Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-25T07:18:31.391Z Has data issue: false hasContentIssue false

Chapter 6 - Primary Production, Cycling of Nutrients, Surface Layer and Plankton

from Assessment of Major Ecosystem Services from the Marine Environment (Other than Provisioning Services)

Published online by Cambridge University Press:  18 May 2017

United Nations
Affiliation:
Division for Ocean Affairs and the Law of the Sea, Office of Legal Affairs
Get access

Summary

Primary Production

Definition and ecological significance

Gross primary production (GPP) is the rate at which photosynthetic plants and bacteria use sunlight to convert carbon dioxide (CO2) and water to the high-energy organic carbon compounds used to fuel growth. Free oxygen (O2) is released during the process. Net primary production (NPP) is GPP less the respiratory release of CO2 by photosynthetic organisms, i.e., the net photosynthetic fixation of inorganic carbon into autotrophic biomass. NPP supports most life on Earth; it fuels global cycles of carbon, nitrogen, phosphorus and other nutrients and is an important parameter of atmospheric CO2 and O2 levels (and, therefore, of anthropogenic climate change).

Global NPP is estimated to be ∼105 Pg C yr-1, about half of which is by marine plants (Field et al., 1998; Falkowski and Raven, 1997; Westberry et al., 2008). Within the euphotic zone of the upper ocean, phytoplankton and macrophytes respectively account for ∼94 per cent (∼50 ﹜ 28 Pg C yr-1) and ∼6 per cent (∼3.0 Pg C yr-1) of NPP (Falkowski et al., 2004; Duarte et al., 2005; Carr et al., 2006; Schneider et al., 2008; Chavez et al., 2011; Ma et al., 2014; Rousseaux and Gregg, 2014). All NPP is not equal in terms of its fate. Marine macrophytes play an important role as carbon sinks in the global carbon cycle, provide habitat for a diversity of animal species, and food for marine and terrestrial consumers (Smith, 1981; Twilley et al., 1992; Duarte et al., 2005; Duarte et al., 2010; Heck et al., 2008; Nellemann et al., 2009; McLeod et al., 2011, Fourqurean et al., 2012). Phytoplankton NPP fuels the marine food webs upon which marine fisheries depend (Pauly and Christensen, 1995; Chassot et al., 2010) and the “biological pump” which transports 2-12 Pg C yr-1 of organic carbon to the deep sea (Falkowski et al., 1998; Muller-Karger et al., 2005; Emerson and Hedges, 2008; Doney, 2010; Passow and Carlson, 2012), where it is sequestered from the atmospheric pool of carbon for 200-1500 years (Craig, 1957; Schlitzer et al., 2003; Primeau and Holzer, 2006; Buesseler, et al., 2007).

Changes in the size structure of phytoplankton communities influence the fate of NPP (Malone, 1980; Legendre and Rassoulzadegan, 1996; Pomeroy et al., 2007; Maranon, 2009).

Type
Chapter
Information
The First Global Integrated Marine Assessment
World Ocean Assessment I
, pp. 119 - 148
Publisher: Cambridge University Press
Print publication year: 2017

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

Adornato, L., Cardenas-Valencia, A., Kaltenbacher, E., Byrne, R.H., Daly, K., and others (2010). In situ nutrient sensors for ocean observing systems. In Proceedings of OceanObs–09: Sustained Ocean Observations and Information for Society(Vol. 2), Venice, Italy, 21-25 September 2009. Hall, J., Harrison, D.E., & Stammer, D., eds. ESA Publication WPP-306, doi:10.5270/OceanObs09.cwp.01.
ACT (Alliance for Coastal Technologies) (2003). State of Technology in the Development and Application of Nutrient Sensors; ACT (2007). Recent Developments in In Situ Nutrient Sensors: Applications and Future Directions, ACT 06-08 (UMCES CBL 07-048).
Agawin, N.S.R., Duarte, C.M., and Agustí, S. (2000). Nutrient and temperature control of the contribution of picoplankton to phytoplankton biomass and production. Limnology and Oceanography, 45(3): 591-600.Google Scholar
Alongi, D. (2008). Mangrove forests: Resilienc., protection from tsunamis, and responses to global climate change. Estuarine Coastal and Shelf Science, 76, 1-13.Google Scholar
Altabet, M.A. (2007). Constraints on oceanic N balance/imbalance from sedimentary 15N record., Biogeosciences, 4: 75–86.Google Scholar
Amarasinghe, M.D., and Balasubramaniam, S. (1992). Net primary productivity of two mangrove forests stands on the northwestern coast of Sri Lanka. Hydrobiologia, 247: 37–47.Google Scholar
Anisimov, O.A., Vaughan, D.G., Callaghan, T.V., Furgal, C., Marchant, H., Prowse, T.D., Vilhjálmsson, H., and Walsh, J.E. (2007). Polar regions (Arctic and Antarctic). Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Parry, M.L., Canziani, O.F., Palutikof, J.P., van der Linden, P.J., and Hanson, C.E., eds. Cambridge University Press, Cambridge, 653-685.
Antoine, D., and Morel, A. (1996). Oceanic primary production. 1. Adaptation of a spectral light-photosynthesis model in view of application to satellite chlorophyll observations. Global Biogeochemical Cycles, 10 (1): 43-55.Google Scholar
Appeltans, W., Ahyong, S.T., Anderson, G., Angel, M.V., Artois, T., et al. (2012). The magnitude of global marine species diversity. Current Biology, 22: 2189 – 2202.Google Scholar
Arrigo, K.R. (2005). Marine microorganisms and global nutrient cycles. Nature, 437, 349-355.Google Scholar
Arrigo, K.R. and van Dijke., G.L. (2011). Secular trends in Arctic Ocean net primary production, Journal of Geophysical Research, 116, C09011, doi:10.1029/2011JC007151.Google Scholar
Arrigo, K.R., van Dijken, G.L., and Pabi, S. (2008). Impact of a shrinking Arctic ice cover on marine primary production, Geophysical Research Letters, 35(19), L19603.Google Scholar
Atkinson, D., Ciotti, B.J., and Montagnes, D.J.S. (2003). Protists decrease in size linearly with temperature. Proceedings of the Royal Society of London B, 270: 2605-2611.Google Scholar
Azam, F., Fenchel, T., Field, J.G., Gray, J.S., Meyer-Reil, L.A., Thingstad, F. (1983). The Ecological Role of Water-Column Microbes in the Sea. Marine Ecology Progress Series, 10: 257-263.Google Scholar
Babin, M., Roesler, C.S., and Cullen, J.J., eds. (2008). Real-Time Coastal Observing Systems for Marine Ecosystem Dynamics and Harmful Algal Blooms: Theory, Instrumentation and Modelling. Monographs on Oceanographic Methodology Series, UNESCO.
Balch, W.M., Drapeau, D.T., Bowler, B.C., Lyczskowski, E., Booth, S., and Alley, D. (2011). The contribution of coccolithophores to the optical and inorganic carbon budgets during the Southern Ocean Gas Exchange Experiment: New evidence in support of the “Great Calcite Belt” hypothesis, Journal of Geophysical Research, 116, C00F06, doi:10.1029/2011JC006941.Google Scholar
Balch, W.M., and Utgoff, P.E. (2009). Potential interactions among ocean acidification, coccolithophores, and the optical properties of seawater. Oceanography, 22 (4): 146-159.Google Scholar
Barber, R.T., Sanderson, M.P., Lindley, S.T., Chai, F., Newton, J., Trees, C.C., Foley, D.G., and Chavez, F.P. (1996). Primary productivity and its regulation in the equatorial Pacific during and following the 1991–92 El Niño. Deep-Sea Research II, 43: 933–969.Google Scholar
Barbier, E.B., Hacker, S.D., Kennedy, C., Koch, E.W., Stier, A.C., and Silliman, B.R. (2011). The value of estuarine and coastal ecosystem services. Ecological Monographs, 81: 169–193Google Scholar
Barcelos Ramos, J., Biswas, H., Schulz, K.G., LaRoche, J., and Riebesel, U. (2007). Effect of rising atmospheric carbon dioxide on the marine nitrogen fixer Trichodesmium. Global Biogeochemical Cycles, 21: GB2028, doi:10.1029/2006GB002898.Google Scholar
Barnard, R., Batten, S.D., Beaugrand, G. et al. (2004). Continuous Plankton Records: Plankton atlas of the North Atlantic Ocean (1958 – 1999). II. Biogeograhical charts. Marine Ecology Progress Series, Supplement: 11-75.
Barnes, C., Maxwell, D., Reuman, D.C., and Jennings, S. (2010). Global patterns in predator–prey size relationships reveal size dependency of trophic transfer efficiency. Ecology, 91:222–232.Google Scholar
Barnes, C., Irigoien, X., De Oliveira, J.A.A., Maxwell, D., and Jennings, S. (2011). Predicting marine phytoplankton community size structure from empirical relationships with remotely sensed variables. Journal of Plankton Research, 33: 13–24.Google Scholar
Barton, A.D., Dutkiewicz, S., Flierl, G., Bragg, J., Follows, M.J. (2010). Patterns of diversity in marine phytoplankton. Science, 327, 1509-1511.Google Scholar
Batten, S.D., and Mackas, D.L. (2009). Shortened duration of the annual Neocalanus plumchrus biomass peak in the Northeast Pacific. Marine Ecology Progress Series, 393: 189–198.Google Scholar
Beardall, J., Stojkovic, S., and Larsen, S. (2009). Living in a high CO2 world: impacts of global climate change on marine phytoplankton. Plant Ecology & Diversity, 2: 191-205.Google Scholar
Beaugrand, G., K.M., Brander, J.A., Lindley, S., Souissi, and P.C., Reid. (2003). Plankton effect on cod recruitment in the North Sea. Nature, 426: 661–664.Google Scholar
Beaugrand, G., Reid, P.C., Ibañez, F., Lindley, J.A., and Edwards, M. (2002). Reorganization of the North Atlantic Marine Copepod Biodiversity and Climate. Science, 296: 1692-1694.Google Scholar
Bednaršek, N., Tarling, G.A., Bakker, D.C.E., Fielding, S., Jones, E.M., Venables, H.J., Ward, P., Kuzirian, A., Lézé, B., Feely, R.A., Murphy, E.J. (2012). Extensive dissolution of live pteropods in the Southern Ocean. Nature, 5: 881-885.Google Scholar
Behrenfeld, M.J. (2011). Biology: Uncertain future for marine algae. Nature Climate Change, 1: 33-34.Google Scholar
Behrenfeld, M.J., O'Malley, R.T., Siegel, D.A., McClain, C.R., et al. (2006). Climatedriven trends in contemporary ocean productivity. Nature 444: 752−755.Google Scholar
Behrenfeld, M.J., and Falkowski, P.G. (1997). Photosynthetic Rates Derived from Satellite-Based Chlorophyll Concentration. Limnology and Oceanography, 42: 1–20.Google Scholar
Benitez-Nelson, C.R. (2000). The biogeochemical cycling of phosphorus in marine systems. Earth-Science Review, 51(1-4): 109-135.Google Scholar
Beusen, A.H.W., Slomp, C.P., and Bouwman, A.F. (2013). Global land–ocean linkage: direct inputs of nitrogen to coastal waters via submarine groundwater discharge. Environmental Research Letters, 8: 034035 (doi:10.1088/1748-9326/8/3/034035).Google Scholar
Bidigare, R.R., Chai, F., Landry, M.R., Lukas, R., Hannides, C.C.S., Christensen, S.J., Karl, D.M., Shi, L., and Chao., Y. (2009). Subtropical ocean ecosystem structure changes forced by North Pacific climate variations. Journal of Plankton Research, 31 (10): 1131-39.Google Scholar
Bijma, J., Pörtner, H-O., Yesson, C., and Rogers, A.D. (2013). Climate change and the oceans – What does the future hold? Marine Pollution Bulletin. 74 (2): 495-505.Google Scholar
Bissinger, J.E., Montagnes, D.J.S., Sharples, J., and Atkinson, D. (2008). Predicting marine phytoplankton maximum growth rates from temperature: Improving on the Eppley curve using quantile regression. Limnology and Oceanography, 53: 487–93.Google Scholar
Bittaker, H.F., and Iverson, R.L. (1976). Thalassia testudinum productivity: A field comparison of measurement methods. Marine Biology, 37: 39-46.Google Scholar
Blanchard, J.L., Jennings, S., Holmes, R., Harle, J., Merino, G., Allen, J.I., Holt, J., Dulvy, N.K., and Barange, M. (2012). Potential consequences of climate change for primary production and fish production in large marine ecosystems. Philosophical transactions of the Royal Society of London, Series B, Biological Sciences, 367: 2979–89.
Bode, A., Hare, J., Li, W.K.W., Morán, X.A.G., and Valdés, L. (2011). Chlorophyll and primary production in the North Atlantic. In: Reid, P.C., and Valdés, L., eds. ICES Status Report on Climate Change in the North Atlantic. International Council for the Exploration of the Sea, Copenhagen, pp. 77-102.
Boesch, D.F., and Turner, R.E. (1984). Dependence of Fishery Species on Salt Marshes: The Role of Food and Refuge. Estuaries and Coasts, 7:460–468.Google Scholar
Boesch, D.F., Anderson, D.M., Horner, R.A., Shumway, S.E., Tester, P.A., and Whitledge, T.E. (1997). Harmful Algal Blooms in Coastal Waters: Options for Prevention, Control and Mitigation. Science for Solutions, NOAA Coastal Ocean Program Decision and Analysis Series, No. 10, NOAA Coastal Ocean Office, Silver Spring, MD, 46 pp. + appendix.
Bouillon, S., Borges, A.V., Castañeda-Moya, E., Diele, K., Dittmar, T., Duke, N.C., Kristensen, E., Lee, S.Y, Marchand, C., Middelburg, J.J., Rivera-Monroy, V.H., Smith III, T.J, and Twilley, R.R. (2008). Mangrove production and carbon sinks: A revision of global budget estimates. Global Biogeochemical Cycles, 22(2), doi:10.1029/2007GB003052.Google Scholar
Boyce, D.G., Dowd, M., Lewis, M.R., Worm, B. (2014). Estimating global chlorophyll changes over the past century. Progress in Oceanography, 122: 163-177.Google Scholar
Boyce, D.G., Lewis, M.R., and Worm, B. (2010). Global phytoplankton decline over the past century. Nature, 466: 591-596.Google Scholar
Boyer, E.W., Howarth, R.W., Galloway, J.N., et al. (2006). Riverine nitrogen export from the continents to the coasts. Global Biogeochemical Cycle, 20: GB1S91; doi:10.1029/2005GB002537.Google Scholar
Braatz, S., Fortuna, S., Broadhead, J., and Leslie, R. (2007). Coastal protection in the aftermath of the Indian Ocean tsunami: What role for forests and trees? Rap publication, 2007/07, Proceedings of the Regional Technical Workshop, Khao Lak, Thailand, 28-31 August, 2006.
