Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-26T19:34:57.381Z Has data issue: false hasContentIssue false

Microzooplankton and meroplanktonic larvae at two seamounts in the subtropical and tropical NE Atlantic

Published online by Cambridge University Press:  20 January 2016

A. Denda*
Affiliation:
Institut für Hydrobiologie und Fischereiwissenschaft, Universität Hamburg, Große Elbstraße 133, 22767 Hamburg, Germany
C. Mohn
Affiliation:
Department of Bioscience, Aarhus University, Frederiksborgvej 399, 4000 Roskilde, Denmark
H. Wehrmann
Affiliation:
Institut für Hydrobiologie und Fischereiwissenschaft, Universität Hamburg, Große Elbstraße 133, 22767 Hamburg, Germany
B. Christiansen
Affiliation:
Institut für Hydrobiologie und Fischereiwissenschaft, Universität Hamburg, Große Elbstraße 133, 22767 Hamburg, Germany
*
Correspondence should be addressed to:A. Denda, Institut für Hydrobiologie und Fischereiwissenschaft, Universität Hamburg, Große Elbstraße 133, 22767 Hamburg, Germany email: anneke.denda@uni-hamburg.de

Abstract

Spatial distribution patterns of microzooplankton (0.055–0.3 mm) biomass and abundance were studied in relation to the hydrographic situation and the local flow field in the waters off Ampère and Senghor, two shallow seamounts in the subtropical and tropical NE Atlantic, in comparison with unaffected open ocean reference sites. Ampère was sampled during November/December 2010 and Senghor during December 2011 and February 2013. The study includes taxonomic composition, abundance of meroplanktonic larvae and an estimation of the respiratory carbon demand. Biomass (dry weight) standing stocks of microzooplankton in the upper 100 m ranged between 30–120 mg m−2 over Ampère and 140–260 mg m−2 over Senghor Seamount, corresponding to 33 and 24% of the total zooplankton (0.055–20 mm). Highest total abundance was always found in the upper 50 m with numbers of 1070–5060 Ind m−3 at Ampère and 5050–20,000 Ind m−3 at Senghor with microzooplankton contributing 70–95%. Zooplankton accumulated mainly at the thermocline coincident with the deep fluorescence maximum and was ascertained by food supply rather than by oxygen limitation. The microzooplankton contribution to the total respiratory carbon demand was ~50% in the subtropical waters off Ampère and ~30% at Senghor, reflecting the important role of microzooplankton in the waters of the NE Atlantic subtropical gyre. Clear evidence of local seamount effects resulting in enhanced microzooplankton biomass compared with the unaffected reference sites were not detected. However, we confirmed Senghor as a hotspot for meroplanktonic larvae, suggesting a retention potential that results in significantly enhanced larval abundance in the seamount waters as compared with the open ocean.

Type
Research Article
Copyright
Copyright © Marine Biological Association of the United Kingdom 2016 

