Blanton, B., Govoni, J., Forward, R., Settle, C., and Hoss, D., 1999. Larval transport during winter in the SABRE study area: results of a coupled vertical larval behavior three-dimensional circulation model. Fish. Oceanogr., 8(S2), 57–76.CrossRefGoogle Scholar
Haury, L. R., 1976a. A comparison of zooplankton patterns in the California Current and North Pacific Central Gyre. Mar. Biol., 37, 159–67.CrossRefGoogle Scholar
Haury, L. R., 1976b. Small-scale pattern of a California current zooplankton assemblage. Mar. Biol., 37(2), 137–57.CrossRefGoogle Scholar
Haury, L. A., J. A. McGowan, and P. H. Wiebe, 1977. Patterns and processes in the time-space of plankton distributions. In Spatial Pattern in Plankton Communities, ed. Steele, J. H.. New York: Plenum Press, 277–327.Google Scholar
Heath, M. and Rankine, P., 1988. Growth and advection of larval herring (Clupea harengus L.) in the vicinity of the Orkney Isles. Est. Coastal Shelf Sci., 27, 547–65.CrossRefGoogle Scholar
Hensen, V., 1887. Über die Bestimmung des Plankton oder des im Meer triebenden Materials an Pflanzen und Thieren. Fünfter Ber. Komm. Wiss. Unters. Deut. Meer Kiel., 1–108.Google Scholar
Hitchcock, G. L., Lessard, E. J., Dorson, D., Fontaine, J., and Rossby, T., 1989. The IFF: The isopycnal float fluorometer. J. Atmosph. Ocean. Tech., 6, 17–26.2.0.CO;2>CrossRefGoogle Scholar
Hitchcock, G. L., Olson, D. B., Knauer, G. A., Pszenny, A. A. P., and Galloway, J. N., 1990. Horizontal diffusion and new production in the Sargasso Sea. Global Biogeochem. Cycles, 4, 253–65.CrossRefGoogle Scholar
Hitchcock, G. L., Mariano, A., and Rossby, T, 1993. Mesoscale pigment fields in the Gulf Stream: Observations in a meander crest and trough. J. Geophys. Res., 98 (C5), 8425–45.CrossRefGoogle Scholar
Hitchcock, G. L., Rossby, T., Lillibridge, J. L, Lessard, E. J.III,, Levine, E. R., Connors, D. N., Børsheim, K. Y., and Mork, M., 1994. Signatures of stirring and mixing near the Gulf Stream front. J. Mar. Res., 52 (5), 797–836.CrossRefGoogle Scholar
Hitchcock, G. L., Wiseman, W. J. Jr., Boicourt, W. C., Mariano, A. J., Walker, N., Nelson, T. A., and Ryan, E., 1997. Property fields in an effluent plume of the Mississippi River. J. Mar. Sys., 12(2), 109–26.CrossRefGoogle Scholar
Hitchcock, G. L., Vargo, G. A., and Dickson, M. L., 2000a. Plankton community composition, production, and respiration in relation to dissolved inorganic carbon on the west Florida Shelf, April, 1996. J. Geophys. Res., 105 (C3), 6579–89.CrossRefGoogle Scholar
Hitchcock, G. L., Key, E., and Masters, J., 2000b. The fate of upwelled waters in the Great Whirl, August, 1995. Deep-Sea Res. II., 47 (7–8), 1605–21.CrossRefGoogle Scholar
Hitchcock, G. L., Chen, R. F., Gardner, G. B., and Wiseman, W. J. Jr., 2004. A Lagrangian view of fluorescent chromophoric dissolved organic matter distributions in the Mississippi River plume. Marine Chemistry, 89, 225–39.CrossRefGoogle Scholar
Hofmann, E. E., Hedström, K. S., Moisan, J. R., Haidvogel, D. B., and Mackas, D. L., 1991. Use of simulated drifter tracks to investigate general transport patterns and residence times in the coastal transition zone. J. Geophys. Res., 96 (C8), 15,041–52.CrossRefGoogle Scholar
Houghton, R. W. and Ho, C., 2001. Diapycnal flow through the Georges Bank tidal front: A dye tracer study. Geophys. Res. Lett., 28(1), 33–6.CrossRefGoogle Scholar