Brodeur, R.D., Sugisaki, H., Hunt, G.L. Jr., (2002). Increases in jellyfish biomass in the Bering Sea: implications for the ecosystem. Marine Ecology Progress Series, 233, 89–103.Google Scholar
Brotz, L., Cheung, W.W.L., Kleisner, K., Pakhomov, E., Pauly, D. (2012). Increasing jellyfish populations: trends in Large Marine Ecosystems. Hydrobiologia, 690: 3-20.Google Scholar
Brown, Z.W., and Arrigo, K.R. (2012). Contrasting trends in sea ice and primary production in the Bering Sea and Arctic Ocean ICES Journal of Marine Science, 69: 1180-1193.Google Scholar
Buitenhuis, E.T., Li, W.K.W., Vaulot, D., Lomas, M.W., Landry, M.R., Partensky, F., Karl, D.M., Ulloa, O., Campbell, L., Jacquet, S., Lantoine, F., Chavez, F., Macias, D., Gosselin, M., and McManus, G.B., (2012). Picophytoplankton biomass distribution in the global ocean. Earth System Science Data, 4: 37–46.Google Scholar
Buesseler, K.O., Antia, A.N., Chen, M., Fowler, S.W., Gardner, W.D., Gustafsson, O., Harada, K., Michaels, A.F., van der Loeff, M.R., Sarin, M., Steinberg, D.K., and Trull, T. (2007). An assessment of the use of sediment traps for estimating upper ocean particle fluxes. Journal of Marine Research, 65: 345–416.Google Scholar
Burkill, P.H., and Reid, P.C. (2010). Plankton biodiversity of the North Atlantic: changing patterns revealed by the Continuous Plankton Recorder survey. In: Proceedings of OceanObs–09: Sustained Ocean Observations and Information for Society, (Vol. 1). Hall, J., Harrison, D.E., and Stammer, D., eds. ESA Publication WPP-306, doi:10.5270/OceanObs09.pp.09.
Calbet, A. (2008). The trophic roles of microzooplankton in marine systems. ICES Journal of Marine Science, 65: 325–331.Google Scholar
Caldeira, K., and Wickett, M.E. (2003). Oceanography: Anthropogenic carbon and ocean pH. Nature, 425: 365-365.Google Scholar
Caldeira, K., and Wickett, M.E. (2005). Ocean model predictions of chemistry changes from carbon dioxide emissions to the atmosphere and ocean. Journal of Geophysical Research, 110:C09S04, doi:10.1029/2004JC002671.Google Scholar
Capone, D.G. (2008). The Marine Nitrogen Cycle. Microbe, 3 (4): 186 – 192.Google Scholar
Capone, D.G., and Knapp, A.N. (2007). Oceanography: A marine nitrogen cycle fix? Nature, 445: 159-160.Google Scholar
Capone, D.G., Bronk, D.A., Mulholland, M.R., and Carpenter, E.J. (2008). Nitrogen in the Marine Environment. Elsevier Inc., Amsterdam, 1729 pp. pp.
Capotondi, A., Alexander, M.A., Bond, N.A., Curchitser, E.N., and Scott, J.D. (2012). Enhanced upper ocean stratification with climate change in the CMIP3 models. Journal of Geophysical Research, 117, C04031, doi:10.1029/2011JC007409.Google Scholar
Carr, M-E., Friedrichs, M.A.M., et al. (2006). A comparison of global estimates of marine primary production from ocean color. Deep-Sea Research Part II, 53: 741-770.Google Scholar
Cebrian, J. (2002). Variability and control of carbon consumption, export, and accumulation in marine communities. Limnology and Oceanography, Inc., 47(1): 11–22.Google Scholar
CENR. (2003). An Assessment of Coastal Hypoxia and Eutrophication in U.S. Waters. Washington DC: National Science and Technology Council, Committee on Environment and Natural Resources.
Chan, F., Barth, J.A., Lubchenco, J., et al. (2008). Emergence of Anoxia in the California Current Large Marine Ecosystem. Science, 319: 920.Google Scholar
Charpy-Roubaud, C., and Sournia, A. (1990). The comparative estimation of phytoplanktonic, microphytobenthic and macrophytobenthic primary production in the oceans. Marine Microbial Food Webs, 4 (1): 31-57.Google Scholar
Chassot, E., Melin, F., Le Pape, O., and Gascuel, D. (2007). Bottom-up control regulates fisheries production at the scale of eco-regions in European seas. Marine Ecology Progress Series, 343: 45–55.Google Scholar
Chassot, E., Bonhommeau, S., Dulvy, N.K., Mélin, F., Watson, R., Gascuel, D., and Le Pape., O. (2010). Global marine primary production constrains fisheries catches. Ecology Letters, 13(4): 495 – 505.Google Scholar
Chavez, F.P., Messié, M., and Pennington, J.T. (2011). Marine Primary Production in Relation to Climate Variability and Change. Annual Reviews Marine Science, 3: 227-260.Google Scholar
Chaalali, A., Beaugrand, G., Raybaud, V., Goberville, E., David, V., et al. (2013). Climatic Facilitation of the Colonization of an Estuary by Acartia tonsa. PLoS ONE 8(9): e74531. doi:10.1371/journal.pone.0074531Google Scholar
Cheung, W., Lam, V., Sarmieno, J.L., Kearney, K., Watson, R., Zeller, D., Pauly, D. (2010). Large scale redistribution of maximum fisheries catch potential in the global ocean unser climate change. Global Change Biology., 16: 24-35. DOI: 10.1111/j.1365-2486.2009.01995.xGoogle Scholar
Chisholm, S.W. (1992). Phytoplankton siz., p. 213–237. In: Falkowski, P.G., and Woodhead, A.D., (eds.), Primary Productivity and Biogeochemical Cycles in the Sea. Plenum Press.
Church, J.A., and White, N.J. (2006). A 20th century acceleration in global sea-level rise. Geophysical Research Letters, 33: L01602.Google Scholar
Chust, G., Allen, J.I., Bopp, L., Schrum, C., Holt, J., Tsiaras, K., Zavatarelli, M., Chifflet, M., Cannaby, H., Dadou, I., Daewel, U., Wakelin, S.L., Machu, E., Pushpadas, D., Butenschon, M., Artioli, Y., Petihakis, G., Smith, C., Garçon, V., Goubanova, K., Le Vu, B., Fach, B.A., Salihoglu, B., Clementi, E. and Irigoien, X. (2014). Biomass changes and trophic amplification of plankton in a warmer ocean. Global Change Biology, 20: 2124–2139. doi: 10.1111/gcb.12562.Google Scholar
Chynoweth, D.P., Owens, J.M., and Legrand, R. (2001). Renewable methane from anaerobic digestion of biomass. Renewable Energy, 22:1-8.Google Scholar
Cleland, E.E. (2011). Biodiversity and Ecosystem Stability. Nature Education Knowledge, 3(10):14Google Scholar
Cloern, J.E. (1987). Turbidity as a control on phytoplankton biomass and productivity in estuaries. Continental Shelf Research, 7: 1367-1381.Google Scholar
Cloern, J.E. (2001). Review. Our evolving conceptual model of the coastal eutrophication problem. Marine Ecology Progress Series, 210: 223–253.Google Scholar
Cloern, J.E., and Jassby, A.D. (2010). Patterns and Scales of Phytoplankton Variability in Estuarine–Coastal Ecosystems. Estuaries and Coasts, 33:230–241.Google Scholar
Cloern, J.E., Foster, S.Q., and Kleckner, A.E. (2013). Review: phytoplankton primary production in the world's estuarine-coastal ecosystems. Biogeosciences Discussion, 10: 17725–17783.Google Scholar
Codispoti, L.A., Brandes, J.A., Christensen, J.P., Devol, A.H., Naqvi, S.W.A., Paerl, H.W., and Yoshinari, T. (2001). The oceanic fixed nitrogen and nitrous oxide budgets:Moving targets as we enter the anthropocene? Scientia Marina, 65(Suppl. 2): 85–105.Google Scholar
Compton, J.S., Mallinson, D., Glenn, C.R., Filippelli, G.M., Föllmi, K.B., Shields-Zhou, G.A., and Zanin, Y. (2000). Variations in the global phosphorus cycle, in Marine Authigenesis: From Global to Microbial, Glenn, C.R., Prévol-Lucas, L., and Lucas, J., eds. SEPM Special Publication, 66, 21-33.
Condon, R.H., Graham, W.M., Duarte, C.M., Pitt, K.A., Lucas, C.H., Haddock, S.H.D., Sutherland, K.R., Robinson, K.L., Dawson, M.N., Decker, M.B., and others. (2012). Questioning the Rise of Gelatinous Zooplankton in the World's Oceans. BioScience, 62: 160-169.Google Scholar
Conkright, M.E., and Gregg, W.W. (2003). Comparison of global chlorophyll climatologies: In situ, CZCS, blended in situ-CZCS and SeaWiFS. International Journal of Remote Sensing, 24 (5): 969–991.Google Scholar
Conservation International. (2008). Economic Values of Coral Reefs, Mangroves, and Seagrasses: A Global Compilation. Center for Applied Biodiversity Science, Arlington, VA.
Cooley, S.R., Kite-Powell, H.L., and Doney, S.C. (2009). Ocean acidification's potential to alter global marine ecosystem services. Oceanography, 22: 172-181.Google Scholar
Corten, A., and Lindley, J.A. (2003). The use of CPR data in fisheries research. Progress in Oceanography, 58: 285-300.Google Scholar
Costanza, R., d'Arge, R., de Groot, R., Farber, S., Grasso, M., Hannon, B., Limburg, K., Naeem, S., O'Neill, R.V., Paruelo, J., Raskin, R.G., Suttonk, P., and van den Belt, M. (1997). The value of the world's ecosystem services and natural capital. Nature, 387: 253-260.Google Scholar
Cousens, R. (1984). Estimation of Annual Production by the Intertidal Brown Alga Ascophyllum nodosum (L.), Le Jolis. Botanica Marina, 27: 217-227.Google Scholar
Craig, H. (1957). The Natural Distribution of Radiocarbon and the Exchange Time of Carbon Dioxide between Atmosphere and Sea. Tellus, 9(1): 1-17.Google Scholar
Crooks, S., Herr, D., Tamelander, J., Laffoley, D., and Vandever, J. (2011). Mitigating Climate Change through Restoration and Management of Coastal Wetlands and Near-shore Marine Ecosystems: Challenges and Opportunities. Environment Department Paper 121, World Bank, Washington DC.
Cullen, J.J. (2008a). Primary production methods. pp. 578-58., In Steele, J.H., Turekian, K.K., and Thorpe, S.A., eds. Encyclopedia of Ocean Science, Elsevier, ISBN: 978-0-12-374473-9.
Cullen, J.J. (2008b). Observation and prediction of harmful algal bloom., In Babin, M., Roesler, C.S., and Cullen, J.J., eds. Real-Time Coastal Observing Systems for Marine Ecosystem Dynamics and Harmful Algal Blooms: Theory, Instrumentation and Modelling. Monographs on Oceanographic Methodology Series, UNESCO, pp. 1-41.
Cullen, J.J., Davis, R.F., and Huot, Y. (2012). Spectral model of depth-integrated water column photosynthesis and its inhibition by ultraviolet radiation. Global Biolgeochemical Cycles, 26, GB1011, doi:10.1029/2010GB003914.Google Scholar
Curcó, A., Ibàñez, C., Day, J.W., and Prat, N. (2002). Net primary production and decomposition of salt marshes of the Ebre delta (Catalonia, Spain). Estuaries and Coasts, 25(3): 309-324.Google Scholar
Cushing, D.H. (1990). Plankton production and year-class strength in fish populations: an update of the match/mismatch hypothesis. Advances in Marine Biology, 26: 249-293.Google Scholar
Dame, R.F., and Kenny, D. (1986). Variability of Spartina alterniflora primary production in the euhaline North Inlet estuary. Marine Ecology -Progress Series, 32: 71-80.Google Scholar
Danielsen, F., Sørensen, M.K., Olwig, M.F., Selvam, V., Parish, F., Burgess, N.D., Hiraishi, T., Karunagaran, V.M., Rasmussen, M.S., Hansen, L.B., Quarto, A., and Suryadiputra, N. (2005). The Asian tsunami: A protective role for coastal vegetation. Science, 310: 643.Google Scholar
Daufresne, M., Lengfellner, K., and Sommer, U. (2009). Global warming benefits the small in aquatic ecosystems. Proceedings of the National Academy of Sciences, USA 106: 12788–12793
Dave, A.C., and Lozier, M.S. (2013). Examining the global record of interannual variability in stratification and marine productivity in the low-latitude and midlatitude ocean. Journal of Geophysical Research. Oceans, 118, 3114–3127, doi:10.1002/jgrc.20224.Google Scholar
Davidson, M., and Malone, T.C., eds. (2006/2007). Stemming the tide of coastal disasters. Marine Technology Society Journal, 40: 1-125.Google Scholar
Day, J.W., Jr., Coronado-Molina, C., Vera-Herrera, F.R., Twilley, R., Rivera-Monroy, V.H., Alvarez-Guillen, H., Day, R., and Conner, W. (1996). A 7 year record of aboveground net primary production in a southeastern Mexican mangrove forest. Aquatic Botany, 55: 39–60.Google Scholar
Dayton, P., Curran, S., Kitchingman, A., Wilson, M., Catenazzi, A., Restrepo, J., Birkeland, C., Blaber, S., Saifullah, S., Branch, G., Boersma, D., Nixon, S., Dugan, P., Davidson, N., and Vörösmarty, C. (2005). Coastal ecosystems In: Ecosystems and Human Well-being: Current State and Trends: Findings of the Condition and Trends Working Group, Hassan, R., Scholes, R., and Ash, N., eds. Island Press, p. 513-550.
