Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-9q27g Total loading time: 0 Render date: 2024-07-21T07:42:10.987Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 August 2014

Catriona L. Hurd
Affiliation:
University of Tasmania
Paul J. Harrison
Affiliation:
University of British Columbia, Vancouver
Kai Bischof
Affiliation:
Universität Bremen
Christopher S. Lobban
Affiliation:
University of Guam
Get access
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2014

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

Abdullah, M.I., and Fredriksen, S. (2004). Production, respiration and exudation of dissolved organic matter by the kelp Laminaria hyperborea along the west coast of Norway. J. Mar. Biol. Assoc. UK 84:887–94.CrossRefGoogle Scholar
Åberg, P. (1992). A demographic study of two populations of the seaweed Ascophyllum nodosum. Ecology 73:1473–87.CrossRefGoogle Scholar
Åberg, P. (1996). Patterns of reproductive effort in the brown alga Ascophyllum nodosum. Mar. Ecol. Prog. Ser. 138:199–207.CrossRefGoogle Scholar
Abreu, M. H., Pereira, R., Sousa-Pinto, I., and Yarish, C. (2011). Ecophysiological studies of the non-indigenous species Gracilaria vermiculophylla (Rhodophyta) and its abundance patterns in Ria de Aveiro lagoon, Portugal. Eur. J. Phycol. 46:453–64.CrossRefGoogle Scholar
Ackerman, J. D. (1998). Is the limited diversity of higher plants in marine systems the result of biophysical limitations for reproduction or evolutionary and physiological constraints? Funct. Ecol. 12:979–82.Google Scholar
Ackland, J. C., West, J. A., and Pickett-Heaps, J. (2007). Actin and myosin regulate pseudopodia of Porphyra pulchella (Rhodophyta) archeospores. J. Phycol. 43:129–38.CrossRefGoogle Scholar
Adams, M. S., Stauber, J. L., Binet, M. T., Molloy, R., and Gregory, D. (2008). Toxicity of a secondary-treated sewage effluent to marine biota in Bass Strait, Australia: Development of action trigger values for a toxicity monitoring program. Mar. Poll. Bull. 57:587–98.CrossRefGoogle ScholarPubMed
Adams, N. M. (1994). Seaweeds of New Zealand: An Illustrated Guide. Christchurch, NZ: Canterbury University Press, 360 pp.Google Scholar
Adams, T. S., and Sterner, R. W. (2000). The effect of dietary nitrogen content on trophic level 15N enrichment. Limnology and Oceanography 45(3):601–7CrossRefGoogle Scholar
Adey, W. H., and Goertemiller, T. (1987). Coral reef algal turfs: master producers in nutrient poor seas. Phycologia 26:374–86.CrossRefGoogle Scholar
Adey, W. H., and Vassar, J. M. (1975). Colonization, succession and growth rates of tropical crustose coralline algae (Rhodophyta, Cryptonemiales). Phycologia 14:55–69.CrossRefGoogle Scholar
Aguilera, J., Dummermuth, A., Kartsen, U., Schrieck, R., and Wiencke, C. (2002). Enzymatic defences against photooxidative stress induced by ultraviolet radiation in Arctic marine macroalgae. Polar Biol. 25:432–41.Google Scholar
Aguirre-Lipperheide, M., Estrada-Rodriguez, F. J., and Evans, L.V. (1995). Facts, problems and need in seaweed tissue culture: an appraisal. J. Phycol. 31:677–88.CrossRefGoogle Scholar
Ahlf, W., Hollert, H., Neumann-Hensel, H., and Ricking, M. (2002). A guidance for the assessment and evaluation of sediment quality a German approach based on ecotoxicological and chemical measurements. J. Soils Sediments 2:37–42.CrossRefGoogle Scholar