Huisman, J., Arrayás, M., Ebert, U., and Sommeijer, B., 2002. How do sinking phytoplankton species manage to persist?Amer. Natural., 159(3), 245–54.CrossRefGoogle ScholarPubMed
Hutchins, J. B. and Pearce, A. F., 1994. Influence of the Leeuwin Current on recruitment of tropical reef fishes at Rottnest Island, Western Australia. Bull. Mar. Science, 54, 245–55.Google Scholar
Itoh, H., Hirata, T. C., Tanaka, M., Inoue, N., and Irie, H., 1979. Short-term fluctuations of zooplankton community in company with movement of a curtain drogue-I. Fluctuations of copepod community neighboring a current drogue followed by YŌKŌ MARU. Bull. Seikai Reg. Fish. Res. Lab., 53, 113–23.Google Scholar
James, M. K., Armstrong, P. R., Mason, L. B., and Bode, L., 2002. The structure of reef fish metapopulations: modeling larval dispersal and retention patterns. Proc. R. Soc. Lond. B, 269, 2079–86.CrossRefGoogle Scholar
Jennings, F. D., 1981. The Coastal Upwelling Ecosystems Analysis program. Epilogue. IDOE Int. Symp. on Coastal Upwelling, Los Angeles, CA. Texas A&M Univ., Sea Grant Cent. for Marine Research, College Station, TX, 13–15.
Johnson, D. R., 1995. Wind forced surface currents at the entrance to Chesapeake Bay: Their effect on blue crab larval dispersion and post-larval recruitment. Bull. Mar. Sci., 57(3), 726–38.Google Scholar
Johnson, K., 2003. Biogecochemcial cycles. In ALPS: Autonomous and Lagrangian Platforms and Sensors, Workshop Report, ed. D. L. Rudnick and M. J. Perry. 22–5.
Jones, B. H., K. H. Brink, R. C. Dugdale, D. W. Stuart, J. C. Van Leer, D. Blasco, and J. C. Kelly, 1983. Observations of a persistent upwelling center off Point Conception, California. In Coastal Upwelling: Its Sediment Record, ed. Suess, S. and Thiede, J.. New York: Plenum Press, 37–60.CrossRefGoogle Scholar
Jones, G. P., Milicich, M. J., Emslie, M. J., and Lunow, C., 1999. Self-recruitment in a coral reef fish population. Nature, 402, 802–4.CrossRefGoogle Scholar
Jones, T. W., Malone, T. C., and Pike, S., 1990. Seasonal contrasts in diurnal carbon incorporation by phytoplankton size classes of the coastal plume of Chesapeake Bay. Mar. Ecol. Prog. Ser., 68, 1–21.CrossRefGoogle Scholar
Joseph, J. and Sender, H., 1958. Uber die horizontale Diffusion im Meere. Dt. Hydrogr. Z., 11, 49–77.CrossRefGoogle Scholar
Kamykowski, D., 1995. Trajectories of autotrophic marine dinoflagellates. J. Phycol., 31, 200–8.CrossRefGoogle Scholar
Kamykowski, D., Reed, R. E., and Kirkpatrick, G. J., 1992. Comparison of sinking velocity, swimming velocity, rotation, and path characteristics among six marine dinoflagellate species. Mar. Biol., 113 (2), 319–28.Google Scholar
Kennedy, V. S. and Boicourt, W. C., 1981. Water circulation and oyster spat settlement in two adjacent tributaries of the Choptank River, Maryland. J. Shellfish Res., 1 (1)118.Google Scholar
Ketchum, B. H. and Corwin, N., 1965. The cycle of phosphorus in a plankton bloom in the Gulf of Maine. Limnol. Oceanogr., 10 (suppl.), R148–61.CrossRefGoogle Scholar
Kirkpatrick, G. J., Curtin, T. B., Kamykowski, D., Freezor, M. D., Sartin, M. D., and Reed, R. E., 1990. Measurement of photosynthetic response to euphotic zone physical forcing. Oceanography, 3(1), 18–22.CrossRefGoogle Scholar