Delille, B., Harlay, J., Zondervan, I., Jacquet, S., Chou, L., et al. (2005). Response of primary production and calcification to changes of pCO2 during experimental blooms of the coccolithophorid Emiliania huxleyi. Global Biogeochemical Cycles, 19, GB2023, doi:10.1029/2004GB002318.Google Scholar
Dennison, W.C., Orth, R.J., Moore, K.A., Stevenson, J.C., Carter, V., Kollar, S., Berg-strom, P.W., and Batiuk, R.A. (1993). Assessing Water Quality with Submersed Aquatic Vegetation. Habitat requirements as barometers of Chesapeake Bay health. BioScience, 43: 86–94.Google Scholar
Depledge, M.H., Galgani, F., Panti, C., Caliani, I., Casini, S., and Fossi, M.C. (2013). Plastic litter in the sea. Marine Environmental Research, 92: 279-281.Google Scholar
Deutsch, C., Sarmiento, J.L., Sigman, D.M., Gruber, N., and Dunne, J.P. (2007). Spatial coupling of nitrogen inputs and losses in the ocean. Nature, 445: 163-167.Google Scholar
DeVries, T., Deutsch, C., Primeau, F., Chang, B., and Devol, A. (2012). Global rates of water-column denitrification derived from nitrogen gas measurements. Nature Geoscience, 5: 547–550.Google Scholar
DeVries, T., Deutsch, C., Rafter, P.A., and Primeau, F. (2013). Marine denitrification rates determined from a global 3-dimensional inverse model. Biogeosciences, 10: 2481–2496.Google Scholar
deYoung, B., Heath, M., Werner, F., Chai, F., Megrey, B., and Monfray, P. (2004). Challenges of Modeling Ocean Basin Ecosystems. Science, 304: 1463-1466.Google Scholar
Diaz, R.J., and Rosenberg, R. (2008). Spreading Dead Zones and Consequences for Marine Ecosystems. Science. 321: 926-929.Google Scholar
Donato, D.C., Kauffman, J.B., Murdiyarso, D., Kurnianto, S., Stidham, M., and Kanninen, M. (2011). Mangroves among the most carbon-rich forests in the tropics. Nature Geoscience, 4: 293-297.Google Scholar
Doney, S.C. (2010). The Growing Human Footprint on Coastal and Open-Ocean Biogeochemistry. Science, 328: 1512-1516.Google Scholar
Doney, S.C., Ruckelshaus, M., Duffy, J.E., Barry, J.P., Chan, F., English, C.A., Galindo, H.M., Grebmeier, J.M., Hollowed, A.B., Knowlton, N., Polovina, J., Rabalais, N.N., Sydeman, W.J., and Talley, L.D. (2012). Climate change impacts on marine ecosystems. Annual Reviews -Marine Science, 4: 11-37.Google Scholar
Doney, S.C., Balch, W.M., Fabry, V.J., and Feely, R.A. (2009). Ocean Acidification: A Critical Emerging Oroblem for the Ocean Sciences. Oceanography, 22 (4): 16-25.Google Scholar
Duarte, C.M. (1989). Temporal biomass variability and production/biomass relationships of seagrass communities. Marine Ecology Progress Series, 51: 269-276.Google Scholar
Duarte, C.M. (1995). Submerged aquatic vegetation in relation to different nutrient regimes. Ophelia, 41: 87-112.Google Scholar
Duarte, C.M., and Kirkman, H. (2001). Methods for the measurement of seagrass abundance and depth distribution. In Global Seagrass Research Methods, Short, F.T., and Coles, R.G., eds. 141–153. Elsevier Science B.V.
Duarte, C.M., Borum, J., Short, F.T., and Walker, D.I. (2008). Seagrass ecosystems: their global status and prospects, p. 281-306. In Aquatic Ecosystems, Polunin, N., ed. Cambridge University Press, Cambridge U.K.
Duarte, C.M., Middelburg, J.J., and Caraco, N.F. (2005). Major role of marine vegetation on the oceanic carbon cycle. Biogeosciences, 2: 1-8.Google Scholar
Duarte, C.M., Marbà, N., Gacia, E., Fourqurean, J.W., Beggins, J., Barrón, C., and Apostolaki, E.T. (2010). Seagrass community metabolism: Assessing the carbon sink capacity of seagrass meadows. Global Biogeochemical Cycles, 24 (4), GB4032, doi:10.1029/2010GB003793Google Scholar
Duce, R.A., LaRoche, J., Altieri, K., Arrigo, K.R., Baker, A.R., et al. (2008). Impacts of Atmospheric Anthropogenic Nitrogen on the Open Ocean. Science, 320: 893-897.Google Scholar
Ducklow, H.W., Fraser, W.R., Meredith, M.P., et al. (2013). West Antarctic Peninsula: An Ice-Dependent Coastal Marine Ecosystem in Transition. Oceanography, 26 (3): 190 – 203.Google Scholar
Duke, N.C., Meynecke, J-O., Dittmann, S., Ellison, A.M., Anger, K., et al. (2007). A world without mangroves? Science, 317: 41-42.Google Scholar
Edwards, M., Reid, P.C., and Planque, B. (2001). Long-term and regional variability of phytoplankton biomass in the north-east Atlantic (1960-1995). ICES Journal of Marine Science 58: 39-49.Google Scholar
Edwards, M., and Richardson, A.J. (2004). Impact of climate change on marine pelagic phenology and trophic mismatch. Nature, 430: 881-884.Google Scholar
Edwards, M., Johns, D.G., Leterme, S.C., Svendsen, E., and Richardson, A.J. (2006). Regional climate change and harmful algal blooms in the northeast Atlantic. Limnology and Oceanography, Inc., 51 (2): 820-829.Google Scholar
Edwards, M., Beaugrand, G., Johns, D.G., Licandro, P., McQuatters-Gollop, A., and Wootton, M. (2010). Ecological status report: results from the CPR survey 2009. SAHFOS Technical Report, 7: 1-8. Plymouth, U.K. ISSN 1744-0750.
Emerson, S., and Hedges, J.I. (2008). Chemical Oceanography and the Marine Carbon Cycle, Cambridge University Press, Cambridge, U.K.
Engel, I., Luo, B.P., Pitts, M.C., Poole, L.R., Hoyle, C.R., Grooß, J.-U., Dörnbrack, A., and Peter, T. (2013). Heterogeneous formation of polar stratospheric clouds – Part 2: Nucleation of ice on synoptic scales. Atmospheric Chemistry and Physics Discussions, 13: 8831-8872.Google Scholar
Falkowski, P.G., (1997) Evolution of the nitrogen cycle and its influence on the biological sequestration of CO2 in the ocean, Nature, 387: 272–275.Google Scholar
Falkowksi, 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 (5682): 354-360.Google Scholar
Falkowski, P.G., and Raven, J.A. (1997). Aquatic photosynthesis. Blackwell Science, Malden, MA.
Falkowski, P.G. Barbe., R.T., and Smetacek, V. (1998). Biogeochemical Controls and Feedbacks on Ocean Primary Production. Science, 281: 200–206.Google Scholar
Feely, R.A., Doney, S.C., and Cooley, S.R. (2009). Ocean Acidification: Present Conditions and Future Changes in a High-CO2 World. Oceanography, 22 (4): 36-47.Google Scholar
Ferreyra, G.A., Mostajir, B., Schloss, I.R., Chatila, K., Ferrario, M.E., Sargian, P., Roy, S., Prod'homme, J., and Demers., S. (2006). Ultraviolet-B radiation effects on the structure and function of lower trophic levels of the marine planktonic food web. Photochem Phtobiol, 82, 887-897.Google Scholar
Field, C.B., Behrenfeld, M.J., Randerson, J.T., and Falkowski, P. (1998). Primary Production of the Biosphere: Integrating Terrestrial and Oceanic Components. Science, 281, 237–240.Google Scholar
Filippelli, G.M., and Delaney, M.L. (1996). Phosphorus geochemistry of equatorial Pacific sediments. Geochimica et Cosmochimica Acta, 60:1479-1495.Google Scholar
Finkel, Z.V., Beardall, J., Flynn, K.J., Quigg, A., Rees, T.A.V., and Raven, J.A. (2010). Phytoplankton in a changing world: cell size and elemental stoichiometry. Journal of Plankton Research, 32: 119–137.Google Scholar
Flores, H., Atkinson, A., Kawaguchi, S., Krafft, B.A., Milinevsky, G., Nicol, S., et al. (2012). Impact of climate change on Antarctic krill. Marine Ecology Progress Series, 458: 1–19.Google Scholar
Forget, M-H., Stuart, V., and Platt, T., eds. (2009). Remote Sensing in Fisheries and Aquaculture. IOCCG Report. No. 8, 120 pp.Google Scholar
Forget, M-H., Platt, T., Sathyendranath, S., Stuart, V., and Delaney, L. (2010). “Societal Applications in Fisheries and Aquaculture Using Remotely-Sensed Imagery -The SAFARI Project” in Proceedings of OceanObs–09: Sustained Ocean Observations and Information for Society (Vol. 2), Venice, Italy, 21-25 September 2009, Hall, J., Harrison, D.E., and Stammer, D., eds., ESA Publication WPP-306. doi:10.5270/ OceanObs09.cwp.30.
Fourqurean, J.W., Duarte, C.M., Kennedy, H., Marbà, N., Holmer, M., Mateo, M.A., Apostolaki, E.T., Kendrick, G.A., Krause-Jensen, D., McGlathery, K.J., and Serrano, O. (2012). Seagrass ecosystems as a globally significant carbon stock. Nature Geoscience, 5: 505–509.Google Scholar
Fraile, I., Schulz, M., Mulitza, S., and Kucera, M. (2008). Predicting the global distribution of planktonic foraminifera using a dynamic ecosystem model. Biogeosciences, 5: 891–911.Google Scholar
Frank, K.T., Petrie, B., Shackell, N.L., and Choi, J.S. (2006). Reconciling differences in trophic control in mid-latitude marine ecosystems. Ecology Letters, 9: 1096–1105.Google Scholar
Franks, P.J.S. (2008). Physics and physical modeling of harmful algal blooms. In Babi., M., Roesler, C.S., and Cullen, J.J., eds. Real-Time Coastal Observing Systems for Marine Ecosystem Dynamics and Harmful Algal Blooms: Theory, Instrumentation and Modelling. Monographs on Oceanographic Methodology Series, UNESCO., pp. 561-598.
Freing, A., Wallace, D.W.R., and Bange, H.W. (2012). Global oceanic production of nitrous oxide. Philosophical Transactions of the Royal Society B, 367: 1245–1255.Google Scholar
Fricke, A., Molis, M., Wiencke, C., Valdivia, N., and Chapman, A.S. (2011). Effects of UV radiation on the structure of Arctic macrobenthic communities. Polar Biology, 34, 995-1009.Google Scholar
Friedland, K.D., Stock, C., Drinkwater, K.F., Link, J.S., Leaf, R.T., et al. (2012). Pathways between Primary Production and Fisheries Yields of Large Marine Ecosystems. PLoS ONE, 7(1): e28945, doi: 10.1371/journal.pone.0028945.Google Scholar
Friedrichs, M.A.M, Carr, M-E., Barber, R.T., Scardi, M., Antoine, D., et al. (2009). Assessing the uncertainties of model estimates of primary productivity in the tropical Pacific Ocean. Journal of Marine Systems, 76: 113–133.Google Scholar
Galloway, J.N., Dentener, F.J., Capone, D.G., Boyer, E.W., Howarth, R.W., Seitzinger, S.P., Asner, G.P., Cleveland, C.C., Green, P.A., Holland, E.A., Karl, D.M., Michaels, A.F., Porter, J.H., Townsend, A.R., and Vörsömarty, C.J. (2004). Nitrogen cycles: past, present and future. Biogeochemistry, 70:153-226.Google Scholar
Garrard, S.L., and N.J., Beaumont. (2014). The effect of ocean acidification on carbon storage and sequestration in seagrass beds; a global and UK context. Marine Pollution Bulletin, 86: 138-146.Google Scholar
Geider, R.J., Delucia, E.H., Falkowski, P.G., Finzi, A.C., Grime, J.P., Grace, J., Kana, T.M., La Roche, J., Long, S.P., Osborne, B.A., Platt, T., Prentice, I.C., Raven, J.A., Schlesinger, W.H., Smetacek, V., Stuart, V., Sathyendrananth, S., Thomas, R.B., Vogelmann, T.C., Williams, P., and Woodward, F.I. (2001). Primary productivity of planet earth: biological determinants and physical constraints in terrestrial and aquatic habitats. Global Change Biology, 7: 849-882.Google Scholar
Gilman, E.L., Ellison, J., Duke, N.C., and Field, C. (2008). Threats to mangroves from climate change and adaptation options: a review. Aquatic Botany, 89: 237–250.Google Scholar
Gitelson, A.A. (2004). Wide dynamic rage vegetation index for remote quantification of biophysical characteristics of vegetation. Journal of plant physiology, 161: 165-173.Google Scholar
Glibert, P.M., Anderson, D.M., Gentien, P., Granéli, E., and Sellner, K.G. (2005). The global, complex phenomena of harmful algal blooms. Oceanography, 18(2): 136-147.Google Scholar
Glibert, P.M., and Bouwman, L. (2012). Land-based Nutrient Pollution and the Relationship to Harmful Algal Blooms in Coastal Marine Systems. Inprint, 2: 5-7.(www.loicz.org).Google Scholar
Goes, J.I., Handa, N., Taguchi, S., and Hama, T. (1995). Changes in the patterns of biosynthesis and composition of amino acids in a marine phytoplankton exposed to ultraviolet-B radiation: nitrogen limitation implicated. Photochemistry and Photobiology, 62, 703-710.Google Scholar
Graham, W.M., Gelcich, S., Robinson, K.L., Duarte, C.M., Brotz, L., Purcell, J.E., Madin, L.P., Mianzan, H., Sutherland, K.R., Uye, S.-I., and others. (2014). Linking human well-being and jellyfish: Ecosystem services, impacts and societal responses. Frontiers in the Ecology and Environment, 12: 515–523.Google Scholar
Gregg, W.W., Conkright, M.E., Ginoux, P., O'Reilly, J.E., and Casey, N.W. (2003). Ocean primary production and climate: global decadal changes. Geophysical Research Letters, 30 (15): 1809, doi:10.1029/2003GL016889.Google Scholar
Gregg, W.W., Casey, N.W., O'Reilly, J.E., and Esaias, W.E. (2009). An empirical approach to ocean color data: Reducing bias and the need for post-launch radiometric re-calibration. Remote Sensing of Environment, 113: 1598–1612.Google Scholar
Greene, C.H., and Pershing, A.J. (2004). Climate and the Conservation Biology of North Atlantic Right Whales: The Right Whale at the Wrong Time? Frontiers in Ecology and the Environment, 2: 29-34.Google Scholar
Gross, M.F., Klemas, V., and Hardisky, M.A. (1990). Long-term remote monitoring of salt marsh biomass. Proceedings SPIE 1300, Remote Sensing of the Biosphere, 59, doi:10.1117/12.21390.Google Scholar
Groβkopf, T., Mohr, W., Baustian, T., Schunck, H., Gill, D., Kuypers, M.M.M., Lavik, G., Schmitz, R.A., Wallace, D.W.R., and LaRoche, J. (2012). Doubling of marine dinitrogen-fixation rates based on direct measurements. Nature, 488, 361-364.Google Scholar
Gruber, N. (2004). The Dynamics of the Marine Nitrogen Cycle and its Influence on Atmospheric CO2 Variations. In The Ocean Carbon Cycle and Climate, NATO Science Series, Follows, M., and Oguz., T., eds. Dordrecht: Kluwer Academic, pp. 97–148.
Gruber, N. (2008). The Marine Nitrogen Cycle: Overview and Challenges. In Nitrogen in the marine environment. Capon., D.G., Bronk, D.A., Mulholland, M.R., and Carpenter, E.J., eds. Chapter 1, 2nd Edition, San Diego, CA, Academic Press.
Gruber, N., and Sarmiento, J.L. (2002). Large-Scale Biogeochemical–Physical Interactions in Elemental Cycles. In The Sea: Biological-Physical Interactions. Robinson, A.R., McCarthy, J.F., Rothschild, B., eds. (Wiley, New York), Vol. 12, pp. 337-399.