Ahn, O., Petrell, R. J., and Harrison, P. J. (1998). Ammonium and nitrate uptake by Laminaria saccharina and Nereocystis leutkeana originating from a salmon sea cage farm. J. Appl. Phycol. 10:333–40.CrossRefGoogle Scholar
Airoldi, L. (1998). Roles of disturbance, sediment stress, and substratum retention on spatial dominance in algal turf. Ecology 79:2759–70.CrossRefGoogle Scholar
Airoldi, L. (2000). Effects of disturbance, life histories, and overgrowth on coexistence of algal crusts and turfs. Ecology 81:798–814.CrossRefGoogle Scholar
Airoldi, L. (2003). The effects of sedimentation on rocky coast assemblages. Oceanogr. Mar. Biol. 41:161–236.Google Scholar
Airoldi, L., and Virgilio, M. (1998). Responses of turf-forming algae to spatial variations in the deposition of sediments. Mar. Ecol. Prog. Ser. 165:271–82.CrossRefGoogle Scholar
Alamsjah, M. A., Hirao, S., Ishibashi, F., et al. (2008). Algicidal activity of polyunsaturated fatty acids derived from Ulva fasciata and U. pertusa (Ulvaceae, Chlorophyta) on phytoplankton. J. Appl. Phycol. 20:713–20.CrossRefGoogle Scholar
Alfaro, A. C., Thomas, F., Sergent, L., and Duxbury, M. (2006). Identification of trophic interactions within an estuarine food web (northern New Zealand) using fatty acid biomarkers and stable isotopes. Estuar. Coast. Shelf Sci. 70(1–2):271–86.CrossRefGoogle Scholar
Algarra, P., and Neill, F. X. (1987). Structural adaptations to light reception in two morphotypes of Corallina elongata EIlis and Soland. PSZNIMar. Ecol. 8:253–61.CrossRefGoogle Scholar
Algarra, P., Thomas, J. C., and Mousseau, A. (1990). Phycobilisome heterogeneity in the red alga Porphyra umbilicalis. Plant Physiol. 92:570–6.CrossRefGoogle ScholarPubMed
Allen, J. F. (2003). State-transitions: a question of balance. Science 299:1530–2.CrossRefGoogle Scholar
Allen, T. F. H. (1977). Scale in microscopic algal ecology: a neglected dimension. Phycologia 16:253–7.CrossRefGoogle Scholar
Almeida, E., Diamantino, T. C., and de Sousa, O. (2007). Marine paints: the particular case of antifouling paints. Prog. Org. Coatings 59:2–20.CrossRefGoogle Scholar
Amat, M. A., and Braud, J.-P. (1990). Ammonium uptake by Chondrus crispus Stackhouse (Gigartinales; Rhodophyta) in culture. Hydrobiologia 204/205:467–71.CrossRefGoogle Scholar
Amsler, C. D. (ed.). (2008). Algal Chemical Ecology. London: Springer, 313 pp.CrossRefGoogle Scholar
Amsler, C. D. (2012). Chemical ecology of seaweeds. In Wiencke, C., and Bischof, K. (eds), Seaweed Biology: Novel Insights Into Ecophysiology, Ecology and Utilization (pp. 177–88). Series: Ecological Studies, Volume 219. Berlin and Heidelberg: Springer.CrossRefGoogle Scholar
Amsler, C. D., and Fairhead, V. A. (2006). Defensive and sensory chemical ecology of brown algae. Adv. Bot. Res. 43:1–91. [, , ]Google Scholar
Amsler, C. D., and Neushul, M. (l989a). Diel periodicity of spore release from the kelp Nereocystis luetkeana (Mertens) Postels et Ruprecht. J. Exp. Mar. Biol. Ecol. 134:117–27.CrossRefGoogle Scholar
Amsler, C. D., and Neushul, M. (1989b). Chemotactic effects of nutrients on spores of the kelps Macrocystis pyrifera and Pterygophora californica. Mar. Biol. 102:557–64.CrossRefGoogle Scholar
Amsler, C. D., and Neushul, M. (1990). Nutrient stimulation of spore settlement in the kelps Pterygophora californica and Macrocystis pyrifera. Mar. Biol. 107:297–304.CrossRefGoogle Scholar
Amsler, C. D., and Searles, R. B. (1980). Vertical distribution of seaweed spores in a water column offshore of North Carolina. J. Phycol. 16:617–19.CrossRefGoogle Scholar