Kirwan, A. D., McNally, G., Chang, M.-S., and Molinari, R., 1975. The effect of wind and surface currents on drifters. J. Phys. Oceanogr., 5, 361–8.2.0.CO;2>CrossRefGoogle Scholar
Koehl, M. A. R. and T. M. Powell, 1994. Turbulent transport of larvae near wave-swept rocky shores: Does water motion overwhelm larval sinking? In Reproduction and Development of Marine Invertebrates, ed. Wilson, W. H. Jr., Stricker, S. A., and Shinn, G. L.. Baltimore: The Johns Hopkins University Press, 261–74.Google Scholar
Laane, R. W. P. M., Manuels, M. W., and Staal, W., 1984. A procedure for enriching and cleaning up Rhodamine B and Rhodamine WT in natural waters. Water Res., 18 (2), 163–5.CrossRefGoogle Scholar
Landry, M. R., Brown, S. L., Selph, K. E., Abbott, M. R., Letelier, R. M., Christensen, S., Bidigare, R. R., and Casciotti, K., 2001. Initiation of the spring phytoplankton increase in the Antarctic Polar Front Zone at 170 degree W. J. Geophys. Res., 106 (C7), 13,903–15.CrossRefGoogle Scholar
Largier, J. L., 2003. Considerations in estimating larval dispersal distances from oceanographic data. Ecol. Appl., 13(1), S71–89.CrossRefGoogle Scholar
Law, C. S., Abraham, E. R., Watson, A. J., and Liddicoat, M. I., 2003. Vertical eddy diffusion and nutrient supply to the surface mixed layer of the Antarctic Circumpolar Current. J. Geophys. Res., 108 (C8).CrossRefGoogle Scholar
Law, C. S., Liddicoat, M. I., Martin, A. P., Richards, K. J., and Woodward, E. M. S., 2000. A Lagrangian SF6 tracer study of an anticyclonic eddy in the North Atlantic: patch evolution, vertical mixing and nitrate supply to the mixed layer. Deep-Sea Res. II., 48 (4–5), 705–24.Google Scholar
Borgne, R. P., 1978. Ammonium formation in Cape Timiris (Mauritania) upwelling. J. Exp. Mar. Biol. Ecol., 31(3), 253–65.CrossRefGoogle Scholar
Ledwell, J. R., Watson, A. J., and Law, C. S., 1998. Mixing of a tracer in the pycnocline. J. Geophys. Res., 103 (C10): 21,499–529.CrossRefGoogle Scholar
Lee, T. N., Clarke, M. E., Williams, E. W., Szmant, A. F., and Berger, T., 1994. Evolution of a Tortugas gyre and its influence on recruitment in the Florida Keys. Bull. Mar. Sci., 54, 621–46.Google Scholar
Leis, J. M., 1991a. The pelagic stage of reef fishes: The larval biology of coral reef fishes. InThe ecology of fishes on coral reefs, ed. Sale, P. F.. San Diego: Academic Press, 183–230.Google Scholar
Leis, J. M., 1991b. Vertical distribution of fish larvae in the Great Barrier Reef Lagoon, Australia. Mar. Biol., 109, 157–66.CrossRefGoogle Scholar
Letelier, R. M., Abbott, M. R., and Karl, D. M., 1995. Southern ocean optical drifter experiment. Antarct. J. U.S., 30(5), 108–10.Google Scholar
Letelier, R. M., Abbott, M. R., and Karl, D. M., 1997. Chlorophyll natural fluorescence response to upwelling events in the Southern Ocean. Geophys Res. Letts., 24(4), 409–12.CrossRefGoogle Scholar
Levin, L. A., 1990. A review of methods for labeling and tracking marine invertebrate larvae. Ophelia., 32 (1–2), 115–44.CrossRefGoogle Scholar
Liu, G, Janowitz, G. S., and Kamykowski, D., 2002. Influence of current shear on Gymnodinium breve (Dinophyceae) population dynamics: a numerical study. Mar. Ecol. Prog. Ser., 231, 47–66.CrossRefGoogle Scholar
Lohrenz, S. E., Cullen, J. J., Phinney, D. A., Olson, D. B., and Yentsch, C. S., 1993. Distributions of pigments and primary production in a Gulf Stream meander. J. Geophys. Res., 98 (C8), 14,545–55.CrossRefGoogle Scholar