Gruber, N., and Galloway, J.N. (2008). An Earth-system perspective of the global nitrogen cycle. Nature, 451: 293 -296.Google Scholar
Gruber, N., C., Hauri, and G-K., Plattner. (2009). High vulnerability of eastern boundary upwelling systems to ocean acidification. Global Change Newsletter (IGBP), Issue No. 73.Google Scholar
Guidi, L., Degl'Innocenti, E., Remorini, D., Biricolti, S., Fini, A., Ferrini, F., Nicese, F.P., and Tattini, M. (2011). The impact of UV-radiation on the physiology and biochemistry of Ligustrum vulgare exposed to different visible-light irradiance. Environmental and Experimental Botany, 70, 88-95.Google Scholar
Ha, S-Y., Joo, H-M., Kang, S-H., Ahn, I-Y., and Shin, K-H. (2014). Effect of ultraviolet irradiation on the production and composition of fatty acids in plankton in a sub-Antarctic environment. Journal of Oceanography, 70: 1-10. DOI 10.1007/ s10872-013-0207-3.Google Scholar
Ha, S-Y., Kim, Y-N., Park, M-O., Kang, S-H., Kim, H-C. and Shin, K-H. (2012). Production of mycosporine-like amino acids of in situ phytoplankton community in Kongsfjorden, Svalbard, Arctic. Journal of Photochemistry and Photobiology, B: Biology, 114, 1-14.Google Scholar
Haddad, N.M., Crutsinger, G.M., Gross, K., Haarstad, J., and Tilman, D. (2011). Plant diversity and the stability of foodwebs. Ecology Letters, 14: 42–46.Google Scholar
Häder, D.-P., Kumar, H.D., Smith, R.C., and Worrest, R.C. (2007). Effects of solar UV radiation on aquatic ecosystems and interactions with climate change. Photochemical and Photobiological Sciences, 6, 267-285.Google Scholar
Hall, J., Harrison, D.E., and Stammer, D., eds. (2010). Proceedings of OceanObs–09: Sustained Ocean Observations and Information for Society (Volumes 1 and 2), ESA Publication WPP-306, doi:10.5270/OceanObs09.
Hallegraeff, G.M. (2010). Ocean climate chang., phytoplankton community responses, and harmful algal blooms: a formidable predictive challenge. Journal of the Phycology, 46: 220−235.Google Scholar
Hamersley, M.R., Lavik, G., Woebken, D., Rattray, J.E., Lam, P., Hopmans, E.C., Damsté, J.S.S., Krüger, S., Graco, M., Gutiérrez, D., and Kuypers, M.M.M. (2007). Anaerobic ammonium oxidation in the Peruvian oxygen minimum zone. Limnology and Oceanography, 52 (3): 923-933.Google Scholar
Xu, H., Miao, X., and Wu, Q. (2006). High quality biodiesel production from a macroalga Chlorella protothecoides by heterotrophic growth in fermenters. Journal of Biotechnology 126:499-507.Google Scholar
Handy, R.D., Owen, R., and Valsami-Jones, E. (2008). The ecotoxicology of nanoparticles and nanomaterials: current status, knowledge gaps, challenges, and future needs. Ecotoxicology 17:315–325.Google Scholar
Harrison, J.A., Bouwman, A.F., Mayorga, E., and Seitzinger, S. (2010). Magnitudes and sources of dissolved inorganic phosphorus inputs to surface fresh waters and the coastal zone: a new global model. Global Biogeochemical Cycles, 24, GB1003. doi: 10.1029/2009GB003590.Google Scholar
Hassan, R., Scholes, R., and Ash, N., eds. (2005). Coastal ecosystems In Ecosystems and Human Well-being: Current State and Trends, Island Press.
Hay, S. (2006). Marine ecology: gelatinous bells may ring change in marine ecosystems. Current Biology 16(17): R679-82.Google Scholar
Hays, G.C., Richardson, A.J., and Robinson, C. (2005). Climate change and marine plankton. Trends in Ecology & Evolution, 20 (6): 337-344.Google Scholar
Hay, S. (2006). Marine Ecology: Gelatinous Bells May Ring Change in Marine Ecosystems. Current Biology, 16: 679-682.Google Scholar
Heck, Jr.K.L., Carruthers, T.J.B., Duarte, C.M., Hughes, A.R., Kendrick, G., Orth, R.J., and Williams, S.W. (2008). Trophic Transfers from Seagrass Meadows Subsidize Diverse Marine and Terrestrial Consumers. Ecosystems, 11: 1198–1210.Google Scholar
Henley, S.F. (2013). Climate-induced changes in carbon and nitrogen cycling in the rapidly warming Antarctic coastal ocean. (http://hdl.handle.net/1842/7626).
Henson, S.A., and Thomas, A.C. (2007). Phytoplankton Scales of Variability in the California Current System: 1. Interannual and Cross-Shelf Variability, Journal of Geophysical Research-Oceans, 112, C07017, doi:10.1029/2006JC004039.Google Scholar
Henson, S.A., Sarmiento, J.L., Dunne, J.P., Bopp, L., Lima, I., Doney, S.C., John, J., and Beaulieu, C. (2010). Detection of anthropogenic climate change in satellite records of ocean chlorophyll and productivity. Biogeosciences, 7: 621–640.Google Scholar
Henson, S., Lampitt, R., and Johns, D. (2012). Variability in phytoplankton community structure in response to the North Atlantic Oscillation and implications for organic carbon flux. Limnology and Oceanography, 57(6): 1591-1601.Google Scholar
Henson, S., Cole, H., Beaulieu, C., and Yool, A. (2013). The impact of global warming on seasonality of ocean primary production. Biogeosciences, 10: 4357–4369Google Scholar
Heumann, B.W. (2011). Satellite remote sensing of mangrove forests: Recent advances and future opportunities. Progress in Physical Geography, 35 (1): 87-108.Google Scholar
Hilligsøe, K.M., Richardson, K., Bendtsen, J., Sørensen, L-L., Nielsen, T.G., and Lyngsgaard, M.M. (2011). Linking phytoplankton community size composition with temperature, plankton food web structure and sea-air CO2 flux. Deep-Sea Research Part I: Oceanographic Research Papers, 58: 826–838.Google Scholar
Hoagland, P., and Scatasta, S. (2006). The Economic Effects of Harmful Algal Blooms. In Granéli, E., and Turner, J.T., eds. Ecology of Harmful Algae. Ecology Studies Series: Springer-Verlag Berlin Heidelber. 189: 391-402.
Hofmann, L.C., Bischof, K. (2014). Ocean acidification effects on calcifying macroalgae. Aquatic Biology, 22: 261-279.Google Scholar
Hood, M., ed. (2009). Ship-based Repeat Hydrography: A Strategy for a Sustained Global Program. UNESCO-IOC Technical Series, 89. IOCCP Reports, 17. ICPO Publication 142. (www.go-ship.org/Docs/IOCTS89_GOSHIP.pdf).
Hooper, D.U., Chapin III, F.S., Ewel, J.J., Hector, A., Inchausti, P., Lavorel, S., Lawton, J.H., Lodge, D.M., Loreau, M., Naeem, S., Schmid, B., Setälä, H., Symstad, A.J., Vandermeer, J., and Wardle, D.A. (2005). Effects of biodiversity on ecosystem functioning: A consensus of current knowledge. Ecological Monographs, 75: 3–35.Google Scholar
Howarth, R.W., et al. (1996). Regional nitrogen budgets and riverine N and P fluxes for the drainages to the North Atlantic Ocean: Natural and human influences. Biogeochemistry, 35, 75– 139.Google Scholar
Howarth, R.W. (2008). Coastal nitrogen pollution: A review of sources and trends globally and regionally. Harmful Alga., 8: 14–20.Google Scholar
Howarth, R.W., Chan, F., Conley, D.J., Garnier, J., Doney, S.C., Marino, R., and Billen, G. (2011). Coupled biogeochemical cycles: eutrophication and hypoxia in temperate estuaries and coastal marine ecosystems. Frontiers in Ecology and the Environment, 9(1): 18–26.Google Scholar
Howarth, R.W., Swaney, D.P., Billen, G., Garnier, J., Hong, B., Humborg, C., Johnes, P.J., Mörth, C-M., and Marino, R. (2012). Nitrogen fluxes from large watersheds to coastal ecosystems controlled by net anthropogenic nitrogen inputs and climate. Frontiers in Ecology and the Environment. 10(1): 37–43.Google Scholar
Hu, C., Feng, L., Lee, Z., Davis, C.O., Mannino, A., McClain, C.R., and Franz, B.A. (2012). Dynamic range and sensitivity requirements of satellite ocean color sensors: learning from the past. Applied Optics, 51 (25): 6045-6062.Google Scholar
Huete-Ortega, M., Marañón, E., Varela, M., Bode, A. (2010). General patterns in the size scaling of phytoplankton abundance in coastal waters during a 10-year time series. Journal of Plankton Research, 32: 1-14.Google Scholar
Hutchins, D.A., Mulholland, M.R., and Fu, F-X. (2009). Nutrient Cycles and Marine Microbes in a CO2-Enriched Ocean. Oceanography, 22: 128–145.Google Scholar
IOCCG. (1998). Minimum Requirements for an Operational Ocean-Colour Sensor for the Open Ocean. Report 1, Morel, A., ed. International Ocean-Colour Coordinating Group (IOCCG) Report 1, pp.48.
Irigoien, X., Huisman, J., and Harris, R.P. (2004). Global biodiversity patterns of marine phytoplankton and zooplankton. Nature, 429: 863-867.Google Scholar
Jang, C.J., Park, J., Park, T., and Yoo, S. (2011) Response of the ocean mixed layer depth to global warming and its impact on primary production: a case for the North Pacific Ocean. ICES Journal of Marine Science. doi:10.1093/icesjms/fsr064.
Jansen, T., Campbell, A., Brunel, T., Clausen, L.M.. (2013). Spatial Segregation within the Spawning Migration of North Eastern Atlantic Mackerel (Scomber scombrus) as Indicated by Juvenile Growth Patterns. PloS One, doi: 10.1371/journal. pone.0058114.
Jia, M., Wang, Z., Liu, D., Ren, C., Tang, X., and Dong, Z. (2013). Monitoring Loss and Recovery of Salt Marshes in the Liao River Delta, China. Journal of Coastal Research (in-press). 31: 371-377. Doi: http://dx.doi.org/10.2112/JCOASTRES-D-13-00056.1.Google Scholar
Jiang, H., and Qiu, B. (2011). Inhibition of photosynthesis by UV-B exposure and its repair in the bloom-forming cyanobacterium Microcystis aeruginosa. Journal of Applied Phycology, 23 (4): 691-696.Google Scholar
Johnson, K.S., and Coletti, L.J. (2002). In situ ultraviolet spectrophotometry for high resolution and long-term monitoring of nitrate, bromide and bisulfide in the ocean. Deep-Sea Research Part I: Oceanography Research Papers, 49: 1291-1305.Google Scholar
Kaldy, J.E., and Dunton, K.H. (2000). Above-and below-ground production, biomass and reproductive ecology of Thalassia testudinum (turtle grass) in a subtropical coastal lagoon. Marine Ecology Progress Series, 193: 271-283.Google Scholar
Kamykowski, D. (1987). A preliminary biophysical model of the relationship between temperature and plant nutrients in the upper ocean. Deep Sea Research Part A. Oceanographic Research Papers, 34 (7): 1067–1079.Google Scholar
Karl, D.M., Christian, J.R., Dore, J.E., Hebel, D.V., Letelier, R.M., Tupas, L.M., and Winn, C.D. (1996). Seasonal and interannual variability in primary production and particle flux at Station ALOHA. Deep Sea Research Part II: Topical Studies in Oceanography, 43: 539-568.Google Scholar
Karl, D., Michaels, A., Bergman, B., Capone, D., Carpenter, E., Letelier, R., Lipschultz, F., Paerl, H., Sigman, D., and Stal, L. (2002). Dinitrogen fixation in the world's oceans. Biogeochemistry, 57/58: 47–98.Google Scholar
Keeling, R.F., Körtzinger, A., and Gruber, N. (2010). Ocean deoxygenation in a warming world. Annual Reviews Marine Science, 2: 199–229.Google Scholar
Kemp, W.M., Lewis, M.R., Jones, T.W. (1986). Comparison of methods for measuring productlon by the submersed macrophyte, Potamogeton perfoliatus. Limnology and Oceanography, 31: 1322-1334.Google Scholar
Kemp, W.M., Testa, J.M., Conley, D.J., Gilbert, D., and Hagy, J.D. (2009). Temporal responses of coastal hypoxia to nutrient loading and physical controls, Biogeosciences, 6: 2985–3008.Google Scholar
Kemp, W.M., Batiuk, R., Bartleson, R., Bergstrom, P., Carter, V., Gallegos, G., Hunley, W., Karrh, L., Koch, E., Landwehr, J., Moore, K., Murray, L., Naylor, M., Rybicki, N., Stevenson, J.C., and Wilcox, D. (2004). Habitat Requirements for Submerged Aquatic Vegetation in Chesapeake Bay: Water Quality, Light Regime, and Physical-Chemical Factors. Estuaries, 27: 363–377.Google Scholar
Kideys, A.E. (2002). Fall and Rise of the Black Sea Ecosystem. Science, 297: 1482-1484.Google Scholar
Kiørboe, T. (1993). Turbulence, phytoplankton cell size, and the structure of pelagic food webs. Advances in Marine Biology 29:72.Google Scholar
Koch, E.W., Barbier, E.B., Silliman, B.R., Reed, D.J., Perillo, G.M.E., Hacker, S.D., Granek, E.F., Primavera, J.H., Muthiga, N., Polasky, S., Halpern, B.S., Kennedy, C.J., Kappel, C.V., and Wolanski, E. (2009). Non-linearity in ecosystem services: temporal and spatial variability in coastal protection. Frontiers in Ecology and the Environment, 7(1): 29–37, doi:10.1890/080126.Google Scholar
Koeller, P., Fuentes-Yaco, C., Platt, T., Sathyendranath, S., Richards, A., Ouellet, P., Orr, D., Skúladóttir, U., Wieland, K., Savard, L., and Aschan, M. (2009). Basin-Scale Coherence in Phenology of Shrimps and Phytoplankton in the North Atlantic Ocean. Science, 324:791-793.Google Scholar
Koeve, W., and Kähler, P.K. (2010). Heterotrophic denitrification vs. autotrophic anammox – quantifying collateral effects on the oceanic carbon cycle. Biogeosciences, 7: 2327–2337.Google Scholar
Kovacs, J.M., King, J.M.L., de Santiago, F.F., Flores-Verdugo., F. (2009). Evaluating the condition of a mangrove forest of the Mexican Pacific based on an estimated leaf area index mapping approach. Environmental Monitoring and Assessment, 157: 137-149.Google Scholar
Kovacs, J., Wang, M.J., Blanco-Correa, M.. (2001). Mapping mangrove disturbances using multi-date Landsat TM imagery. Environmental Management, 27: 763–776.Google Scholar
Krishnamurthy, A., Moore, J.K., Mahowald, N., Luo, C., Doney, S.C., Lindsay, K., Zender, C.S. (2009). Impacts of increasing anthropogenic soluble iron and nitrogen deposition on ocean biogeochemistry. Global Biogeochemical Cycles, 23, GB3016, doi:10.1029/2008GB003440.Google Scholar
Krishnamurthy, A., Moore, J.K., Mahowald, N., Luo, C., and Zender., C.S. (2010). Impacts of atmospheric nutrient inputs on marine biogeochemistry. Journal of Geophysical Research, 115: 13 pp., G01006, doi:10.1029/2009JG001115.Google Scholar
Kroeker, K.J., Kordas, R.L., Crim, R.N., and Singh, G.G. (2010). Meta-analysis reveals negative yet variable effects of ocean acidification on marine organisms. Ecology Letters, 13: 1419–1434.Google Scholar
Kroeze, C., and Seitzinger, S.P. (1998). Nitrogen inputs to rivers, estuaries and continental shelves and related nitrous oxide emissions in 1990 and 2050: a global model. Nutrient Cycling in Agroecosystems, 52: 195-212.Google Scholar
Laffoley, D., and Grimsditch, G. (2009). The management of natural coastal carbon sinks. Gland: IUCN.