Amsler, C. D., Reed, D. C., and Neushul, M. (1992). The microclimate inhabited by macroalgal propagules. Brit. Phycol. J. 27:253–70.CrossRefGoogle Scholar
Amsler, C. D., McClintock, J. B., and Baker, B. J. (2008). Macroalgal chemical defenses in polar marine communities. In Amsler, C. D. (ed.), Algal Chemical Ecology (pp. 91–104). London: Springer.CrossRefGoogle Scholar
Amsler, C. D., Amsler, M. O., McClintock, J. B., and Baker, B. J. (2009). Filamentous algal endophytes in macrophytic Antarctic algae: prevalence in hosts and palatability to mesoherbivores. Phycologia 48:324–34.CrossRefGoogle Scholar
Anderberg, H. I., Danielson, J., and Johanson, U. (2011). Algal MIPs, high diversity and conserved motifs. BMC Evol. Biol. 11:110.CrossRefGoogle ScholarPubMed
Andersen, R. A. (2004). Biology and systematics of heterokont and haptophyte algae. Am. J. Bot. 91:1508–22.CrossRefGoogle ScholarPubMed
Anderson, B. S., and Hunt, J. W. (1988). Bioassay methods of evaluating the toxicity of heavy metals, biocides and sewage effluent using microscopic stages of giant kelp Macrocystis pyrifera (Agardh): a preliminary report. Mar. Environ. Res. 26:113–34.CrossRefGoogle Scholar
Anderson, B. S., Hunt, J. W., Turpen, S. L., Conlon, A. R., and Martin, M. (1990). Copper toxicity to microscopic stages of giant kelp Macrocystis pyrifera: interpopulation comparisons and temporal variability. Mar. Ecol. Prog. Ser. 68:147–56.CrossRefGoogle Scholar
Anderson, F. E., and Green, J. (1980). Estuaries. The New Encyclopaedia Britannica. Macropaedia vol. 6, (pp. 968–76). London: Encyclopaedia Britannica.Google Scholar
Anderson, R. J., and Bolton, J. J. (1989). Growth and fertility, in relation to temperature and photoperiod, in South African Desmarestia firma (Phaeophyceae). Bot. Mar. 32:149–58.Google Scholar
Anderson, R. J., Anderson, D. R., and Anderson, J. S. (2008). Survival of sand-burial by seaweeds with crustose bases or life-history stages structures the biotic community on an intertidal rocky shore. Bot. Mar. 51:10–20.CrossRefGoogle Scholar
Anderson, S. M., and Charters, A. C. (1982). A fluid dynamics study of seawater flow through Gelidium nudifrons. Limnol. Oceanogr. 27:399–412.CrossRefGoogle Scholar
Andersson, S., and Kautsky, L. (1996). Copper effects on reproductive stages of Baltic Sea Fucus vesiculosus. Mar. Biol. 125:171–6.CrossRefGoogle Scholar
Andrade, L.R., Farina, M., and Amado Filho, G. M. (2004). Effects of copper on Enteromorpha flexuosa (Chlorophyta) in vitro. Ecotox. Envir. Safety 58:117–25.CrossRefGoogle ScholarPubMed
Andrade, S., Medina, M. H., Moffet, J. W., and Correa, J. A. (2006). Cadmium-copper antagonism in seaweeds inhabiting coastal areas affected by copper mine waste disposal. Environ. Sci. Technol. 40:4382–7.CrossRefGoogle Scholar
Andrade, S., Pulido, M. J., and Correa, J. A. (2010). The effect of organic ligands exuded by intertidal seaweeds on copper complexation. Chemosphere 78:397–401. [, , ]CrossRefGoogle ScholarPubMed
Andreakis, N., and Schaffelke, B. (2012). Invasive marine seaweeds: pest or prize? In Wiencke, C., and Bischof, K. (eds), Seaweed Biology: Novel Insights into Ecophysiology, Ecology and Utilization, Ecological Studies 219 (pp. 235–62). Berlin and Heidelberg: Springer-Verlag.CrossRefGoogle Scholar
Andréfouët, S., Payri, C., Hochberg, E. J., et al. (2004). Use of in situ and airborne reflectance for scaling-up spectral discrimination of coral reef macroalgae from species to communities. Mar. Ecol. Prog. Ser. 283:161–77.CrossRefGoogle Scholar