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

REFERENCES

Arístegui, J., Hernández-León, S., Montero, M.F. and Gómez, M. (2001) The seasonal planktonic cycle in coastal waters of the Canary Islands. Scientia Marina 65(Suppl. 1), 5158.Google Scholar
Arnsberg, A.J. (2001) Arthropoda, Cirripedia: the barnacles. In Shanks, A.L. (ed.) An identification guide to the larval marine invertebrates of the Pacific Northwest. Corvallis, OR: Oregon State University Press, pp. 157177.Google Scholar
Beck, T., Metzger, T. and Freiwald, A. (2006) BIAS. Biodiversity inventorial atlas of macrobenthic seamount animals. OASIS Deliverable 25 Final Report, 125 pp.Google Scholar
Beckmann, A. and Mohn, C. (2002) The upper ocean circulation at Great Meteor Seamount. Part II: retention potential of the seamount induced circulation. Ocean Dynamics 52, 194204.Google Scholar
Boehlert, G.W. and Mundy, B.C. (1993) Ichthyoplankton assemblages at seamounts and oceanic islands. Bulletin of Marine Science 53, 336361.Google Scholar
Böttger-Schnack, R. (1992) Community structure and vertical distribution of cyclopoid and poecilostomatoid copepods in the Red Sea. III. Re-evaluation for separating a new species of Oncaea . Marine Ecology Progress Series 80, 301304.Google Scholar
Böttger-Schnack, R. (1994) The microcopepod fauna in the Eastern Mediterranean and Arabian Seas: a comparison with the Red Sea fauna. Hydrobiologia 292/293, 271282.Google Scholar
Böttger-Schnack, R. (1996) Vertical structure of small metazoan plankton, especially noncalanoid copepods. I. Deep Arabian Sea. Journal of Plankton Research 18, 10731101.Google Scholar
Böttjer, D. and Morales, C.E. (2005) Microzooplankton grazing in a coastal embayment off Concepción, Chile, (~36° S) during non-upwelling conditions. Journal of Plankton Research 27, 383391.Google Scholar
Calbet, A. (2008) The trophic roles of microzooplankton in marine systems. ICES Journal of Marine Science 65, 325331.Google Scholar
Calbet, A., Garrido, S., Saiz, E., Alcaraz, M. and Duarte, C.M. (2001) Annual zooplankton succession in coastal NW Mediterranean waters: the importance of the smaller size fractions. Journal of Plankton Research 23, 319331.Google Scholar
Calbet, A. and Landry, M.R. (1999) Mesozooplankton influences on the microbial food web: direct and indirect trophic interactions in the oligotrophic open-ocean. Limnology and Oceanography 44, 13701380.Google Scholar
Castelin, M., Lorion, J., Brisset, J., Cruaud, C., Maestrati, P., Utge, J. and Samadi, S. (2012) Speciation patterns in gastropods with longlived larvae from deep-sea seamounts. Molecular Ecology 21, 48284853.Google Scholar
Castro, J.J., Santiago, J.A. and Hernández-García, V. (1999) Fish associated with fish aggregation devices off the Canary Islands (Central-East Atlantic). Scientia Marina 63, 191198.Google Scholar
Chivers, A.J., Narayanaswamy, B.E., Lamont, P.A., Dale, A. and Turnewitsch, R. (2013) Changes in polychaete standing stock and diversity on the northern side of Senghor Seamount (NE Atlantic). Biogeosciences 10, 35353546.Google Scholar
Christiansen, B., Martin, B. and Hirch, S. (2009) The benthopelagic fish fauna on the summit of Seine Seamount, NE Atlantic: composition, population structure and diets. Deep-Sea Research II 56, 27052712.Google Scholar
Clark, M.R., Rowden, A.A., Schlacher, T.A., Williams, A., Consalvey, M., Stocks, K.I., Rogers, A.D., O'Hara, T.D., White, M., Shank, T.M. and Hall-Spencer, J.M. (2010) The ecology of seamounts: structure, function, and human impacts. Annual Review of Marine Science 2, 253278.Google Scholar
Clarke, K.R. and Gorley, R.N. (2006) PRIMER v6: user manual/tutorial. Plymouth: PRIMER-E Ltd, 192 pp.Google Scholar
Clarke, K.R., Sommerfield, P.J. and Gorley, R.N. (2008) Testing of null hypotheses in exploratory community analyses: similarity profiles and biota-environment linkage. Journal of Experimental Marine Biology and Ecology 366, 5669.Google Scholar