Lam, P., Lavik, G., Jensen, M.M., van de Vossenberg, J., Schmid, M., Woebken, D., Gutiérrez, D., Amann, R., Jetten, M.S.M., and Kuypers, M.M.M. (2009). Revising the nitrogen cycle in the Peruvian oxygen minimum zone. Proceedings of the National Academy of Sciences (USA), 106: 4752-4757.Google Scholar
Lamarque, J.-F., Dentener, F., McConnell, J., et al. (2013). Multimodel mean nitrogen and sulfur deposition from the Atmospheric Chemistry and Climate Model Intercomparison roject (ACCMIP): evaluation of historical and projected future changes. Atmospheric Chemistry and Physics, 13: 7997-8018.Google Scholar
Landolfi, A., Dietze, H., Koeve, W., and Oschlies, A. (2013). Overlooked runaway feedback in the marine nitrogen cycle: the vicious cycle. Biogeosciences, 10: 1351–1363.Google Scholar
Landry, M.R., Barber, R.T., Bidigare, R.R., Chai, F., Coale, K.H., et al. (1997). Iron and grazing constraints on primary production in the central equatorial Pacific: An EqPac synthesis. Limnology and Oceanography, 42: 405-418.Google Scholar
Laufkötter, C., Vogt, M., and Gruber, N. (2013). Long-term trends in ocean plankton production and particle export between 1960-2006. Biogeosciences, 10: 7373–7393.Google Scholar
Laws, E.A., Falkowski, P.G., Smith, Jr.W.O., Ducklow, H., and McCarthy, J.J. (2000). Temperature effects on export production in the open ocean. Global Biogeochemical Cycles, 14: 1231–1246.Google Scholar
Lebrato, M., and Jones, D.O.B. (2011). Jellyfish biomass in the biological pump: Expanding the oceanic carbon cycle. The Biochemical Society Journal, p. 35-39.Google Scholar
Lee, Z.-P., ed. (2006). Remote sensing of inherent optical properties: Fundamentals, Tests of algorithms, and applications. IOCCG Report No. 5, 126 pp., Dartmouth, Canada.
Legendre, L., and Rassoulzadegan, F. (1996). Food-web mediated export of biogenic carbon in oceans: hydrodinamic control. Marine Ecology Progress Series, 145: 179–193.Google Scholar
Le Quéré, C., Takahashi, T., Buitenhuis, E.T., Rödenbeck, C., and Sutherland, S.C. (2010). Impact of climate change and variability on the global oceanic sink of CO2. Global Biogeochemical Cycles, 24 (4): GB4007, doi: 10.1029/2009GB003599.Google Scholar
Licandro, P., Conway, D.V.P., Daly Yahia, M.N., Fernandez de Puelles, M.L., Gasparini, S., Hecq, J.H., Tranter, P., Kirby, R.R. (2010). A blooming jellyfish in the northeast Atlantic and Mediterranean. Biology Letters, doi:10.1098/rsbl.2010.0150.
Lilly, E.L., Halanych, K.M., and Anderson, D.M. (2007). Species boundaries and global biogeography of the Alexandrium tamarense complex (Dinophyceae). Journal of Phycology, 43: 1329–1338.Google Scholar
Liu, J., Liu, S., Loveland, T.R., Tieszen, L.L. (2008). Integrating remotely sensed land cover observations and a biogeochemical model for estimating forest ecosystem carbon dynamics. Ecological Modelling, 219: 361-372.Google Scholar
Lohrenz, S.E., Fahnensteil, G.L., Redalje, D.G., Lang, G.A., Chen, X., and Dagg, M.J. (1997). Variations in primary production of northern Gulf of Mexico continental shelf waters linked to nutrient inputs from the Mississippi River. Marine Ecology Progress Series, 155: 45-54.Google Scholar
Lomas, M.W., Bates, N.R., Johnson, R.J., Knap, A.H., Steinberg, D.K., and Carlson, C.A. (2013). Two decades and counting: 24-years of sustained open ocean biogeochemical measurements in the Sargasso Sea. Deep Sea Research Part II: Topical Studies in Oceanography. 93: 16-32. doi: 10.1016/j.dsr2.2013.01.008.Google Scholar
Long, S., Jones, M.B., and Roberts, M.J., eds. (1992). Primary Productivity of Grass Ecosystems of the Tropics and Sub-tropics. Chapman and Hall, London, 267 pp.
Luo, Y.-W., Doney, S.C., Anderson, L.A., Benavides, M., Bode, A., et al. (2012). Database of diazotrophs in global ocean: abundances, biomass and nitrogen fixation rates. Earth System Science Data, 4: 47-73.Google Scholar
Lynam, C.P., Gibbons, M.J., Axelsen, B.E., Sparks, C.A.J., Coetzee, J., Heywood, B.G., Brierley, A.S. (2006). Jellyfish overtake fish in a heavily fished ecosystem. Curr. Biol., 16: 492−493.Google Scholar
Ma, S., Tao, Z., Tang, X., Yu, Y., Zhou, X., Ma, W., and Li, Z. (2014). Estimation of Marine Primary Productivity From Satellite-Derived Phytoplankton Absorption Data. IEEE Journal of Selected Topics in Applied Earth Observations and Remote Sensing, 7 (7): 3084-3092.Google Scholar
Mackas, D.L. (2011). Does blending of chlorophyll data bias temporal trend? Nature, 472: E4-E5.Google Scholar
Mackas, D.L., and Beaugrand, G. (2010). Comparisons of zooplankton time series. Journal of Marine Systems, 79: 286-304.Google Scholar
Mackas, D.L., Batten, S.D., and Trudel, M. (2007). Effects on zooplankton of a warming ocean: recent evidence from the Northeast Pacific. Progress in Oceanography, 75(2): 223–252.Google Scholar
MacKenzie, R.A., and Dionne, M. (2008). Habitat heterogeneity: importance of salt marsh pools and high marsh surfaces to fish production in two Gulf of Maine salt marshes. Marine Ecology Progress Series, 368: 217–230.Google Scholar
Mahaffey, C., Michaels, A.F., Capone, D.G. (2005). The conundrum of marine N2 fixation. American Journal of Science, 305 (6-8): 546–595. doi: 10.2475/ajs.305.6-8.546.Google Scholar
Mahowald, N.M., Baker, A.R., Bergametti, G., Brooks, N., Duce, R.A., Jickells, T.D., Kubilay, N., Prospero, J.M., and Tegen, I. (2005). Atmospheric global dust cycle and iron inputs to the ocean, Global Biogeochemical Cycles, 19, GB4025, doi:10.1029/2004GB002402.Google Scholar
Mahowald, N., Jickells, T.D., Baker, A., Artaxo, P., Benitez-Nelson, C.R., et al. (2008). Global distribution of atmospheric phosphorus sources, concentrations and deposition rates, and anthropogenic impacts, Global Biogeochemical Cycles, 22, GB4026, doi:10.1029/2008GB003240.Google Scholar
Malone, T.C. (1971). Diurnal rhythms in netplankton and nanoplankton assimilation ratios. Marine Biology, 10: 285-289.Google Scholar
Malone, T.C. (1980). Algal size and phytoplankton ecology. pp. 433-464. In: Morri., I., ed. The Physiological Ecology of Phytoplankton, Blackwell Scientific Publications, London.
Malone, T.C. (1991). River flo., phytoplankton production and oxygen depletion in Chesapeake Bay, In Modern and Ancient Continental Shelf Anoxia, Tyson, R.V., and Pearson, T.H., eds. Geological Society Publication No. 58, p. 83-94.
Malone, T.C. (2008). Ecosystem dynamic., harmful algal blooms and operational oceanography. In Babin, M., Roesler, C.S., and Cullen, J.J., eds. Real-Time Coastal Observing Systems for Marine Ecosystem Dynamics and Harmful Algal Blooms, Monographs on Oceanographic Methodology Series, UNESCO, pp. 527-560.
Malone, T.C., Conley, D.J., Fisher, T.R., Glibert, P.M., Harding, L.W., Sellner, K. (1996). Scales of Nutrient-Limited Phytoplankton Productivity in Chesapeake Bay. Estuaries, 19 (2B): 371-385.Google Scholar
Malone, T.C., DiGiacomo, P.M., Gonçalves, E., Knap, A.H., Talaue-McManus, L., and de Mora, S. (2014a). A Global Ocean Observing System Framework for Sustainable Development. Marine Policy, 43:262-272.Google Scholar
Malone, T.C., DiGiacomo, P.M., Gonçalves, E.J., Knap, A.H., Talaue-McManus, L., de Mora, S., and Muelbert., J. (2014b). Enhancing the Global Ocean Observing System to meet evidence based needs for ecosystem-based management of coastal ecosystem services. Natural Resources Forum, 38: 168–181.Google Scholar
Malone, T.C., Hopkins, T.S., Falkowski, P.G., and Whitledge, T.E. (1983). Production and transport of phytoplankton biomass over the continental shelf of the New York Bight. Continental Shelf Research, 1: 305-337.Google Scholar
Malone, T.C., Malej, A., Harding, L.W., Smodlaka, N., eds. (1999). Ecosystems at the Land-Sea Margin: Drainage Basin to the Coastal Sea. AGU, Coastal and Estuarine Studies, No. 55: 377 p.
Mann, K.H. (1972). Ecological energetics of the sea-weed zone in a marine bay on the Atlantic coast of Canada II. Productivity of the seaweeds. Marine Biology, 14, 199-209.Google Scholar
Marañón, E. (2009). Phytoplankton Size Structure. In Encyclopedia of Ocean Science, 2nd Edition, Steele, J.H., Turekian, K.K., and Thorpe, S.A., eds. Academic Press.
Marañón, E., Behrenfeld, M.J., González, N., Mouriño, B., and Zubkov, M.V. (2003). High variability of primary production in oligotrophic waters of the Atlantic Ocean: uncoupling from phytoplankton biomass and size structure. Marine Ecology Progress Series, 257: 1-11Google Scholar
Margalef, R. (1978). Life-forms of phytoplankton as survival alternatives in an unstable environment. Oceanologica Acta, 1: 493-509.Google Scholar
Marra, J., and Ducklow, H.W. (1995). Primary Production in the North Atlantic: Measurements, Scaling, and Optical Determinants. Philosophical Transactions: Biological Sciences, Royal Society London B, 348: 153-160.
Marra, J. (2002). Approaches to the Measurement of Plankton Production. pp. 78-10., In Williams, P.J.L., Thomas, D.N., and Reynolds, C.S., eds. Phytoplankton Productivity: Carbon Assimilation in Marine and Freshwater Ecosystems, Blackwell Science, Oxford.
Martin, J.H., Coale, K.H., Johnson, K.S., Fitzwater, S.E., et al. (1994). Testing the iron hypothesis in ecosystems of the equatorial Pacific Ocean. Nature, 371: 123-129.Google Scholar
Martínez, M.L., Intralawan, A., Vázquez, G., Pérez-Maqueo, O., Sutton, P., and Landgrave, R. (2007). The coasts of our world: Ecological, economic and social importance. Ecological Economics 63: 254-272.Google Scholar
Martiny, A.C., Pham, C.T.A., Primeau, F.W., Vrugt, J.A., Moore, J.K., Levin, S.A., Lomas, M.W. (2013), Strong latitudinal patterns in the elemental ratios of marine plankton and organic matter. Nature geosciences, 6: 279-283. doi: 10.1038/ ngeo1757.Google Scholar
Mathis, J.T., and Feely, R.A. (2013). Building an integrated coastal ocean acidification monitoring network in the U.S. Elementa: Science of the Anthropocene, 1: 000007 doi: 10.12952/journal.elementa.000007.Google Scholar
McCann, K.S. (2000). The diversity–stability debate. Nature, 405: 228-233.Google Scholar
McClain, C.R. (2009). A Decade of Satellite Ocean Color Observations. Annual Reviews Marine Science, 1: 19-42.Google Scholar
McGregor, H.V., Dima, M., Fischer, H.W., and Mulitza, S. (2007). Rapid 20th-century increase in coastal upwelling off northwest Africa. Science, 315: 637-639.Google Scholar
McLeod, K., and Leslie, H., eds. (2009). Ecosystem-Based Management for the Oceans. Washington, DC: Island Press. 392 pp.
McLeod, E., Chmura, G.L., Bouillon, S., Salm, R., Björk, M., Duarte, C.M., Lovelock, C.E., Schlesinger, W.H., and Silliman, B.R. (2011). A blueprint for blue carbon: toward an improved understanding of the role of vegetated coastal habitats in sequestering CO2. Frontiers in Ecology and the Environment 9: 552-560.Google Scholar
McQuatters-Gollop, A., Reid, P.C., Edwards, M., Burkill, P.H., Castellani, C., Batten, S., Gieskes, W., et al. (2011). Is there a decline in marine phytoplankton? Nature, 472: E6-E7.Google Scholar
Meier, M.F., Dyurgerov, M.B., Rick, U.K., O'Neel, S., Pfeffer, W.T., Anderson, R.S., Anderson, S.P., and Glazovsky, A.F. (2007). Glaciers Dominate Eustatic Sea-Level Rise in the 21st Century. Science, 317 (5841): 1064-1067.Google Scholar
Meybeck, M. (1982). Carbon, Nitrogen, and Phosphorus Transport by World Rivers. American Journal of Science, 282: 401-450.Google Scholar
Millennium Ecosystem Assessment. (2005). Ecosystems and Human Well-being: Synthesis. Island Press.