Lorenzen, C. J., 1968. Carbon/Chlorophyll relationships in an upwelling area. Limnol. Oceanogr., 13, 202–4.CrossRefGoogle Scholar
MacIsaac, J. J., Dugdale, R. C., Barber, R. T., Blasco, D., and Packard, T. T., 1985. Primary production cycle in an upwelling center. Deep-Sea Res., 32 (5), 503–29.CrossRefGoogle Scholar
Malone, T. C. and Ducklow, H. W., 1990. Microbial biomass in the coastal plume of Chesapeake Bay: Phytoplankton-bacterioplankton relationships. Limnol. Oceanogr., 35 (2), 296–312.CrossRefGoogle Scholar
Martin, J. H., 1990. Glacial-interglacial CO2 change: The iron hypothesis. Paleoceanography, 3, 1–13.CrossRefGoogle Scholar
Martin, J. H., Coale, K. H., Johnson, K. S., Fitzwater, S. E., Gordon, R. M., Tanner, S. J., Hunter, C. N., Elrod, V. A., Nowicki, J. L., Coley, T. L., Barber, R. T., Lindley, S., Watson, A. J., Scoy, K., and Law, C. S., 1994. Testing the iron hypothesis in ecosystems of the Equatorial Pacific Ocean. Nature, 371, 123–9.CrossRefGoogle Scholar
Martin, A. P., Wade, I. P., Richards, K. J., and Heywood, K. J., 1998. The PRIME eddy. J. Mar. Res., 56, 439–62.CrossRefGoogle Scholar
McGehee, D. and Jaffe, J. S., 1996. Three-dimensional swimming behaviour of individual zooplankters: observations using the acoustical imaging system FishTV. ICES J. Mar. Sci., 53 (2), 363–9.CrossRefGoogle Scholar
Miller, C. B., 1970. Some environmental consequences of vertical migration in marine zooplankton. Limnol. Oceanogr., 15 (5), 727–41.CrossRefGoogle Scholar
Mitchell, B. G., M. Kahru, and J. Sherman, 2000. Autonomous temperature-irradiance profiler resolves the spring bloom in the Sea of Japan. Proceedings Ocean Optics XV, Monaco, Oct 2000.
Moloney, C. L. and Field, J. G., 1991. The size-based dynamics of plankton food webs. 1. A simulation model of carbon and nitrogen flows. J. Plankton Res., 13 (5), 1003–38.CrossRefGoogle Scholar
Morel, A., Ahn, Y.-W., Partensky, F., Vaulot, D., and Claustre, H., 1993. Prochlorococcus and Synechococcus: A comparative study of their optical properties in relation to their size and pigmentation. J. Mar. Res., 51 (3), 617–49.CrossRefGoogle Scholar
Morgan, S. G., 1995. Life and death in the plankton: larval mortality and adaptation. In Ecology of Marine Invertebrate Larvae, ed. McEdward, L.. Boca Raton: CRC Press, 279–321.Google Scholar
Moriarty, D. J. W., 1979. Biomass of suspended bacteria over coral reefs. Mar. Biol., 53(2), 193–200.CrossRefGoogle Scholar
Munk, P., Wright, P. J., and Pihl, N. J., 2002. Distribution of the early larval stages of cod, plaice and lesser sandeel across haline fronts in the North Sea. Estuar. Coast. Shelf Sci., 55 (1), 139–49.CrossRefGoogle Scholar
Nakata, H. and Hirano, T., 1978. Dye-diffusion experiments in a narrow passage and approaches. Bull. Jap. Soc. Fish. Oceanogr., 32, 1–14.Google Scholar
Natunewicz, C. C., Epifanio, C. E., and Garvine, R. W., 2001. Transport of crab larval patches in the coastal ocean. Mar. Ecol. Prog. Ser., 222, 143–54.CrossRefGoogle Scholar
Niiler, P. P., Davis, R. E., and White, H. J., 1987. Water-following characteristics of a mixed layer drifter. Deep-Sea Res. I., 34(11), 1867–81.CrossRefGoogle Scholar
Ohman, M. D., 1988. Behavioral responses of zooplankton to predation. Bull. Mar. Sci., 43, 530–50.Google Scholar