Clarke, K.R. and Warwick, R.M. (2001) Change in marine communities: an approach to statistical analysis and interpretation. 2nd edition. Plymouth: PRIMER-E Ltd, 176 pp.Google Scholar
Denda, A. (2015) Zooplankton dynamics, fish zonation and trophic interactions at two seamounts in contrasting regimes of the Eastern Atlantic . Dissertation, Universität Hamburg, Hamburg, 213 pp.Google Scholar
Denda, A. and Christiansen, B. (2014) Zooplankton distribution patterns at two seamounts in the subtropical and tropical NE Atlantic. Marine Ecology 35, 159179.Google Scholar
Dower, J.F., Freeland, H. and Juniper, S.K. (1992) A strong biological response to oceanic flow past Cobb Seamount. Deep-Sea Research 39, 11391145.Google Scholar
Dower, J.F. and Mackas, D.L. (1996) ‘Seamount effects’ in the zooplankton community near Cobb Seamount. Deep-Sea Research I 43, 837858.Google Scholar
Dumont, M., Kiriakoulakis, K., Legg, S., Mohn, C., Peine, F., Wolff, G. and Turnewitsch, R. (submitted) Tidally-induced enhancement of carbon export about a tall seamount detected using the thorium-234 method. Deep-Sea Research I, submitted.Google Scholar
Edwards, E.S., Burkill, P.H. and Stelfox, C.E. (1999) Zooplankton herbivory in the Arabian Sea during and after the SW monsoon, 1994. Deep-Sea Research II 46, 843863.Google Scholar
Fabian, H., Koppelmann, R. and Weikert, H. (2005) Full-depths zooplankton composition at two deep sites in the western and central Arabian Sea. Indian Journal of Marine Sciences 34, 174187.Google Scholar
Fernández, E. and Pingree, R.D. (1996) Coupling between physical and biological fields in the North Atlantic subtropical front southeast of the Azores. Deep-Sea Research I 43, 13691393.Google Scholar
Firing, E. and Hummon, J.M. (2010) Shipboard ADCP measurements. The go-ship repeat hydrography manual: a collection of expert reports and guidelines. IOCCP Report 14, 111.Google Scholar
Firing, E., Ranada, J. and Caldwell, P. (1995) Processing ADCP data with the CODAS software system version 3.1. Honolulu: Joint Institute for Marine and Atmospheric Research/NODC, University of Hawaii at Manoa.Google Scholar
Gallienne, C.P. and Robins, D.B. (2001) Is Oithona the most important copepod in the world's ocean? Journal of Plankton Research 23, 14211432.CrossRefGoogle Scholar
Genin, A. (2004) Bio-physical coupling in the formation of zooplankton and fish aggregations over abrupt topographies. Journal of Marine Systems 50, 320.CrossRefGoogle Scholar
Genin, A. and Boehlert, G.W. (1985) Dynamics of temperature and chlorophyll structures above a seamount: an oceanic experiment. Journal of Marine Research 43, 907924.Google Scholar
Genin, A. and Dower, J.F. (2007) Seamount plankton dynamics. In Pitcher, T.J., Morato, T., Hart, P.J.B., Clark, M.R., Haggan, N. and Santos, R.S. (eds) Seamounts: ecology, conservation and management. Oxford: Blackwell, pp. 85100.Google Scholar
Go, Y.B., Oh, B.C. and Terazaki, M. (1998) Feeding behavior of the poecilostomatoid copepods Oncaea spp. on chaetognaths. Journal of Marine Systems 15, 475482.Google Scholar
Greene, C.H. (1990) A brief review and critique of zooplankton sampling methods: copepodology for the larval ecologist. Ophelia 32, 109113.CrossRefGoogle Scholar
Hanel, R., John, H.-C., Meyer-Klaeden, O. and Piatkowski, U. (2010) Larval fish abundance, composition and distribution at Senghor Seamount (Cape Verde Islands). Journal of Plankton Research 32, 15411556.Google Scholar
Hatzky, J. (2005) Ampère seamount. In Wille, P.C. (ed.) Sound Images of the ocean in research and monitoring. Berlin: Springer, p. 471.Google Scholar
Hays, G.C. (1996) Large-scale patterns of diel vertical migration in the North Atlantic. Deep-Sea Research I 43, 16011615.Google Scholar