Miller, R.J., Bennett, S., Keller, A.A., Pease, S., and Lenihan, H.S. (2012). TiO2 Nanoparticles are phototoxic to marine phytoplankton. PLoS ONE 7 (1): e30321. doi:10.1371/journal.pone.0030321.Google Scholar
Mills, M.M., and Arrigo, K.R. (2010). Magnitude of oceanic nitrogen fixation influenced by the nutrient uptake ratio of phytoplankton. Nature Geoscience, 3: 412-416.Google Scholar
Mishra, D.R., Cho, H.J., Ghosh, S., Fox, A., Downs, C., Merani, P.B.T., Kirui, P., et al. (2012). Post-spill state of the marsh: Remote estimation of the ecological impact of the Gulf of Mexico oil spill on Louisiana Salt Marshes. Remote Sensing of Environment, 118: 176-185.Google Scholar
Mitsch, W.J., and Gosselink, J.G. (2008). Wetlands. Van Nostrand Reinhold, New York, New York, USA.
Moloney, C.L., St John, M.A., Denman, K.L., Karl, D.M., Köster, F.W., Sundby, S., and Wilson, R.P. (2011). Weaving marine food webs from end to end under global change. Journal of Marine Systems, 84: 106-116.Google Scholar
Montoya, J.M., and Raffaelli, D. (2010). Climate change, biotic interactions and ecosystem services. Philosophical Transactions of the Royal Society B, 365, doi: 10.1098/rstb.2010.0114.Google Scholar
Moore, C.M., Mills, M.M., Arrigo, K.R., Berman-Frank, I., Bopp, L., Boyd, P.W., Galbraith, E.D., Geider, R.J., Guieu, C., Jaccard, S.L., Jickells, T.D., La Roche, J., Lenton, T.M., Mahowald, N.M., Marañón, E., Marinov, I., Moore, J.K., Nakatsuka, T., Oschlies, A., Saito, M.A., Thingstad, T.F., Tsuda, A., Ulloa, O. (2013). Processes and patterns of oceanic nutrient limitation. Nature Geoscience, 6: 701-710. doi: 10.1038/ ngeo/1765.Google Scholar
Moore, J.K., Doney, S.C., Lindsay, K., Mahowald, N., Michaels, A.F. (2006). Nitrogen fixation amplifies the ocean biogeochemical response to decadal timescale variations in mineral dust deposition. Tellus, Series B. 58: 560-572. doi:10.1111/j.1600-0889.2006.00209.x.Google Scholar
Moore, S.K., Trainer, V.L., Mantua, N.J., Parker, M.S., Laws, E.A., Backer, L.C., and Fleming, L.E. (2008). Impacts of climate variability and future climate change on harmful algal blooms and human health. Environmental Health, 7(Suppl 2): S4.Google Scholar
Morán, X.A.G., López-Urrutia, A., Calvo-Díaz, A., and Li, W.K. (2010). Increasing importance of small phytoplankton in a warmer ocean. Global Change Biology, 16: 1137–1144.Google Scholar
Morel, A., and Berthon, J-F. (1989). Surface pigments, algal biomass profiles, and potential production of the euphotic layer: Relationships reinvestigated in view of remote-sensing applications. Limnology and Oceanography, 34: 1545-1562.Google Scholar
Morris, J.T. (2007). Ecological engineering in intertidal saltmarshes. Hydrobiologia, 577: 161-168.Google Scholar
Moy, A.D., Howard, W.R., Bray, S.G., and Trull, T.W. (2009). Reduced calcification in modern Southern Ocean planktonic foraminifera. Nature Geoscience, 2: 276-280.Google Scholar
Muller-Karger, F.E., Varela, R., Thunell, R.C., Luerssen, R., Hu, C., Walsh, J.J. (2005). The importance of continental margins in the global carbon cycle. Geophysical Research Letters, 32, L01602, doi:10.1029/2004GL021346.Google Scholar
Mustapha, S.B., Bélanger, S., and Larouche, P. (2012). Evaluation of ocean color algorithms in the southeastern Beaufort Sea, Canadian Arctic: New parameterization using SeaWiFS, MODIS, and MERIS spectral bands. Canadian Journal of Remote Sensing, 38(05): 535-556.Google Scholar
Naqvi, W. (2012). Marine nitrogen cycle. The Encyclopedia of Earth. (http://www.eoearth.org/view/article/154479).
Naqvi, S.W.A., Voss, M., and Montoya, J.P. (2008). Recent advances in the biogeochemistry of nitrogen in the ocean. Biogeosciences, 5: 1033–1041.Google Scholar
National Research Council. (2004). Groundwater Fluxes across Interfaces. Washington D.C.: National Academies Press.
Negreanu-Pîrjol, B., Negreanu-Pîrjol, T., Paraschiv, G., Bratu, M., Sîrbu, R., Roncea, F., and Meghea, A. (2011). Physical-chemical characterization of some green and red macrophyte algae from the Romanian Black Sea littoral. St. Cerc. St. CICBIA, 12 (2): 173-184.Google Scholar
Nellemann, C., Corcoran, E., Duarte, C.M., Valdés, L., De Young, C., Fonseca, L., Grimsditch, G., eds. (2009). Blue Carbon. A Rapid Response Assessment. United Nations Environment Programme, GRID-Arendal, http://www.grida.no.
Nicholls, R.J., and Cazenave, A. (2010). Sea-Level Rise and its Impact on Coastal Zones. Science, 328 (5985): 1517–1520.Google Scholar
Nixon, S.W. (1992). Quantifying the relationship between nitrogen input and the productivity of marine ecosystems. In Takahash., M., Nakata, K., and Parsons, T.R., eds. Proceedings of Advanced Marine Technology Conference (AMTEC), Vol. 5: 57-83.Google Scholar
Ohashi, R., Yamaguchi, A., Matsuno, K., Saito, R., Yamada, N., Iijima, A., Shiga, N., and Imai, I. (2013). Interannual changes in the zooplankton community structure on the southeastern Bering Sea shelf during summers of 1994–2009. Deep Sea Research Part II: Topical Studies in Oceanography, 94: 44-56.Google Scholar
Okin, G.S., Baker, A.R., Tegen, I., Mahowald, N.M., Dentener, F.J., Duce, R.A., et al. (2011). Impacts of atmospheric nutrient deposition on marine productivity: Roles of nitrogen, phosphorus, and iron. Global Biogeochemical Cycles, 25, GB2022, doi:10.1029/2010GB003858.Google Scholar
Onabid, M.A. (2011). Improved ocean chlorophyll estimate from remote sensed data: The modified blending technique. African Journal of Environmental Science and Technology, 5(9): 732-747.Google Scholar
O'Reilly, J.E., Busch, D.A. (1983). Phytoplankton primary production for the Northwestern Atlantic shelf. Rapports et procès-verbaux des reunions, Conseil permanent international pour l'exploration de la mer, 183: 255-268.Google Scholar
O'Reilly, J.E., Maritorena, S., Mitchell, B.G., Siegel, D.A., Carder, K.L., Garver, S.A., Kahru, M., and McClain, C. (1998). Ocean color chlorophyll algorithms for SeaWiFS. Journal of Geophysical Research, 103 (C11): 24937-24953.Google Scholar
Orhan, I., Sener, B., Atici, T., Brun, R., Perozzo, R., and Tasdemir, D. (2006). Turkish freshwater and marine macrophyte extracts show in vitro antiprotozoal activity and inhibit Fabl, a key enzyme of Plasmodium falciparum fatty acid biosynthesis. Phytomedicine, 13(6): 388-393.Google Scholar
Orr, J.C., V.J., Fabry, O., Aumont. L., Bopp, S.C., Doney, R.A., Feely, A., Gnanadesikan, N., Gruber, A., Ishida, F., Joos, R.M., Key, K., Lindsay, E., Maier-Reimer, and others. (2005). Anthropogenic ocean acidification over the twenty-first century and its impact on calcifying organisms. Nature, 437: 681-686.Google Scholar
Orth, R.J., Carruthers, T.J.B., Dennison, W.C., Duarte, C.M., Fourqurean, J.W., Heck, Jr.K.L., Hughes, A.R., Kendrick, G.A., Kenworthy, W.J., Olyarnik, S., Short, F.T., Waycott, M., and Williams, S.L. (2006). A global crisis for seagrass ecosystems. BioScience 56:987-996.Google Scholar
Oschlies, A., Schulz, K.G., Riebesell, U., and Schmittner, A. (2008). Simulated 21st century's increase in oceanic suboxia by CO2-enhanced biotic carbon export. Global Biogeochemical Cycles, 22, GB4008, doi:10.1029/2007GB003147.Google Scholar
Passow, U., and Carlson, C.A. (2012). The biological pump in a high CO2 world. Marine Ecology Progress Series, 470: 249-271.Google Scholar
Paulmier, A., and Ruiz-Pino, D. (2009). Oxygen minimum zones (OMZs) in the modern ocean. Progress in Oceanography, 80: 113-28.Google Scholar
Pauly, D., and Christensen, V. (1995). Primary production required to sustain global fisheries. Nature, (374): 255-257.Google Scholar
Paytan, A., and McLaughlin, K. (2007). The oceanic phosphorus cycle. Chemical Reviews, 107, 563–576.Google Scholar
Peierls, B.L., Caraco, N.F., Pace, M.L., and Cole, J.J. (1991). Human influence on river Nitrogen. Nature, 350: 386–87.Google Scholar
Pennington, J.T., Mahoney, K.L., Kuwahara, V.S., Kolber, D.D., Calienes, R., and Chavez, F.P. (2006). Primary production in the eastern tropical Pacific: A review. Progress in Oceanography, 69: 285–317.Google Scholar
Pennock, J.R., and Sharp, J.H. (1986). Phytoplankton production in the Delaware Estuary: temporal and spatial variability. Marine Ecology Progress Series, 34: 143-155.Google Scholar
Pereira, H., Barreira, L., Figueiredo, F., Custódio, L., Vizetto-Duarte, C., Polo, C., Rešek, E., Engelen, A., and Varela, J. (2012). Polyunsaturated Fatty Acids of Marine Macroalgae: Potential for Nutritional and Pharmaceutical Applications. Marine Drugs, 10(9): 1920-1935.Google Scholar
Pernetta, J.C. (1993). Mangrove forest., climate change and sea level rise: hydrological influences on community structure and survival, with examples from the Indo- West Pacific. Marine Conservation Development Report, International Union for Conservation of Nature, Gland, Switzerland.
Perrings, C., Naeem, S., Ahrestani, F., Bunker, D.E., Burkill, P., et al. (2010). Ecosystem services for 2020. Science, 330: 323-324.Google Scholar
Perry, M.J. (1986). Assessing marine primary production from space. BioScience, 36 (7): 461-467.Google Scholar
PICES. 2004. Marine Ecosystems of the North Pacific. PICES Special Publication 1, 280 p.
Pikitch, E.K., Santora, C., Babcock, E.A., Bakun, A., Bonfil, R., et al. (2004). Ecosystembased fishery management. Science, 305: 346–347.Google Scholar
Pilskaln, C.H., Paduan, J.B., Chavez, F.P., Anderson, R.Y., and Berelson, W.M. (1996). Carbon export and regeneration in the coastal upwelling system of Monterey Bay, central California. Journal of Marine Research, 54: 1149-1178.Google Scholar
Pingree, R.D., Pugh, P.R., Holligan, P.M., and Forster, G.R. (1975). Summer phytoplankton blooms and red tides along tidal fronts in the approaches to the English Channel. Nature, 258: 672-677.Google Scholar
Pitt, K.A., Welsh, D.T., Condon, R.H. (2009). Influence of jellyfish blooms on carbon, nitrogen and phosphorus cycling and plankton production. Hydrobiologia, 616 (1): 133-149.Google Scholar
Platt, T., and Sathyendranath, S. (1993). Estimators of primary production for the interpretation of remotely sensed data on ocean color. Journal of Geophysical Research, 98: 14561-14576.Google Scholar
Platt, T., Fuentes-Yaco, C., and Frank, K. (2003). Marine ecology: Spring algal bloom and larval fish survival. Nature, 423: 398-399.Google Scholar
Platt, T., Hoepffner, N., Stuart, V., and Brown, C., eds. (2008). Why ocean colour? The Societal benefits of ocean-colour technology. IOCCG Report, No. 7, 141 pp.Google Scholar
Platt, T., White III, G.N., Zhai, L., Sathyendranath, S., and Roy, S. (2009). The phenology of phytoplankton blooms: Ecosystem indicators from remote sensing. Ecological Modelling, 220 (21): 3057-3069.Google Scholar
Polovina, J.J., Howell, E.A., and Abecassis, M. (2008). The ocean's least productive waters are expanding. Geophysical Research Letters, 35, L03618, http://dx.doi.org/10.1029/2007GL031745Google Scholar
Polovina, J.J., Dunne, J.P., Woodworth, P.A., Howell, E.A. (2011). Projected expansion of the subtropical biome and contraction of the temperate and equatorial upwelling biomes in the North Pacific under global warming. ICES Journal of Marine Science, 68 (6), 986–995.Google Scholar
Polovina, J.J., and Woodworth, P.A. (2012). Declines in phytoplankton cell size in the subtropical oceans estimated from satellite remotely-sensed temperature and chlorophyll, 1998–2007. Deep Sea Research Part II: Topical Studies in Oceanography, 77–80, 82-88.Google Scholar
Pomeroy, L.R., Williams, P.J.leB., Azam, F., and Hobbie, J.E. (2007). The microbial loop. Oceanography, 20 (2): 28-33.Google Scholar
Powell, T.M., (1989). Physical and biological scales of variability in lakes, estuaries, and the coastal ocean, p. 157-180. In Roughgarden, J., May, R.M., and Levin, S.A., eds. Perspectives in Ecological Theory, Princeton University Press, Princeton N.J.
Plus, M., Deslous-Paoli, J-M., Auby, I., and Dagault, F. (2001). Factors influencing primary production of seagrass beds (Zostera noltii Hornem.) in the Thau Lagoon (French Mediterranean coast). Journal of Experimental Marine Biology and Ecology, 259:63–84.Google Scholar
Primeau, F.W., and Holzer, M. (2006). The Ocean's Memory of the Atmosphere: Residence-Time and Ventilation-Rate Distributions of Water Masses. Journal of Physical Oceanography, 36: 1439-1456.Google Scholar
Purcell, J.E., (2012). Jellyfish and ctenophore blooms coincide with human proliferations and environmental perturbations. Annual Reviews, Marine Science, 4: 209-235.Google Scholar
Rabalais, N.N., Turner, R.E., Díaz, R.J., and Justíc, D. (2009). Global change and eutrophication of coastal waters. ICES Journal of Marine Science, 66: 1528–1537.Google Scholar
Rahmstorf, S., Foster, G., and Cazenave, A. (2012). Comparing climate projections to observations up to 2011. Environmental Research Letters, 7: 1-5.Google Scholar
Raven, J.A. (1998). Small is beautiful: The picophytoplankton. Functional Ecology, 12: 503–513.Google Scholar
Redfield, A.C., Ketchum, B.H., and Richards, F.A. (1963). The influence of organisms on the composition of sea-water. In The Sea, Vol. 2 M.N., Hill (ed.), pp. 26-77. Interscience, New York.