Okubo, A., Hasegawa, S., Amano, M., and Takeda, I., 1957. Report of the observation concerning the diffusion of dye patch in the sea off the coast of Tokai-mura. Research papers Japan Atomic Energy res. Inst., 2, 17–21.Google Scholar
Okubo, A., 1971. Oceanic diffusion diagrams. Deep-Sea Res., 18 (8), 789–802.Google Scholar
Okubo, A., 1994. The role of diffusion and related physical processes in dispersal and recruitment of marine populations. In The Biophysics of Marine Larval Dispersal, ed. Sammarco, P. and Heron, M. L.. Coastal Estuarine Studies, 45. New York: Springer Verlag, 5–31.CrossRefGoogle Scholar
Olaizola, M., LaRoche, J., Kolber, Z., and Falkowski, P. G., 1994. Non-photochemical fluorescence quenching and the diadinoxanthin cycle in a marine diatom. Photosynth. Res., 41, 357–70.CrossRefGoogle Scholar
Olson, R. R., 1985. The consequences of short-distance larval dispersal in a sessile marine invertebrate. Ecology, 66(1), 30–9.CrossRefGoogle Scholar
Oudot, C., Raul, P., and Wauthy, B., 1979. Western Pacific equatorial upwelling: physical and chemical distributions and standing crop following a drifting drogue. Cahiers Indo-Pac., 1(1), 39–81.Google Scholar
Palumbi, S. R., 2001. The ecology of marine protected areas. In Marine Ecology: the New Synthesis, ed. Bertness, M., Gaines, S. D., and Hay, M. E.. Sunderland, MA: Sinauer, 509–30.Google Scholar
Paris, C. B. and Cowen, R. K., 2004. Direct evidence of a biophysical retention mechanism for coral reef fish larvae. Limnol. Oceanogr., 49(6), 1964–79.CrossRefGoogle Scholar
Paris, C. B., Cowen, R. K., Lwiza, K. M. M., Wang, D.-P., and Olson, D. B., 2002. Multivariate objective analysis of the coastal circulation of Barbados, West Indies: implication for larval transport. Deep-Sea Res. I., 49, 1363–86.CrossRefGoogle Scholar
Parker, G. G. Jr., 1973. Tests of rhodamine WT dye for toxicity to oysters and fish. J. Res. U. S. Geol. Surv., 1 (4), 499.Google Scholar
Pelegrí, J. L. and Csanady, G. T., 1990. Nutrient transport and mixing in the Gulf Stream. J. Geophys. Res., 96 (C2), 2577–83.CrossRefGoogle Scholar
Pepin, P. and Helbig, J. A., 1997. Distribution and drift of Atlantic cod (Gadus morhua) eggs and larvae on the Northeast Newfoundland Shelf. Can. J. Fish. Aquat. Sci., 54 (03), 670–85.CrossRefGoogle Scholar
Price, H. J., 1989. Swimming behavior of krill in response to algal patches: A mesocosm study. Limnol. Oceanogr., 34 (4), 649–59.CrossRefGoogle Scholar
Raabe, T. U., Brockmann, U. H., Duerselen, C.-D., Krause, M., and Rick, H. J., 1997. Nutrient and plankton dynamics during a spring drift experiment in the German Bight. Mar. Ecol. Prog. Ser., 156, 275–88.CrossRefGoogle Scholar
Raven, J. A., 1986. Physiological consequences of extremely small size for autotrophic organisms in the sea. In Photosynthetic picoplankton. (Can. Bull. Fish. Aquat. Sci. 214), ed. Platt, T. and Li, W. K. W.. Canadian Government Publishing Centre, 1–70.Google Scholar
Raven, J. A. and Richardson, K., 1984. Dinophyte flagella: A cost-benefit analysis. New Phytol., 98(2), 259–76.CrossRefGoogle Scholar
Rick, S., 1999. The spring bloom in the German Bight: Effects of high inorganic N:P ratios on the phytoplankton development. 305. Berichte aus dem Institut für Meereskunde an der Christian-Albrechts-Universität Kiel.