Hernández-León, S. and Ikeda, T. (2005) A global assessment of mesozooplankton respiration in the ocean. Journal of Plankton Research 27, 153158.Google Scholar
Hernández-León, S., Postel, L., Arístegui, J., Gómez, M., Montero, M.F., Torres, S., Almeida, C., Kühner, E., Brenning, U. and Hagen, E. (1999) Large-scale and mesoscale distribution of plankton biomass and metabolic activity in the northeastern central Atlantic. Journal of Oceanography 55, 471482.Google Scholar
Hirch, S., Martin, B. and Christiansen, B. (2009) Zooplankton metabolism and carbon demand at two seamounts in the NE Atlantic. Deep-Sea Research II 56, 26562670.Google Scholar
Huskin, I., Anadón, R., Medina, G., Head, R.N. and Harris, R.P. (2001) Mesozooplankton distribution and copepod grazing in the subtropical Atlantic near the Azores: influence of mesoscale structures. Journal of Plankton Research 23, 671691.Google Scholar
Ikeda, T., Kanno, Y., Ozaki, K. and Shinada, A. (2001) Metabolic rates of epipelagic marine copepods as a function of body mass and temperature. Marine Biology 139, 587596.Google Scholar
Ikeda, T., Sano, F., Yamaguchi, A. and Matsuishi, T. (2006) Metabolism of mesopelagic and bathypelagic copepods in the western North Pacific Ocean. Marine Ecology Progress Series 322, 199211.Google Scholar
Ikeda, T., Torres, J.J., Hernandez-Leon, S. and Geiger, S.P. (2000) Metabolism. In Harris, R.P., Wiebe, P.H., Lenz, J., Skjoldal, H.R. and Huntley, M. (eds) ICES zooplankton methodology manual. San Diego, CA: Academic Press, pp. 455532.Google Scholar
IOC, IHO, BODC (2003) Centenary Edition of the GEBCO Digital Atlas, published on CD ROM on behalf of the Intergovernmental Oceanographic Commission and the International Hydrographic Organization as part of the General Bathymetric Chart of the Oceans. British Oceanographic Data Centre. Available via British Oceanographic Data Centre.Google Scholar
Jackson, G.A. (1980) Phytoplankton growth and zooplankton grazing in oligotrophic oceans. Nature 284, 439441.Google Scholar
Kattner, G., Albers, C., Graeve, M. and Schnack-Schiel, S.B. (2003) Fatty acid and alcohol composition of the small polar copepods, Oithona and Oncaea: indication on feeding modes. Polar Biology 26, 666671.Google Scholar
Kaufmann, M. (2004) Der Einfluss von Seamounts auf die klein- und mesoskalige Verteilung des Phytoplanktons im zentralen, subtropischen Nordostatlantik . Dissertation, Christian-Albrechts-Universität zu Kiel, Kiel, 157 pp.Google Scholar
King, F.D., Devol, A.H. and Packard, T.T. (1978) Plankton metabolic activity in the eastern tropical North Pacific. Deep-Sea Research 25, 689704.Google Scholar
Kleppel, G.S. (1993) On the diets of calanoid copepods. Marine Ecology Progress Series 99, 183195.Google Scholar
Koppelmann, R. and Weikert, H. (2007) Spatial and temporal distribution patterns of deep-sea mesozooplankton in the eastern Mediterranean – indications of a climatically induced shift? Marine Ecology 28, 117.Google Scholar
Koski, M., Kiørboe, T. and Takahashi, K. (2005) Benthic life in the pelagic: aggregate encounter and degradation rates by pelagic harpacticoid copepods. Limnology and Oceanography 50, 12541263.Google Scholar
Kuhn, T., Halbach, P. and Maggiulli, M. (1996) Formation of ferromanganese microcrusts in relation to glacial/interglacial stages in Pleistocene sediments from Ampere Seamount (Subtropical NE Atlantic). Chemical Geology 130, 217232.Google Scholar
Landry, M.R., Constantinou, J. and Kirshtein, J. (1995) Microzooplankton grazing in the central equatorial Pacific during February and August, 1992. Deep-Sea Research II 42, 657671.Google Scholar
Lathuilière, C., Echevin, V. and Lévy, M. (2008) Seasonal and intraseasonal surface chlorophyll-a variability along the northwest African coast. Journal of Geophysical Research 113, C05007.CrossRefGoogle Scholar