Andrew, N. L. (1993). Spatial heterogeneity, sea urchin grazing, and habitat structure on reefs in temperate Australia. Ecology 74:292–302.CrossRefGoogle Scholar
Andrews, J. H. (1976). The pathology of marine algae. Biol. Rev. 51:211–53.CrossRefGoogle Scholar
Ang, P. O. (1992). Cost of reproduction in Fucus distichus. Mar. Ecol. Prog. Ser. 89:25–35.CrossRefGoogle Scholar
Ang, P. O., and DeWreede, R. E. (1993). Simulation and analysis of the dynamics of a Fucus distichus (Phaeophyceae, Fucales) population. Mar. Ecol. Prog. Ser. 93:253–65.CrossRefGoogle Scholar
Anthony, K. R. N., Maynard, J. A., Diaz-Pulido, G., et al. (2011). Ocean acidification and warming will lower coral reef resilience. Global Change Biology 17:1798–808.CrossRefGoogle Scholar
Antia, N. J., Harrison, P. J., and Oliveira, L. (1991). The role of dissolved organic nitrogen in phytoplankton nutrition, cell biology and ecology. Phycologia 30:1–89.CrossRefGoogle Scholar
Antizar-Ladislao, B. (2008). Environmental levels, toxicity and human exposure to tributyltin (TBT) contaminated marine environment. A review. Environ. Internat. 34:292–308.CrossRefGoogle ScholarPubMed
Ar Gall, E., Küpper, F. C., and Kloareg, B. (2004). A survey of iodine content in Laminaria digitata. Bot. Mar. 47:30–7.Google Scholar
Archibald, J. M. (2009). The origin and spread of eukaryotic photosynthesis: evolving views in light of genomics. Bot. Mar. 52:95–103.CrossRefGoogle Scholar
Arenas, F., and Fernández, C. (2000). Size structure and dynamics in a population of Sargassum muticum (Phaeophyceae). J. Phycol. 36:1012–20.CrossRefGoogle Scholar
Arenas, F., Sánchez, I., Hawkins, S. J., and Jenkins, S. R. (2006). The invisibility of marine algal assemblages: role of functional diversity and identity. Ecology 87:2851–61.CrossRefGoogle Scholar
Arkema, K. K., Reed, D. C., and Schroeter, S. C. (2009). Direct and indirect effects of giant kelp determine benthic community structure and dynamics. Ecology 90:3126–37.CrossRefGoogle ScholarPubMed
Armbrust, E., Berges, J., Bowler, C., et al. (2004). The genome of the diatom Thalassiosira Pseudonana: Ecology, evolution, and metabolism. Science 306: 76–86.CrossRefGoogle ScholarPubMed
Armstrong, S. L. (1987). Mechanical properties of the tissues of the brown alga Hedophyllum sessile (C. Ag.) Setchell: variability with habitat. J. Exp. Mar. Biol. Ecol. 114:143–51.CrossRefGoogle Scholar
Arnold, K. E., and Manley, S. L. (1985). Carbon allocation in Macrocystis pyrifera (Phaeophyta): intrinsic variability in photosynthesis and respiration. J. Phycol. 21:154–67.CrossRefGoogle Scholar
Arnold, K. E., and Murray, S. N (1980). Relationships between irradiance and photosynthesis for marine benthic green algae (Chlorophyta) of differing morphologies. J. Exp. Mar. Biol. Ecol. 43: 183–92.CrossRefGoogle Scholar
Arnold, T. M., Targett, N. M., Tanner, C. E., Hatch, W. I., and Ferrari, K. E. (2001). Evidence for methyl jasmonate-induced phlorotannin production in Fucus vesiculosus (Phaeophyceae). J. Phycol. 37:1026–9.CrossRefGoogle Scholar
Arntz, W. E., Gallardo, V. A., Gutiérrez, D., et al. (2006). El Niño and similar perturbation effects on the benthos of the Humboldt, California, and Benguela Current upwelling ecosystems. Adv. Geosci. 6:243–65.CrossRefGoogle Scholar
Aro, E. M., Virgin, I., and Andersson, B. (1993). Photoinhibition of photosystem II. Inactivation, protein damage and turnover. Biochim. Biophys. Acta 1143:113–34.CrossRefGoogle ScholarPubMed