Roberts, C. M., 1997. Connectivity and management of Caribbean coral reefs. Science, 278, 1454–7.CrossRefGoogle ScholarPubMed
Roman, M. R. and Boicourt, W. C., 1990. Temporal and spatial variations in the abundance of blue crab larvae in the Chesapeake Bay plume and surrounding shelf waters. Bull. Mar. Sci., 46 (1), 249–50.Google Scholar
Roman, M. R. and Boicourt, W. C., 1999. Dispersion and recruitment of crab larvae in the Chesapeake Bay plume: physical and biological controls. Estuaries, 22 (3A), 563–74.CrossRefGoogle Scholar
Rossby, T., Bower, A. S., and Shaw, P.-T., 1985. Particle pathways in the Gulf Stream. Bull. Am. Meteorol. Soc., 66 (9), 1106–10.2.0.CO;2>CrossRefGoogle Scholar
Rossby, T., Dorson, D., and Fontaine, J., 1986. The RAFOS system. J. Atmos. Oceanic Techn., 3, 672–9.2.0.CO;2>CrossRefGoogle Scholar
Rossby, T. and Webb, D., 1970. Observing abyssal motion by tracking Swallow floats in the SOFAR channel. Deep-Sea Res., 17, 359–65.Google Scholar
Rudnick, D. L., and M. J. Perry, eds, 2003. ALPS: Autonomous and Lagrangian Platforms and Sensors, Workshop Report. www.geo-prose.com/ALPS
Ruiz, J., Garcia, C. M., and Rodriguez, J., 1996. Sedimentation loss of phytoplankton cells from the mixed layer: Effects of turbulence levels. J. Plankton Res., 18(9), 1727–34.CrossRefGoogle Scholar
Rumrill, S. S., 1991. Natural mortality of marine invertebrate larvae. Ophelia, 32, 163–98.CrossRefGoogle Scholar
Ryther, J. H., D. W. Menzel, E. M. Hulburt, et al. 1971. Production and utilization of organic matter in the Peru coastal current. Anton Bruun Report No. 4. Texas A&M University. College Station, TX.
Sambrotto, R. N. and Langdon, C., 1994. Water column dynamics of dissolved inorganic carbon (DIC), nitrogen and O2 on Georges Bank during April, 1990. Continental Shelf Research, 14 (7/8), 765–89.CrossRefGoogle Scholar
Savidge, G. and Williams, P. J. le B, 2001. The PRIME 1996 cruise: an overview. Deep-Sea Res. II. Top. Stud. Oceanogr., 48 (4–5), 687–704.CrossRefGoogle Scholar
Scheltema, R. S., 1986. On dispersal and planktonic larvae of benthic invertebrates: An eclectic overview and summary of problems. Bull. Mar. Sci., 39, 290–322.Google Scholar
Schlee, S., 1973. The Edge of an Unfamiliar World: A History of Oceanography. New York: E. P. Dutton & Co., Inc.Google Scholar
Seligman, H., 1955. The discharge of radioactive waste products into the Irish Sea. Proceedings of the International Conference of Peaceful Uses of Atomic Energy. Geneva. 701–11.
Shanks, A. L., Largier, J., Brink, L., Brubaker, J., and Hooff, R., 2000. Demonstration of the onshore transport of larval invertebrates by the shoreward movement of an upwelling front. Limnol. Oceanogr., 45(1), 230–6.CrossRefGoogle Scholar
Shanks, A. L., Grantham, B. A., and Carr, M. H., 2003. Propagule dispersal distance and the size and spacing of marine reserves. Ecol. Applications, 13(1), S159–69.CrossRefGoogle Scholar
Sieburth, J. McN., Smetacek, V., and Lenz, J., 1978. Pelagic ecosystem structure: Heterotrophic components of the plankton and their relationship to plankton size-fractions. Limnol. Oceanogr., 23, 1256–63.CrossRefGoogle Scholar
Siegel, D. A., Kinlan, B. P., Gaylord, B., and Gaines, S. D., 2003. Lagrangian descriptions of marine larval dispersion. Mar. Ecol. Prog. Ser., 260, 83–96.CrossRefGoogle Scholar
Smayda, T. J., 1970. The suspension and sinking of phytoplankton in the sea. Oceanogr. Mar. Biol. Ann. Rev., 8, 353–414.Google Scholar