Lavelle, J.W. and Mohn, C. (2010) Motion, commotion, and biophysical connections at deep ocean seamounts. Oceanography 23(Sp. Issue 1), 90103.Google Scholar
Lo, W.-T., Shih, C-T. and Hwang, J.-S. (2004) Diel vertical migration of the planktonic copepods at an upwelling station north of Taiwan, western North Pacific. Journal of Plankton Research 26, 8997.Google Scholar
Lozán, J.L. and Kausch, H. (2004) Angewandte Statistik für Naturwissenschaftler, 3. Auflage, Hamburg: Wissenschaftliche Auswertungen, 300 pp.Google Scholar
Martin, B. and Christiansen, B. (2009) Distribution of zooplankton biomass at three seamounts in the NE Atlantic. Deep-Sea Research II 56, 26712682.Google Scholar
Mason, E., Colas, F., Molemaker, J., Shchepetkin, A.F., Troupin, C., McWilliams, J.C. and Sangrà, P. (2011) Seasonal variability of the Canary Current: a numerical study. Journal of Geophysical Research 116, 120.Google Scholar
McEwen, G.F., Johnson, M.W. and Folsom, T.R. (1954) A statistical analysis of the performance of the plankton splitter, based on test observations. Archiv für Meteorologie, Geophysik und Bioklimatologie Serie A 7, 502527.Google Scholar
Metaxas, A. (2011) Spatial patterns of larval abundance at hydrothermal vents on seamounts: evidence for recruitment limitation. Marine Ecology Progress Series 437, 103117.Google Scholar
Mittelstaedt, E. (1991) The ocean boundary along the northwest African coast: circulation and oceanographic properties at the sea surface. Progress in Oceanography 26, 307355.Google Scholar
Mohn, C. and Beckmann, A. (2002) The upper ocean circulation at Great Meteor Seamount. Part I: Structure of density and flow fields. Ocean Dynamics 52, 179193.Google Scholar
Morel, A. (1996) An ocean flux study in eutrophic, mesotrophic and oligotrophic situations: the EUMELI program. Deep-Sea Research I 43, 11851190.Google Scholar
Mouriño, B., Fernandez, E., Serret, P., Harbour, D., Sinha, B. and Pingree, R. (2001) Variability and seasonality of physical and biological fields at the Great Meteor Tablemount (subtropical NE Atlantic). Oceanologica Acta 24, 120.Google Scholar
Müller, B. and Siedler, G. (1992) Multi-year current time series in the eastern North Atlantic Ocean. Journal of Marine Research 50, 6398.Google Scholar
Mullineaux, L.S. and Mills, S.W. (1997) A test of the larval retention hypothesis in seamount-generated flows. Deep-Sea Research I 44, 745770.CrossRefGoogle Scholar
Nellen, W. (1973) Untersuchungen zur Verteilung von Fischlarven und Plankton im Gebiet der Großen Meteorbank. ‘Meteor’ Forschungschungsergebnisse Reihe D 13, 4769.Google Scholar
Ohman, M.D. (1990) The demographic benefits of diel vertical migration by zooplankton. Ecological Monographs 60, 257281.Google Scholar
Ohtsuka, S. and Kubo, N. (1991) Larvaceans and their houses as important food for some pelagic copepods. Proceedings of the Fourth International Conference on Copepoda. Bulletin of Plankton Society Japan Sp. Vol., 535551.Google Scholar
Ohtsuka, S., Kubo, N., Okada, M. and Gushima, K. (1993) Attachment and feeding of pelagic copepods on larvacean houses. Journal of Oceanography 49, 115120.Google Scholar
Omori, M. and Ikeda, T. (1984) Methods in marine zooplankton ecology. New York, NY: John Wiley and Sons, 332 pp.Google Scholar
Onken, R. and Klein, B. (1991) A model of baroclinic instability and waves between the ventilated gyre and the shadow zone of the North Atlantic Ocean. Journal of Physical Oceanography 21, 5366.Google Scholar
Parker, T. and Tunnicliffe, V. (1994) Dispersal strategies of the biota on an oceanic seamount: implications for ecology and biogeography. Biological Bulletin 187, 336345.CrossRefGoogle Scholar
Pastor, M.V., Pelegrí, J.L., Hernández-Guerra, A., Font, J., Salat, J. and Emelianov, M. (2008) Water and nutrient fluxes off northwest Africa. Continental Shelf Research 28, 915936.Google Scholar