Aronson, R. B., Precht, W. F., Macintyre, I. G., and Toth, L. T. (2012). Catastrophe and the life span of coral reefs. Ecology 93:303–13.CrossRefGoogle ScholarPubMed
Asada, K., and Takahashi, M. (1987). Production and scavenging of active oxygen in photosynthesis. In Kyle, D. J., Osmond, C. B., and Arntzen, C. J. (eds) Photoinhibition. Topics in Photosynthesis 9 (pp. 89–109). Amsterdam: Elsevier Science Publishers.Google Scholar
Ashen, J. B., Cohen, J. D., and Goff, L. J. (1999). GC-SIM-MS detection and quantification of free indole-3-acetic acid in bacterial galls on the marine alga Prionitis lanceolata (Rhodophyta). J. Phycol. 35:493–500.CrossRefGoogle Scholar
Asimgil, H., and Kavakli, I. H. (2012). Purification and characterization of five members of photolyase/cryptochrome family from Cyanidioschyzon merolae. Plant Sci. 185:190–8.CrossRefGoogle ScholarPubMed
Ask, E. I., and Azanza, R. V. (2002). Advances in cultivation technology of commercial eucheumatoid species: a review with suggestions for future research. Aquaculture 206:257–77.CrossRefGoogle Scholar
ASTM (2004). Standard Guide for Conducting Sexual Reproduction Tests with Seaweeds. ASTME1498–92.Google Scholar
Atalah, J., and Crowe, T. P. (2010). Combined effects of nutrient enrichment, sedimentation and grazer loss on rock pool assemblages. J. Exp. Mar. Biol. Ecol. 388:51–7.CrossRefGoogle Scholar
Ateweberhan, M., Bruggemann, J. H., and Breeman, A. M. (2005). Seasonal dynamics of Sargassum ilicifolium (Phaeophyta) on a shallow reef flat in the southern Red Sea (Eritrea). Mar. Ecol. Prog. Ser. 292:159–71.CrossRefGoogle Scholar
Ateweberhan, M., Bruggemann, J. H., and Breeman, A. M. (2006). Seasonal module dynamics of Turbinaria Triquetra (Fucales, Phaeophyceae) in the southern Red Sea. J. Phycol. 42:990–1001.CrossRefGoogle Scholar
Ateweberhan, M., Bruggemann, J. H., and Breeman, A. M. (2009). Seasonal changes in size structure of Sargassum and Turbinaria populations (Phaeophyceae) on tropical reef flats in the southern Red Sea. J. Phycol. 45:69–80.CrossRefGoogle ScholarPubMed
Atkinson, M. J., and Grigg, R. W. (1984). Model of a coral reef ecosystem. 11. Gross and net benthic primary production at French Frigate Shoals, Hawaii. Coral Reefs 3:13–22.CrossRefGoogle Scholar
Atkinson, M. J., and Smith, S. V. (1983). C: N: P ratios of benthic marine plants. Limnol. Ocenaogr. 28:568–74.CrossRefGoogle Scholar
Audibert, L., Fauchon, M., Blanc, N., et al. (2010). Phenolic compounds in the brown seaweed Ascophyllum nodosum: distribution and radical-scavenging activities. Phytochem. Anal. 21:399–405.CrossRefGoogle ScholarPubMed
Austin, A. P., and Pringle, J. D. (1969). Periodicity of mitosis in red algae. Proc. Intl. Seaweed Symp. 6:41–52.Google Scholar
Ávila, E., and Carballo, J. L. (2006). Habitat selection by larvae of the symbiotic sponge Haliclona caerulea (Hechtel, 1965) (Demospongiae, Haplosclerida). Symbiosis 41:21–9.Google Scholar
Azanza-Corrales, R., and Dawes, C. J. (1989). Wound healing in cultured Eucheuma alvarezii var. tambalang Doty. Bot. Mar. 32:229–34.CrossRefGoogle Scholar
Azimzadeh, J., and Marshall, W. F. (2010). Building the centriole. Curr. Biol. 20:R816–25.CrossRefGoogle ScholarPubMed
Baardseth, E. (1970). A Square-Scanning, Two-Stage Sampling Method of Estimating Seaweed Quantities. Norwegian Institute of Seaweed Research, Report 33, 41 pp.