Stoner, D. S., 1990. Recruitment of a tropical colonial ascidian: Relative importance of pre-settlement vs. post-settlement processes. Ecology, 71(5), 1682–90.CrossRefGoogle Scholar
Strathmann, R. R., 1990. Why life histories evolve differently in the sea. Am. Zool., 30 (1), 197–207.CrossRefGoogle Scholar
Strickler, J. R., Squires, K. D., Yamakazi, H., and Abib, A. H., 1997. Combining analog turbulence with digital turbulence. Sci. Mar. (Barc.)., 61 (1), 197–204.Google Scholar
Suijlen, J. M. and Buyse, J. J., 1994. Potentials of photolytic rhodamine WT as a large-scale water tracer assessed in a long-term experiment in the Loosdrecht Lakes. Limnol. Oceanogr., 39, 1411–23.CrossRefGoogle Scholar
Swallow, J. C., 1955. A neutral-buoyancy float for measuring deep currents. Deep-Sea Res., 3, 74–81.Google Scholar
Swearer, S. E., Caselle, J. E., Lea, D. W., and Warner, R. R., 1999. Larval retention and recruitment in an island population of a coral-reef fish. Nature, 402 (6763), 799–802.CrossRefGoogle Scholar
Taggart, C. T. and Leggett, W. C., 1987a. Short-term mortality in post-emergent larval capelin Mallotus villosus. 1. Analysis of multiple in situ estimates. Mar. Ecol. Prog. Ser., 41 (3), 205–17.CrossRefGoogle Scholar
Taggart, C. T. and Leggett, W. C., 1987b. Short-term mortality in post-emergent larval capelin Mallotus villosus. 2. Importance of food and predator density, and density-dependence. Mar. Ecol. Prog. Ser., 41 (3), 219–29.CrossRefGoogle Scholar
Talbot, J. W. and Talbot, G. A., 1974. Diffusion in shallow seas and in English coastal and estuarine waters. Rapp. P.-v. Réun. Cons. Int. Explor. Mer., 167, 93–110.Google Scholar
Talbot, J. W., 1977. The dispersal of plaice eggs and larvae in the Southern Bight of the North Sea. J. Cons. Int. Explor. Mer., 37(3), 221–48.CrossRefGoogle Scholar
Taylor, A. H., Harbour, D. S., Harris, R. P., Burkill, P. H., and Edwards, E. S., 1993. Seasonal succession in the pelagic ecosystem of the North Atlantic and the utilization of nitrogen. J. Plankton Res., 15 (8), 875–91.CrossRefGoogle Scholar
Taylor, F. J. R., 1980. Phytoplankton ecology before 1900: Supplementary notes to the “Depths of the Ocean”. In Oceanography: The Past. ed. Sears, M. and Merriman, D.. New York: Springer-Verlag, 509–21.CrossRefGoogle Scholar
Tegner, M. J. and Butler, R. A., 1985. Drift-tube study of the dispersal potential of green abalone (Haliotus fulgens) larvae in the Southern California Bight – Implications for recovery of depleted populations. Mar. Ecol. Prog. Ser., 26, 73–84.CrossRefGoogle Scholar
Thorrold, S. R. and J. A. Hare, 2002. Otolith applications in reef fish ecology. In Coral Reef Fishes: Dynamics and Diversity in a Complex Ecosystem, ed. Sale, P. F.. New York: Academic Press, 243–64.Google Scholar
Thorrold, S. R., Latkoczy, C., Swart, P. K., and Jones, C. M., 2001. Natal homing in a marine fish metapopulation. Science, 291 (5502), 297–9.CrossRefGoogle Scholar
Thorrold, S. R., Jones, G. P., Hellberg, M. E., Burton, R. S., Swearer, S. E., Neigel, J. E., Morgan, S. G., and Warner, R. R., 2002. Quantifying larval retention and connectivity in marine populations with artificial and natural markers. Bull. Mar. Sci., 70 (1), 291–308.Google Scholar
Tsuda, A.et al., 2003. A mesoscale iron enrichment in the western Subarctic Pacific induces a large centric diatom bloom. Science, 300, 958–61.CrossRefGoogle ScholarPubMed
Upstill-Goddard, R. C., Suijlen, J. M., Malin, G., and Nightingale, P. D., 2001. The use of photolytic Rhodamines WT and sulpho G as conservative tracers of dispersion in surface waters. Limnol. Oceanogr., 46(4), 927–34.CrossRefGoogle Scholar