Pfaffenhöfer, G.A. (1993) On the ecology of marine cyclopoid copepods (Crustacea, Copepoda). Journal of Plankton Research 15, 3755.Google Scholar
Pierre, C., Vangriesheim, A. and Laube-Lenfant, E. (1994) Variability of water masses and of organic production-regeneration systems as related to eutrophic, mesotrophic and oligotrophic conditions in the northeast Atlantic ocean. Journal of Marine Systems 5, 159170.Google Scholar
Richardson, P.L. (1980) Anticyclonic eddies generated near the Corner Rise Seamount. Journal of Marine Research 38, 673686.Google Scholar
Richardson, P.L. (1981) Gulf Stream trajectories measured with free-drifting buoys. Journal of Physical Oceanography 11, 9991010.Google Scholar
Robinson, C., Serret, P., Tilstone, G., Teira, E., Zubkov, M.V., Rees, A.P., Malcolm, E. and Woodward, S. (2002) Plankton respiration in the Eastern Atlantic Ocean. Deep-Sea Research I 49, 787813.Google Scholar
Roden, G.I. (1987) Effects of seamounts and seamount chains on ocean circulation and thermohaline structure. In Keating, B., Fryer, P., Batzia, R. and Boehlert, G. (eds) Seamounts, islands and atolls. Washington, DC: American Geophysical Union, pp. 335354. [Geophysical Monograph, no 43.]Google Scholar
Roden, G.I. (1994) Effects of the Fieberling seamount group upon flow and thermohaline structure in the spring of 1991. Journal of Geophysical Research 99, 99419961.Google Scholar
Roe, H.S.J. (1988) Midwater biomass profiles over the Madeira Abyssal Plain and the contribution of copepods. Hydrobiologia 167/168, 169181.Google Scholar
Rogers, A.D. (1994) The biology of seamounts. Advances in Marine Biology 30, 305350.Google Scholar
Rowden, A.A., Schlacher, T.A., Williams, A., Clark, M.R., Stewart, R., Althaus, F., Bowden, D.A., Consalvey, M., Robinson, W. and Dowdney, J. (2010) A test of the seamount oasis hypothesis: seamounts support higher epibenthic megafaunal biomass than adjacent slopes. Marine Ecology 31(Suppl. 1), 112.Google Scholar
Saltzmann, J. and Wishner, K.F. (1997a) Zooplankton ecology in the eastern tropical Pacific oxygen minimum zone above a seamount: 1. General trends. Deep-Sea Research I 44, 907930.Google Scholar
Saltzmann, J. and Wishner, K.F. (1997b) Zooplankton ecology in the eastern tropical Pacific oxygen minimum zone above a seamount: 2. Vertical distribution of copepods. Deep-Sea Research I 44, 931954.Google Scholar
Sánchez-Velasco, L. and Shirasago, B. (1999) Spatial distribution of some groups of microzooplankton in relation to oceanographic processes in the vicinity of a submarine canyon in the north-western Mediterranean Sea. ICES Journal of Marine Science 56, 114.Google Scholar
Schmoker, C., Hernández-León, S. and Calbet, A. (2013) Microzooplankton grazing in the oceans: impacts, data variability, knowledge gaps and future directions. Journal of Plankton Research 35, 691706.Google Scholar
Shank, T.M. (2010) Seamounts: deep-ocean laboratories of faunal connectivity, evolution, and endemism. Oceanography 23(Spec. Issue 1), 108122.Google Scholar
Sherr, E.B. and Sherr, B.F. (2007) Heterotrophic dinoflagellates: a significant component of microzooplankton biomass and major grazers of diatoms in the sea. Marine Ecology Progress Series 352, 187197.Google Scholar
Siedler, G. and Onken, R. (1996) Eastern recirculation. In Krauss, W. (ed.) The warm water sphere of the North Atlantic Ocean. Berlin: Gebrüder Bornträger, pp. 339364.Google Scholar
Smith, W.H.F. and Sandwell, D.T. (1997) Global seafloor topography from satellite altimetry and ship depth soundings. Science 277, 19571962.Google Scholar
Sokal, R.R. and Rohlf, F.J. (2009) Introduction to biostatistics. Dover edition. New York, NY: W.H. Freeman, 361 pp.Google Scholar
Southward, A.J. (1998) New observations on barnacles (Crustacea: Cirripedia) of the Azores region. Arquipélago 16A, 1127.Google Scholar