Babcock, R. C., Shears, N. T., Alcala, A. C., et al. (2010). Decadal trends in marine reserves reveal differential rates of change in direct and indirect effects. PNAS 107:18,256–61.CrossRefGoogle ScholarPubMed
Bäck, S., Collins, J. C., and Russell, G. (1992). Effects of salinity on growth of Baltic and Atlantic Fucus vesiculosus. Br. Phycol. J. 27:39–47.CrossRefGoogle Scholar
Bacon, L. C., and Vadas, R. L. (1991). A model for gamete release in Ascophyllum nodosum (Phaeophyta). J. Phycol. 27: 166–73.CrossRefGoogle Scholar
Badger, M. (2003). The roles of anhydrases in photosynthetic CO2 concentrating mechanisms. Photosynth. Res. 77:83–94.CrossRefGoogle Scholar
Badger, M. R., Andrews, T. J., Whitney, S. M., et al. (1998). The diversity and co-evolution of Rubisco, plastids, pyrenoids and chloroplast-based CO2 concentrating mechanisms in the algae. Can. J. Bot. 76:1052–71.Google Scholar
Baker, S. M., and Bohling, M. H. (1916). On the brown seaweeds of the salt marsh. Part II. Their systematic relationships, morphology, and ecology. J. Linn. Soc. Bot. 43:325–80.CrossRefGoogle Scholar
Balasse, M., Mainland, I., and Richards, M. (2009). Stable isotope evidence for seasonal consumption of marine seaweed by modern and archeological sheep in the Orkney archipelago (Scotland). Environ. Archeol. 14:1–14.CrossRefGoogle Scholar
Balata, D., Piazzi, L., and Benedetti-Cecchi, L. (2007). Sediment disturbance and loss of beta diversity on subtidal rocky reefs. Ecology 88:2455–61.CrossRefGoogle ScholarPubMed
Balata, D., Piazzi, L., and Rindi, F. (2011). Testing a new classification of morphological functional groups of marine macroalgae for the detection of responses to stress. Mar. Biol. 158:2459–69.CrossRefGoogle Scholar
Ballesteros, E., Garrabou, J., Hereu, B., et al. (2009). Deep-water stands of Cystoseira zosteroides C. Agardh (Fucales, Ochrophyta) in the Northwestern Mediterranean: insights into assemblage structure and population dynamics. Estuar. Coast. Shelf Sci. 82: 477–84.CrossRefGoogle Scholar
Barbrook, A. C., Howe, C. J., Kurniawan, D. P., and Tarr, S. J. (2010). Organization and expression of organellar genomes. Phil. Trans. R. Soc. B. 365:785–97.CrossRefGoogle ScholarPubMed
Barile, P. J., and Lapointe, B. E. (2005). Atmospheric nitrogen deposition from a remote source enriches macroalgae in coral reef ecosystems near Green Turtle Cay, Abacos, Bahamas. Mar. Poll. Bull. 50:1262–72.CrossRefGoogle ScholarPubMed
Barner, A. K., Pfister, C. A., and Wootton, T. (2010). The mixed mating system of the sea palm kelp Postelsia palmaeformis: few costs to selfing. P. Roy. Soc. B. Bio. 278:1347–55.CrossRefGoogle ScholarPubMed
Barnes, D. J., and Chalker, B. E. (1990). Calcification and photosynthesis in reef-building corals and algae. In Dubinsky, Z. (ed.), Ecosystems of the World 25 (pp. 109–31). Amsterdam: Elsevier.Google Scholar
Barnett, B. E., and Ashcroft, C. R. (1985). Heavy metals in Fucus vesiculosus in the Humber Estuary. Environ. Pollut. B-9:193–201.CrossRefGoogle Scholar
Barr, N. G., and Rees, T. A. V. (2003) Nitrogen status and metabolism in the green seaweed Enteromorpha intestinalis: an examination of three natural populations. Mar. Ecol. Prog. Ser. 249:133–44.CrossRefGoogle Scholar
Barr, N. G., Kloeppel, A., Rees, T. A. V., et al. (2008). Wave surge increases rates of growth and nutrient uptake in the green seaweed Ulva pertusa maintained at low bulk flow velocities. Aquat. Biol. 3:179–86.CrossRefGoogle Scholar
Barrington, K., Chopin, T., and Robinson, S. (2009). Integrated multi-trophic aquaculture (IMTA) in marine temperate waters. In Soto, D. (ed.) Integrated Mariculture: A Global Review. FAO Fisheries and Aquaculture Technical Paper No. 529:7–49.Google Scholar