Villareal, T. A. and Carpenter, E. J., 1994. Chemical composition and photosynthetic characteristics of Ethmodiscus rex (Bacillariophyceae): evidence for vertical migration. J. Phycology, 30, 1–8.CrossRefGoogle Scholar
Walsh, J. J., Weisberg, R. H., Dieterle, D. A., He, R., Darrow, B. P., Jolliff, J. K., Lester, K. M., Vargo, G. A., Kirkpatrick, G. J., Fanning, K. A., Sutton, T. T., Jochens, A. E., Biggs, D. C., Nababan, B., Hu, C., and Muller-Karger, F. E., 2003. Phytoplankton response to intrusions of slope water on the West Florida Shelf: Models and observations. J. Geophys. Res., 108 (C6), 3190 10.1029/2002JC001406.CrossRefGoogle Scholar
Wanninkhof, R., Ledwell, J. R., and Broecker, W. S., 1985. Gas-exchange wind-speed relation measured with sulfur-hexafluoride on a lake. Science, 227, 1224–26.CrossRefGoogle ScholarPubMed
Wanninkhof, R., Asher, W., Weppernig, R., Chen, H., Schlosser, P., Langdon, C., and Sambrotto, R., 1993. Gas transfer experiment on Georges Bank using two volatile deliberate tracers. J. Geophys. Res., 98 (C11), 20,237–48.CrossRefGoogle Scholar
Wanninkhof, R., Hitchcock, G. L., Wiseman, W. J., Vargo, G., Ortner, P. B., Asher, W., Ho, D. T., Schlosser, P., Dickson, M.-L., Masserini, R., Fanning, K., and Zhang, J.-Z., 1997. Gas exchange, dispersion and biological productivity on the west Florida Shelf: results from a Lagrangian tracer study. Geophys. Res. Letts., 24 (14), 1767–70.CrossRefGoogle Scholar
Watson, A. J. and Ledwell, J. R., 2000. Oceanographic tracer release experiments using sulphur hexafluoride. J. Geophys. Res., 105 (C6), 14,325–37.CrossRefGoogle Scholar
Watson, A. J., Ledwell, J. R., and Sutherland, S. C., 1991. The Santa Monica Basin tracer experiment: Comparison of release methods and performance of perfluorodecalin and sulfur hexafluoride. J. Geophys. Res., 96 (C5), 8719–25.CrossRefGoogle Scholar
Watson, A. J., Law, C. S., Scoy, K. A., Millero, F. J., Yao, W., Friedderich, G. E., Liddicoat, M. I., Wanninkhof, R. H., Barber, R. T., and Coale, K. H., 1994. Minimal effect of iron fertilization on sea-surface carbon dioxide concentrations. Nature, 371, 143–5.CrossRefGoogle Scholar
Werner, F. E., Blanton, B. O., Quinlan, J. A., and Luettich, R. A. Jr., 1999. Physical oceanography of the North Carolina continental shelf during the fall and winter seasons: implications for the transport of larval menhaden. Fish. Oceanogr., 8(S2), 7–21.CrossRefGoogle Scholar
Wolcott, T. G. and D. L. Wolcott, 1998. Estuarine export and egress of larvae: test with larval mimics. Abstracts of the Annual Meeting of the Society for Integrative and Comparative Biology. Boston MA.
Wyatt, B., Burt, W. V., and Pattullo, J. G., 1972. Surface currents off Oregon as determined from drift bottle returns. J. Phys. Oceanogr., 2, 286–93.2.0.CO;2>CrossRefGoogle Scholar
Yeung, C. and Lee, T. N., 2002. Larval transport and retention of the spiny lobster, Panulirus argus, in the coastal zone of the Florida Keys, USA. Fish. Oceanogr., 11(5), 286–309.CrossRefGoogle Scholar
Zaret, T. M. and Suffern, J. S, 1976. Vertical migration in zooplankton as a predator avoidance mechanism. Limnol. Oceanogr., 21(6), 804–13.CrossRefGoogle Scholar
Zenitani, H., Nakata, K., and Kimura, R., 1996. Survival and growth of sardine larvae in the offshore side of the Kuroshio. Fish. Oceanogr., 5 (1), 56–62.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
×