Richardson, A.J. (2008). In hot water: zooplankton and climate change. ICES Journal of Marine Science, 65, 279-295.Google Scholar
Richardson, A.J., Bakun, A., Hays, G.C., Gibbons, M.J. (2009). The jellyfish joyride: causes, consequences and management responses to a more gelatinous future. Trends in Ecology and Evolution 24(6), 312-322.Google Scholar
Richardson, A.J., and Shoeman, D.S. (2004). Climate impact on plankton ecosystems in the Northeast Atlantic. Science, 305: 1609-1612.Google Scholar
Rockstram, J., Steffen, W., Noone, K., Persson, A., Chapin, F.S.III., Lambin, E., Lenton, T.M., Scheffer, M., Folke, C., Schellnhuber, H.J., Nykvist, B., de Wit, B.A., Hughes, T., van der Leeuw, S., Rodhe, H., Sorlin, S., Snyder, P.K., Coastanza, R., Sve-Ecoldin, U., Falkenmark, M., Karlberg, L., Corell, R.W., Fabry, V.J., Hansen, J., Walker, B., Liverman, D., Richardson, K., Crutzen, P., Foley, J. (2009). Planetary Boundaries: Exploring the Safe Operating Space for Humanity. Ecology and Society, 14: 32. [online] URL: http://www.ecologyandsociety.org/vol14/iss2/art32/.Google Scholar
Rockström, J., Steffen, W. Noone, K., Persson, Å., Chapin, F.S.III, Lambin, E.F., Lenton, T.M., Scheffer, M., Folke, C., Schellnhuber, H.J., Nykvist, B., de Wit, C.A., Hughes, T., van der Leeuw, S., Rodhe, H., Sörlin, S., Snyder, P.K., Costanza, R., Svedin, U., Falkenmark, M., Karlberg, L., Corell, R.W., Fabry, V.J., Hansen, J., Walker, B.H., Liverman, D., Richardson, K., Crutzen, P., and Foley, J.A. (2009). A safe operating space for humanity. Nature 461:472-475.Google Scholar
Roelfsema, C.M., Phinn, S.R., Udy, N., and Maxwell, P. (2009). An Integrated Field and Remote Sensing Approach for Mapping Seagrass Cover, Moreton Bay, Australia. Spatial Science, 54 (1): 45-62.Google Scholar
Ross, M.S., Ruiz, P.L., Telesnicki, G.J., and Meeder, J.F. (2001). Estimating above-ground biomass and production in mangrove communities of Biscayne National Park, Florida (U.S.A.). Wetlands Ecology and Management, 9: 27–37.Google Scholar
Rost, B., Zondervan, I., and Wolf-Gladrow, D. (2008). Sensitivity of phytoplankton to future changes in ocean carbonate chemistry: current knowledge, contradictions and research directions. Marine Ecology Progress Series, 373: 227–237.Google Scholar
Rousseaux, C.S., and Gregg, W.W. (2014). Interannual variations in phytoplankton primary production at a global scale. Remote Sensing, 6: 1-19.Google Scholar
Rudnick, D.L., Davis, R.E., Eriksen, C.C., Fratantoni, D.M., and Perry, M.J. (2004). Underwater gliders for ocean research. Marine Technology Society Journal, 38(1): 48-59.Google Scholar
Ruttenberg, K.C. (2004). The global phosphorus cycle. In: Hollan., H.D. and Turekian, K.K., eds. Treatise on Geochemistry. Elsevier, New York, NY, pp. 585–643.
Rykaczewski, R.R., and Dunne, J.P. (2011). A measured look at ocean chlorophyll trends. Nature, 472: E5-E6.Google Scholar
Saba, V.S., Friedrichs, M.A.M., Carr, M-E., Antoine, D., Armstrong, R.A., et al. (2010). Challenges of modeling depth-integrated marine primary productivity over multiple decades: A case study at BATS and HOT. Global Biogeochemical Cycles, 24, GB3020, doi:10.1029/2009GB003655.Google Scholar
Saba, V.S., Friedrichs, M.A.M., Antoine, D., Armstrong, R.A., Asanuma, I., Behrenfeld, M.J., Ciotti, A.M., Dowell, M., Hoepffner, N., Hyde, K.J.W., Ishizaka, J., Kameda, T., Marra, J., Mélin, F., Morel, A., O'Reilly, J., Scardi, M., Smith, W.O., Jr., Smyth, T.J., Tang, S., Uitz, J., Waters, K., and Westberry, T.K. (2011). An evaluation of ocean color model estimates of marine primary productivity in coastal and pelagic regions across the globe. Biogeosciences, 8, 489–503.Google Scholar
Sadeghi, A., Dinter, T., Vountas, M., Taylor, B., Altenburg-Soppa, M., and Bracher, A. (2012). Remote sensing of coccolithophore blooms in selected oceanic regions using the PhytoDOAS method applied to hyper-spectral satellite data. Biogeosciences, 9: 2127–2143.Google Scholar
Sakamoto, C.M., et al. (2004). Influence of Rossby waves on nutrient dynamics and the plankton community structure in the North Pacific subtropical gyre. Journal of Geophysical Research, 109, doi:10.1029/2003JC001976.Google Scholar
Sakshaug, E. (2004). Primary and secondary production in the Arctic Seas. In The Organic Carbon Cycle in the Arctic Ocean, Stein, R., and MacDonald, R.W., eds. pp. 57-81. Springer, Berlin.
Santos, A.L., Oliveira, V., Baptista, I., Henriques, I., Gomes, N.C.M., Almeida, A., Correia, A., and Cunha, A. (2012a). Effects of UV-B radiation on the structureal and physiological diversity of bacterioneuston and bacterioplankton. Applied and Environmental Microbiology, 78: 2066-2069.Google Scholar
Sarmento, H., Montoya, J.M., Vázquez-Domínguez, E., Vaqué, D., and Gasol, J.M. (2010). Warming effects on marine microbial food web processes: how far can we go when it comes to predictions? Philosophical Transactions of the Royal Society B, 365: 2137-2149.Google Scholar
Sathyendranath, S., ed. (2000). Remote sensing of ocean colour in coastal and other optically complex waters. Reports of the International Ocean-Colour Coordinating Group, No. 3, 140 p., Dartmouth, Canada.
Sathyendranath, S., Ahanhanzo, J., Bernard, S., Byfield, V., Delaney, L., and others. (2010). ChloroGIN: Use of satellite and in situ data in support of ecosystembased management of marine resources in Proceedings of OceanObs–09: Sustained Ocean Observations and Information for Society (Vol. 2), Venice, Italy, 21-25 September 2009, Hall, J., Harrison, D.E., & Stammer, D., eds. ESA Publication WPP-306.
Scavia, D., and Bricker, S.B. (2006). Coastal eutrophication assessment in the United States. Biogeochemistry, 79: 187–208.Google Scholar
Schlitzer, R., Usbeck, R., and Fischer, G. (2003). Inverse modeling of particulate organic carbon fluxes in the South Atlantic. In: The South Atlantic in the Late Quaternary: Reconstruction of material budget and current systems. Wefer, G., Mulitza, S., and Ratmeyer, V., eds. Springer, Berlin, Heidelberg, New York, pp. 1-19.
Schlitzer, R., Monfray, P., Hoepffner, N., Fischer, G., Gruber, N., Lampitt, R., Lévy, M., Schmidt, A.L., Wysmyk, J.K.C., Craig, S.E., and Lotze, H.K. (2012). Regional-scale effects of eutrophication on ecosystem structure and services of seagrass beds. Limnology and Oceanography. 57(5): 1389–1402.Google Scholar
Schmidt, A.L., Wysmyk, J.K.C., Craig, S.E., and Lotze, H.K. (2012). Regional-scale effects of eutrophication on ecosystem structure and services of seagrass beds. Limnology and Oceanography. 57(5): 1389–1402.Google Scholar
Schneider, B., Bopp, L., Gehlen, M., Segschneider, J., Frolicher, T.L., Cadule, P., Friedlingstein, P., Doney, S.C., Behrenfeld, M.J., and Joos, F. (2008). Climate-induced interannual variability of marine primary and export production in three global coupled climate carbon cycle models. Biogeosciences, 5: 597–614.Google Scholar
Schulz, M., Prospero, J.M., Baker, A.R., Dentener, F., et al. (2012). Atmospheric transport and deposition of mineral dust to the ocean: implications for research needs. Environmental Science & Technology, 46: 10,390-10,404.Google Scholar
Seitzinger, S.P., Harrison, J.A., Dumont, E., Beusen, A.H.W., Bouwman, A.F. (2005). Sources and delivery of carbon, nitrogen, and phosphorus to the coastal zone: An overview of the Global Nutrient Export from Watersheds (NEWS) models and their application, Global Biogeochemical Cycles, 19, GB4S01, doi:10.1029/2005GB002606.Google Scholar
Seitzinger, S.P., Mayorga, E., Bouwman, A.F., et al. (2010). Global nutrient export: A scenario analysis of past and future trends. Global Biogeochemical Cycles, 23, GB0A08, doi: 10.1029/2009GB003587.Google Scholar
Sharma, K.K., Schuhmann, H., and Schenk, P.M. (2012). High lipid induction in microalgae for biodiesel production. Energies, 5: 1532-1553.Google Scholar
Sharpies, J., Middelburg, J.J., Fennel, K., Jickells, T.D. (2013), Riverine delivery of nutrients and carbon to the oceans, Nature, 504:61-70. doi:10.1038/nature12857.Google Scholar
Sherman, K., and Hempel, G. (2009). The UNEP Large Marine Ecosystem Report: A perspective on changing conditions in LMEs of the world's Regional Seas. UNEP Regional Seas Reports and Studies, 182. United Nations Environmental Programme, Nairobi, 872 pp.
Sherman, K., Alexander, L.M., and Gold, B.D., eds. (1993). Large Marine Ecosystems: Stress, Mitigation, and Sustainability, AAAS Press, Washington D.C., 376 pp.
Sherman, K., Smith, W., Morse, W., Berman, M., Green, J., and Ejsymont, L. (1984). Spawning strategies of fishes in relation to circulation, phytoplankton production, and pulses in zooplankton off the northeastern United States. Marine Ecology Progress Series, 18: 1-19.Google Scholar
Shi, D., Xu, Y., and Morel, F.M.M. (2009). Effects of the pH/pCO2 control method on medium chemistry and phytoplankton growth. Biogeosciences, 6: 1199-1207.Google Scholar
Shi, D., Xu, Y., Hopkinson, B.M., Morel, F.M.M. (2010). Effect of ocean acidification on iron availability to marine phytoplankton. Science, 327(5966): 676-679.Google Scholar
Shi, D., Kranz, S.A., Kim, J-M., and Morel, F.M.M. (2012). Ocean acidification slows nitrogen fixation and growth in the dominant diazotroph Trichodesmium under low-iron conditions. Proceedings of the National Academy of Sciences, 109 (45): E3094-100.Google Scholar
Short, F.T., and Neckles, H.H. (1999). The effects of global climate change on seagrasses. Aquatic Botany, 63: 169-196.Google Scholar
Shurin, J.G., Gruner, D.S., and Hillebrand, H. (2006). All wet or dried up? Real differences between aquatic and terrestrial food webs. Proceedings of the Royal Society B, 273: 1-9.Google Scholar
Siegel, D.A., Behrenfeld, M.J., Maritorena, S., McClain, C.R., Antoine, D., and others. (2013). Regional to global assessments of phytoplankton dynamics from the Sea-WiFS mission. Remote Sensing of Environment 135: 77–91.Google Scholar
Siegel, D.A., Doney, S.C., and Yoder, J.A. (2002). The North Atlantic spring phytoplankton bloom and Sverdrup's critical depth hypothesis. Science, 296: 730– 733.Google Scholar
Silliman, B.R., Grosholz, T., and Bertness, M.D. (2009). Salt marshes under global siege. Pages 103–114 in Silliman, B.R., Grosholz, T., and Bertness, M.D., eds. Human impacts on salt marshes: a global perspective. University of California Press, Berkeley, California, USA.
Silva, J., Sharon, Y., Santos, R., and Beer, S. (2009). Measuring seagrass photosynthesis: methods and applications. Aquatic Biology, 7: 127–141.Google Scholar
Slomp, C.P. (2011). Phosphorus cycling in the estuarine and coastal zones: source., sinks, and transformations. In: Wolanski, E., and McLusky, D.S., eds. Treatise on Estuarine and Coastal Science, Vol 5, pp. 201–229. Waltham: Academic Press.
Smetacek, V. (1999). Diatoms and the ocean carbon cycle. Protist, 150 (1): 25-32.Google Scholar
Smith, R.C., Baker, K.S., Dierssen, H.M., Stammerjohn, S.E., and Vernet, M. (2001). Variability of primary production in an Antarctic marine ecosystem as estimated using a multi-scale sampling strategy. American Zoology, 41: 40–56.Google Scholar
Smith, S.V. (1981). Marine macrophytes as a global carbon sink. Science, 211: 838-840.Google Scholar
Smith, Jr.W.O., and Demaster, D.J. (1996). Phytoplankton biomass and productivity in the Amazon River plume: correlation with seasonal river discharge. Continental Shelf Research, 16 (3): 291-319.Google Scholar
Sohm, J.A., Webb, E.A., and Capone, D.G. (2011). Emerging patterns of marine nitrogen fixation. Nature Reviews: Microbiology, 9: 499-508.Google Scholar
Son, N-T., and Chen, C-F. (2013). Remote sensing of mangrove forests in Central America. SPIE Newsroom, doi: 10.1117/2.1201304.004771 (http://spie.org/x93599.xml)
Stal, L.J. (2009). Is the distribution of nitrogen-fixing cyanobacteria in the oceans related to temperature? Environmental Microbiology, 11 (7): 1632-45.Google Scholar
Steele, J.H. (1985). A comparison of terrestrial and marine ecological systems. Nature, 313, 355-358.Google Scholar
Steeman-Nielsen, E. (1963). Productivity, definition and measurement. In Hill, M.W. ed. The Sea, Vol. 1, pp 129-164. New York, John Wiley.