SPSS Inc. (1999) Systat 8.0: statistics. Upper Saddle River, NJ: Prentice Hall.Google Scholar
Ssalto/Duacs User Handbook (2006) (M)SLA and (M)ADT Near-Real Time and Delayed Time Products, SALP-MU-P-EA-21065-CLS, Edition 4.4. http://www.aviso.oceanobs.com/fileadmin/documents/data/tools/hdbk_duacs.pdf Google Scholar
Stadler, L.C. and Marcus, N.H. (1997) Zooplankton responses to hypoxia: behavioral patterns and survival of three species of calanoid copepods. Marine Biology 127, 599607.Google Scholar
Steedmann, H.F. (1976) Examination, sorting and observation fluid. In Steedmann, H.F. (ed.) Zooplankton fixation and preservation. Paris: UNESCO Press, pp. 182183.Google Scholar
Stocks, K.I. and Hart, P.J.B. (2007) Biogeography and biodiversity of seamounts. In Pitcher, T.J., Morato, T., Hart, P.J.B., Clark, M.R., Haggan, N. and Santos, R.S. (eds) Seamounts: ecology, fisheries and conservation. Oxford: Blackwell Publishing, pp. 255281.Google Scholar
Suzuki, K., Nakamura, Y. and Hiromi, J. (1999) Feeding by the small calanoid copepod Paracalanus sp. on heterotrophic dinoflagellates and ciliates. Aquatic Microbial Ecology 17, 99103.Google Scholar
Tomczak, M. Jr (1981) An analysis of mixing in the frontal zone of South and North Atlantic Central Water off North-West Africa. Progress in Oceanography 10, 173192.Google Scholar
Tranter, D.J. (1962) Zooplankton abundance in Australasian waters. Australian Journal of Marine and Freshwater Research 13, 106142.Google Scholar
Troupin, C., Machín, F., Ouberdous, M., Sirjacobs, D., Barth, A. and Beckers, J.-M. (2010) High-resolution climatology of the Northeast Atlantic using Data – Interpolating Variational Analysis (DIVA). Journal of Geophysical Research 115, C08005. doi: 10.1029/2009JC005512.Google Scholar
Turner, J.T. (2004) The importance of small planktonic copepods and their roles in pelagic marine food webs. Zoological Studies 43, 255266.Google Scholar
Uye, S., Aoto, I. and Onbé, T. (2002) Seasonal population dynamics and production of Microsetella norvegica, a widely distributed but little-studied marine planktonic harpacticoid copepod. Journal of Plankton Research 24, 143153.Google Scholar
Vangriesheim, A., Bournot-Marec, C. and Fontan, A. (2003) Flow variability near the Cape Verde frontal zone (subtropical Atlantic Ocean). Oceanologica Acta 26, 149159.Google Scholar
Weikert, H. (1977) Copepod carcasses in the upwelling region South of Cap Blanc, N.W. Africa. Marine Biology 42, 351355.Google Scholar
Weikert, H. and John, H.C. (1981) Experiences with a modified Bé multiple opening-closing plankton net. Journal of Plankton Research 3, 167176.Google Scholar
Wheeler, E.H. (1967) Copepod detritus in the deep sea. Limnology and Oceanography 12, 697702.Google Scholar
White, M., Bashmachnikov, I., Arístegui, J. and Martins, A. (2007) Physical processes and seamount productivity. In Pitcher, T.J., Morato, T., Hart, P.J.B., Clark, M.R., Haggan, N. and Santos, R.S. (eds) Seamounts: ecology, conservation and management. Oxford: Blackwell, pp. 6584.Google Scholar
Wiborg, K.F. (1951) The whirling vessel, an apparatus for the fractioning of plankton samples. Fiskeridirektoratets Skrifter Serie Havundersøkelser 9, 116.Google Scholar
Wiebe, P.H., Copley, N.J. and Boyd, S.H. (1992) Coarse-scale horizontal patchiness and vertical migration of zooplankton in Gulf Stream warm-core ring 82-H. Deep-Sea Research 39(Suppl. 1), S247S278.Google Scholar
Zenk, W., Klein, B. and Schröder, M. (1991) Cape Verde frontal zone. Deep-Sea Research 38(Suppl. 1), S505S530.Google Scholar
Supplementary material: Image

Denda supplementary material

Figure S1

Download Denda supplementary material(Image)
Image 532.4 KB
Supplementary material: File

Denda supplementary material

Figure S1 Legend and Tables S1-S3

Download Denda supplementary material(File)
File 239.1 KB