Bartsch, I., Wiencke, C., Bischof, K., et al. (2008). The genus Laminaria sensu lato: recent insights and developments. Eur. J. Phycol. 43: 1–86. [, , , , , , , , , ]CrossRefGoogle Scholar
Basu, S., Sun, H. G., Brian, L., Quatrano, R. L., and Muday, G. K. (2002). Early embryo development in Fucus distichus is auxin sensitive. Plant Physiol. 130:292–302.CrossRefGoogle ScholarPubMed
Bates, C. R., Saunders, G. W., and Chopin, T. (2009). Historical versus contemporary measures of seaweed biodiversity in the Bay of Fundy. Botany 87:1066–76.CrossRefGoogle Scholar
Baumann, H. A., Morrison, L., and Stengel, D. B. (2009). Metal accumulation and toxicity measured by PAM chlorophyll fluorescence in seven species of marine macroalgae. Ecotox. Environ. Safety 72:1063–75.CrossRefGoogle ScholarPubMed
Baweja, P., Sahoo, D., Garcia-Jiménez, P., and Robaina, R. R. (2009). Seaweed tissue culture as applied to biotechnology: problems, achievements and prospects. Phycol. Res. 57:45–58.CrossRefGoogle Scholar
Bayer-Giraldi, M., Uhlig, C., John, U., Mock, T., and Valentin, K. (2010). Antifreeze proteins in polar sea ice diatoms: diversity and gene expression in the genusFragilariopsis. Environ Microbiol 12:1041–52.CrossRefGoogle ScholarPubMed
Bayne, B. L. (1989). The biological effects of marine pollutants. In Albaigés, J. (ed.), Marine Pollution (pp. 131–51). New York: Hemisphere.Google Scholar
Beck, C. B. (2010). An Introduction to Plant Structure and Development. Plant Anatomy for the the Twenty-First Century. Cambridge: Cambridge University Press, 464 pp.CrossRefGoogle Scholar
Becker, S., Walter, B., and Bischof, K. (2009). Freezing tolerance and photosynthetic performance of polar seaweeds at low temperatures. Bot. Mar. 52: 609–16. [, , ]CrossRefGoogle Scholar
Becker, S., Graeve, M., and Bischof, K. (2010). Photosynthesis and lipid composition of the Antarctic endemic rhodophyte Palmaria decipiens: effects of changing light and temperature levels. Polar Biol. 33:945–55. [, , ]CrossRefGoogle Scholar
Beer, S., and Axelsson, L. (2004). Limitations in the use of PAM fluorometry for measuring photosynthetic rates of macroalgae at high irradiances. Eur. J. Phycol. 39:1–7.CrossRefGoogle Scholar
Beer, S., and Eshel, A. (1983). Photosynthesis of Ulva sp. I. Effects of desiccation when exposed to air. J. Exp. Mar. Biol. Ecol. 70:91–7.Google Scholar
Bégin, C., and Scheibling, R. E. (2003). Growth and survival of the invasive green alga Codium fragile ssp. tomentosoides in tide pools on a rocky shore in Nova Scotia. Bot. Mar. 46:404–12.CrossRefGoogle Scholar
Bekheet, I. A., Kandil, K. M., and Shaban, N. Z. (1984). Studies on urease extracted fromUlva lactuca. Hydrobiologia 116/117:580–3.CrossRefGoogle Scholar
Bell, E. C. (1995). Environmental and morphological influences on thallus temperature and desiccation of the intertidal alga Mastocarpus papillatus Kützing. J. Exp. Mar. Biol. Ecol. 191:29–55.CrossRefGoogle Scholar
Bell, E. C. (1999). Applying flow tank measurements to the surf zone: predicting dislodgment of the Gigartinaceae. Phycol. Res. 47:159–66.CrossRefGoogle Scholar
Bell, E. C., and Denny, M. W. (1994). Quantifying wave exposure: a simple device for recording maximum velocity and results of its use at several field sites. J. Exp. Mar. Biol. Ecol. 181:9–29. [, , ]CrossRefGoogle Scholar
Bell, G. (1997). The evolution of the life cycle of brown seaweeds. Biol. J. Linn. Soc. 60:21–38.CrossRefGoogle Scholar
Bellgrove, A., Kihara, H., Iwata, A., Aoki, M. N., and Heraud, P. (2009). Fourier transform infrared microspectroscopy as a tool to identify macroalgal propagules. J. Phycol. 45:560–70.CrossRefGoogle ScholarPubMed
Belliveau, D. J., Garbary, D. J., and McLachlan, J. L. (1990). Effects of fluorescent brighteners on growth and morphology of the red alga Antithamnion kylinii. Stain Technol. 65:303–11.CrossRefGoogle ScholarPubMed