Steinacher, M., Joos, F., Frölicher, T.L., Bopp, L., Cadule, P., Cocco, V., Doney, S.C., Gehlen, M., Lindsay, K., Moore, J.K., Schneider, B., and Segschneider, J. (2010). Projected 21st century decrease in marine productivity: a multi-model analysis, Biogeosciences, 7: 979-1005.Google Scholar
Steinberg, D.K., Carlson, C.A., Bates, N.R., Johnson, R.J., Michaels, A.F., and Knap, A.H. (2001). Overview of the US JGOFS Bermuda Atlantic Time-series Study (BATS): a decade-scale look at ocean biology and biogeochemistry. Deep-Sea Research II, 48: 1405-1447.Google Scholar
Steinberg, D.K., Lomas, M.W., Cope, J.S., (2012). Long-term increase in mesozooplankton biomass in the Sargasso Sea: Linkage to climate and implications for food web dynamics and biogeochemical cycling. Global Biogeochemical Cycles 26 (1), GB1004.Google Scholar
Stier, P., Feichter, J., Kinne, S., Kloster, S., Vignati, E., Wilson, J., Ganzeveld, L., Tegen, I., Werner, M., Balkanski, Y., Schulz, M., Boucher, O., Minikin, A., and Petzold, A. (2005). The aerosol-climate model ECHAM5-HAM. Atmospheric Chemistry and Physics, 5: 1125–1156.Google Scholar
Strous, M., Pelletier, E., Mangenot, S., et al. (2006). Deciphering the evolution and metabolism of an anammox bacterium from a community genome. Nature, 440, 790–794.Google Scholar
Subramaniam, A., Yager, P.L., Carpenter, E.J., Mahaffey, C., Björkman, K., Cooley, S., Kustka, A.B., Montoya, J.P., Sañudo-Wilhelmy, S.A., Shipe, R., and Capone, D.G. (2008). Amazon River enhances diazotrophy and carbon sequestration in the tropical North Atlantic Ocean. PNAS, 105 (3): 10460-10465.Google Scholar
Suntharalingam, P., Buitenhuis, E., Le Quéré, C., Dentener, F., Nevison, C., Butler, J.H., Bange, H.W., and Forster, G. (2012). Quantifying the impact of anthropogenic nitrogen deposition on oceanic nitrous oxide, Geophysical Research Letters, 39: L07605, doi:10.1029/2011GL050778.Google Scholar
Suikkanen, S., Pulina, S., Engström-Öst, J., Lehtiniemi, M., Lehtinen, S., et al. (2013) Climate Change and Eutrophication Induced Shifts in Northern Summer Plankton Communities. PLoS ONE 8 (6): e66475. doi:10.1371/journal.pone.0066475.Google Scholar
Taalas, P., Kaurola, J., Kylling, A., Shindell, D., Sausen, R., Dameris, M., Grewe, V., Herman, J., Damski, J. and Steil, B. (2000). The impact of greenhouse gases and halogenated species on future solar UV radiation doses. Geophysical Research Letters, 27: 1127-1130.Google Scholar
Taucher, J. and Oschlie., A. (2011). Can we predict the direction of marine primary production change under global warming? Geophysical Research Letters, 38: L02603.Google Scholar
Taucher, J., Schulz, K.G., Dittmar, T., Sommer, U., Oschlies, A., and Riebesell, U. (2012) Enhanced carbon overconsumption in response to increasing temperatures during a mesocosm experiment. Biogeosciences, 9: 3479–3514.Google Scholar
Teira, E., Mouriño, B., Marañón, E., Pérez, V., Pazó, M.J., Serret, P., de Armas, D., Escánez, J., Woodward, E.M.S., and Fernández, E. (2005). Variability of chlorophyll and primary production in the eastern north Atlantic subtropical gyre: potential factors affecting phytoplankton activity. Deep-Sea Research, 52: 569–588.Google Scholar
Testor, P., Meyers, G., Pattiaratchi, C., Bachmayer, R., Hayes, D., Pouliquen, S., et al. (2010). Gliders as a component of future observing systems in Proceedings of OceanObs–09: Sustained Ocean Observations and Information for Society (Vol. 2), Hall, J., Harrison, D.E., and Stammer, D., eds. ESA Publication WPP-306, doi:10.5270/OceanObs09.
Teuten, E.L., Saquing, J.M., Knappe, D.R.U., Barlaz, M.A., Jonsson, S., Bjorn, A., Rowland, S.J., Thompson, R.C., Galloway, T.S., Yamashita, R., Ochi, D., Watanuki, Y., Moore, C., Viet, P.H., Tana, T.S., Prudente, M., Boonyatumanond, R., Zakaria, M.P., Akkhavong, K., Ogata, Y., Hirai, H., Iwasa, S., Mizukawa, K., Hagino, Y., Imamura, A., Saha, M., Takada, H. (2009). Transport and release of chemicals from plastics to the environment and to wildlife. Philosophical Transactions of the Royal Society of London B: Biological Science, 364: 2027-2045.Google Scholar
Thamdrup, B., Dalsgaard, T., Jensen, M.M., et al. (2006). Anaerobic ammonium oxidation in the oxygen-deficient waters off northern Chile. Limnology and Oceanography, 51: 2145–2156.Google Scholar
Tian, T., Wei, H., Su, J., and Chung, C. (2005). Simulations of annual cycle of phytoplankton production and the utilization of nitrogen in the Yellow Sea. Journal of Oceanography, 61: 343-357.Google Scholar
Titus, J.G., Anderson, K.E., Cahoon, D.R., Gesch, D.B., Gill, S.K, Gutierrez, B.T., Thieler, E.R., and Williams, S.J. (2009). Coastal sensitivity to sea level rise: A Focus on the Mid-Atlantic Region. U.S. Climate Change Science Program, Synthesis and Assessment Product 4.1.
Tremblay, J.E., and Gagnon, J. (2009). The effect of irradiance and nutrient supply on the productivity of Arctic waters: A perspective on climate change, pp. 73–94, In Influence of Climate Change on the Changing Arctic and Sub-Arctic Conditions, Springer, New York.
Trilla, G.G., Pratolongo, P., Beget, M.E., Kandus, P., Marcovecchio, J., and Di Bella, C. (2013). Relating biophysical parameters of coastal marshes to hyperspectral reflectance data in the Bahia Blanca Estuary, Argentina. Journal of Coastal Research, 29 (1): 231-238.Google Scholar
Twilley, R.R., Chen, R.H., and Hargis, T. (1992). Carbon sinks in mangroves and their implications to carbon budget of tropical coastal ecosystems. Water Air and Soil Pollution, 64 (1): 265-265.Google Scholar
Ueyama, R., and Monger, B.C. 2005. Wind-induced modulation of seasonal phytoplankton blooms in the North Atlantic derived from satellite observations, Limnology and Oceanography, 50: 1820– 1829.Google Scholar
Ulloa, O., Canfieldb, D.E., DeLong, E.F., Letelier, R.M., and Stewart, F.J. (2012). Microbial oceanography of anoxic oxygen minimum zones. Proceedings of the National Academy of Sciences (USA), 109 (40): 15996-16003.Google Scholar
UNEP-WCMC. (2006). In the front line: Shoreline protection and other ecosystem services from mangroves and coral reefs. UNEP-WCMC, Cambridge, UK 33 pp.
UNESCO. (2006). A Coastal Theme for the IGOS Partnership for the Monitoring of our Environment from Space and from Earth. Paris, 60 pp.
UNESCO. (2008). Filling Gaps in LME Nitrogen Loadings Forecast for 64 LMEs. GEF/ LME
UNESCO-IOC. (2012). Requirements for Global Implementation of the Strategic Plan for Coastal GOOS. GOOS Report 193. Paris: Intergovernmental Oceanographic Commission. (http://www.ioc-goos.org/index.php?option=com_oe&task=viewDocum entRecord&docID=7702&lang=en)
Uye, S. (2008). Blooms of the giant jellyfish Nemopilema nomurai: A threat to the fisheries sustainability of the East Asian Marginal Seas. Plankton & Benthos Researc., 3: 125–131.Google Scholar
Valiela, I., Bowen, J.L., and York, J.K. (2001). Mangrove forests: One of the world's threatened major tropical environments. BioScience, 51 (10): 807–815.Google Scholar
Valiela, I., Kinney, E., Culbertson, J., Peacock, E., and Smith, S. (2009). Global losses of mangroves and salt marshes: Magnitudes, causes and consequences. Pp. 107-138 in Global Loss of Coastal Habitats: Rates, Causes, and Consequences, Duarte, C. ed. Fundación BBVA. Bilbao.
Vancoppenolle, M., Bopp, L., Madec, G., Dunne, J., Ilyina, T., Halloran, P.R., and Steiner, N. (2013), Future Arctic Ocean primary productivity from CMIP5 simulations: Uncertain outcome, but consistent mechanisms, Global Biogeochemical Cycles, 27, doi:10.1002/gbc.20055.Google Scholar
Van Dolah, F.M. (2000). Marine algal toxins: origins, health effects, and their increased occurrence. Environmental Health Perspectives, 108 (suppl 1): 133-141.Google Scholar
Vantrepotte, V., and Mélin, F. (2009). Temporal variability of 10-year global SeaWiFS time-series of phytoplankton chlorophyll a concentration. ICES Journal of Marine Science, 66: 1547–1556.Google Scholar
Vargas, M., Brown, C.W., and Sapiano, M.R.P. (2009). Phenology of marine phytoplankton from satellite ocean color measurements. Geophysical Research Letters, 36: L01608, doi: 10.1029/2008GL036006.Google Scholar
Vernet, M., and Smith, R.C. (2007) Measuring and Modeling Primary Production in Marine Pelagic Ecosystems, in Principles and Standards for measuring Primary Production, Fahey, J.T., and Knapp, A.K., eds. Oxford University Press, Ch.9, pp.142-174.
Vernet, M., Martinson, D., Iannuzzi, R., Stammerjohn, S., Kozlowski, W., Sines, K., Smith, R.C., and Garibotti, I. (2008). Primary production within the sea-ice zone west of the Antarctic Peninsula: I – Sea ice, summer mixed layer, and irradiance. Deep Sea Research Part II: Topical Studies in Oceanography, 55: 2068-2085.Google Scholar
Villar-Argaiz, M., Median-Sánchez, J.M., Bullejos, F.J., Delgado-Molina, J.A., Ruíz Pérez, O., Navarro, J.C., and Carrillo, P. (2009). UV radiation and phosphorus interact to influence the biochemical composition of phytoplankton. Freshwater Biology, 54, 1233-1245.Google Scholar
Vonthron-Sénécheau, C., Kaiser, M., Devambez, L., Vastel, A., Mussio, I., Rusig, A.M. (2011). Antiprotozoal activities of organic extracts from French marine seaweeds. Marine Drugs, 9 (6):922-33.Google Scholar
Voss, M., Bange, H.W., Dippner, J.W., Middelburg, J.J., Montoya, J.P., and Ward, B.B. (2013). The marine nitrogen cycle: recent discoveries, uncertainties and the potential relevance of climate change. Philosophical Transactions of the Royal Society B, Biological Sciences, 368 (1621): 20130121, doi: 10.1098/rstb.2013.0121.Google Scholar
Walker, B., Holling, C.S., Carpenter, S.R., Kinzig, A. (2004). “Resilience, adaptability and transformability in social–ecological systems”. Ecology and Society, 9 (2): 5.Google Scholar
Walsh, P.J., Smith, S.L., Fleming, L.E., Solo-Gabriele, H.M., and Gerwick, W.H., eds. (2008). Oceans and Human Health: Risk and Remedies from the Sea. Elsevier, 644 p.
Wang, M., and Overland, J. (2009). An sea ice free summer Arctic within 30 years. Geophysical Research Letters, 36, L07502. http://dx.doi.org/10.1029/2009GL037820.Google Scholar
Ward, B.A., Dutkiewicz, S., Jahn, O., and Follows, M.J. (2012). A size-structured foodweb model for the global ocean. Limnology and Oceanography, 57(6): 1877– 1891.Google Scholar
Ward, B.B. (2012). The Global Nitrogen Cycle. In: Knol., A.H., Canfield, D.E., and Konhauser, K.O., eds, Fundamentals of Geomicrobiology, Wiley-Blackwell, Chichester, UK, pp. 36-48.
Ward, B.B. (2013). How is nitrogen lost? Science, 341: 352 – 352.Google Scholar
Ward, B.B., Capone, D.G., and Zehr, J.P. (2007). What's New in the Nitrogen Cycle? Oceanography, 20 (2): 101-109.Google Scholar
Ward, B.B., Devol, A.H., Rich, J.J., Chang, B.X., Bulow, S.E., Naik, H., Pratihary, A., and Jayakumar, A. (2009). Denitrification as the dominant nitrogen loss process in the Arabian Sea, Nature, 461:78–81.Google Scholar
Ware, D.M. (2000). Aquatic ecosystems: Properties and models. In: Harrison, P.J., Parsons, T.R., eds. Fisheries oceanography: An integrative approach to fisheries ecology and management. Oxford: Balckwell Science. pp. 161–200.
Ware, D.M. and Thomso., R.E. (2005). Bottom-up ecosystem trophic dynamics determine fish production in the northeast Pacific. Science, 308: 1280–1284.Google Scholar
Waycott, M., Duarte, C.M., Carruthers, T.J.B., Orth, R.J., Dennison, W.C., Olyarnik, S., Calladine, A., Fourqurean, J.W., Heck, K.L., Hughes, A.R., Kendrick, G.A., Kenworthy, W.J., Short, F.T., and Williams, S.L. (2009). Accelerating loss of seagrasses across the globe threatens coastal ecosystems. Proceedings of the National Academy of Sciences USA 106:12377–12381.
Wegner, I.A., Besseling, E., Foekema, E.M., Kamermans, P., and Koelmans, A.A. (2012). Effects of nanopolystyrene on the feeding behavior of the blue mussel (Mytilus edulis L.). Environmental Toxicology and Chemistry, 31 (11): 2490-2497.Google Scholar
Wernand, M.R., van der Woerd, H.J., Gieskes, W.W.C. (2013). Trends in ocean colour and chlorophyll concentration from 1889 to 2000, worldwide. PLoS One, 8(6).Google Scholar
Wernberg, T., Russell, B.D., Thomsen, M.S., Gurgel, C.F.D., Bradshaw, C.J.A., Poloczanska, E.S., Connell, S.D. (2011). Seaweed communities in retreat from ocean warming. Current Biology, 21: 1828-1832.Google Scholar
Westberry, T., Behrenfeld, M.J., Siegel, D.A., and Boss, E. (2008). Carbon based primary productivity modeling with vertically resolved photoacclimation. Global Biogeochemical Cycles, 22, GB2024, doi:10.1029/2007GB003078.Google Scholar
Wollast, R. (1998). Evaluation and comparison of the global carbon cycle in the coastal zone and the open ocean. In The Sea, Brink, K.H. and A.R., Robinson, eds. 10: 213-253.Google Scholar
Worm, B., Barbier, E.B., Beaumont, N., Duffy, J.E., Folke, C., Halpern, B.S., Jackson, J.B.C., Lotze, H.K., Micheli, F., and Palumbi, S.R. (2006). Impacts of biodiversity loss on ocean ecosystem services. Science, 314: 787-790.Google Scholar
Wright, S.L., Thompson, R.C., and Galloway, T.S. (2013). The physical impacts of microplastics on marine organisms: A review. Environmental Pollution, 178, 483e492.Google Scholar
Yusuf, C. (2007). Biodiesel from microalgae. Biotechnology Advances, 25:294-306.Google Scholar
Zehr, J.P., Kudela, R.M. (2011), Nitrogen Cycle of the Open Ocean: From Genes to Ecosystems. Annual Reviews Marine Science, 3: 197–225.Google Scholar
Zhang, M., Ustin, S.L., Rejmankova, E., and Sanderson, E.W. (1997). Monitoring Pacific coast salt marshes using remote sensing. Ecological Applications, 7 (3): 1039–1053.Google Scholar
Zhai, L., Platt, T., Tang, C., Sathyendranath, S., and Walne, A. (2013). The response of phytoplankton to climate variability associated with the North Atlantic Oscillation. Oceanography, 93: 159-168.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

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
×