Bellwood, D. R., Hughes, T. P., Folke, C., and Nyström, M. (2004). Confronting the coral reef crisis. Nature 429:827–33.CrossRefGoogle ScholarPubMed
Bender, E. A., Case, T. J., and Gilpin, M. E. (1984). Perturbation experiments in community ecology: theory and practice. Ecology 65: 1–13.CrossRefGoogle Scholar
Benedetti-Cecchi, L. (2001). Variability in abundance of algae and invertebrates at different spatial scales on rocky sea shores. Mar. Ecol. Prog. Ser. 215:79–92. [, , ]CrossRefGoogle Scholar
Benedetti-Cecchi, L. (2005). Unanticipated impacts of spatial variance of biodiversity on plant productivity. Ecol. Lett. 8:791–9.CrossRefGoogle Scholar
Benedetti-Cecchi, L., Bertocci, I., Vaselli, S., and Maggi, E. (2006). Morphological plasticity and variable spatial patterns in different populations of the red alga Rissoella verrucosa. Mar. Ecol. Prog. Ser. 315:87–98.CrossRefGoogle Scholar
Benson, E. E., Rutter, J. C., and Cobb, A. H. (1983). Seasonal variation in frond morphology and chloroplast physiology of the intertidal alga Codium fragile (Suringar) Hariot. New Phytol. 95:569–80.CrossRefGoogle Scholar
Berges, J. A. (1997). Algal nitrate reductases. Eur. J. Phycol. 32:3–8. [, , ]CrossRefGoogle Scholar
Berges, J. A., and Harrison, P. J. (1995). Nitrate reductase activity quantitatively predicts the rate of nitrate incorporation under steady state light limitation: a revised assay and characterization of the enzyme in three species of marine phytoplankton. Limnol. Oceanogr. 40:82–93.CrossRefGoogle Scholar
Berges, J. A., and Mulholland, M. (2008). Enzymes and cellular N cycling. In Capone, D. G., Bronk, D. A., Mulholland, M. R., and Carpenter, E. J. (eds), Nitrogen in the Marine Environment (pp. 1361–420). New York: Academic Press.Google Scholar
Bergström, L., Berger, R., and Kautsky, L. (2003). Negative direct effects on nutrient enrichment on the establishment of Fucus vesiculosus in the Baltic Sea. Eur. J. Phycol. 38:41–6.CrossRefGoogle Scholar
Berndt, M–L., Callow, J. A., and Brawley, S. H. (2002). Gamete concentrations and timing and success of fertilization in a rocky shore seaweed. Mar. Ecol. Prog. Ser. 226:273–85.CrossRefGoogle Scholar
Bernstein, B. B., and Jung, N. (1979). Selective pressures and coevolution in a kelp canopy community in southern California. Ecol. Monogr. 49:335–55.CrossRefGoogle Scholar
Bertness, M. D., and Callaway, R. (1994). Positive interactions in communities. Trends Ecol. Evol. 9:191–3.CrossRefGoogle ScholarPubMed
Bertness, M. D., Leonard, G. H., Levine, J. M., Schmidt, P. R., and Ingraham, A. O. (1999). Testing the relative contribution of positive and negative interactions in rocky intertidal communities. Ecology 80:2711–26.CrossRefGoogle Scholar
Bertness, M. D., Gaines, S. D., and Hay, M. E. (eds) (2001). Marine Community Ecology. Sunderland, MA: Sinauer Associates, Inc., 550 pp. [, 4.0]Google Scholar
Bertness, M. D., Trussell, G. C., Ewanschuk, P. J., and Silliman, B. R. (2002). Do alternate stable community states exist in the Gulf of Maine rocky intertidal zone? Ecology 83:3434–48.CrossRefGoogle Scholar
Besada, V., Andrade, J. M., Schultze, F., and González, J. J. (2009). Heavy metals in edible seaweeds commercialized for human consumption. J. Mar. Systems 75:305–15.CrossRefGoogle Scholar
Bessho, K., and Iwasa, Y. (2009). Heteromorphic and isomorphic alternations of generations in macroalgae as adaptations to a seasonal environment. Evol. Ecol. Res. 11:691–711.Google Scholar
Bessho, K., and Iwasa, Y. (2010). Optimal seasonal schedules and the relative dominance of heteromorphic and isomorphic life cycles in macroalgae. J. Theor. Biol. 267:201–12.CrossRefGoogle ScholarPubMed
Beutlich, A., Borstelmann, B., Reddemann, R., Speckenbach, K., and Schnetter, R. (1990). Notes on the life histories of Boergesenia and Valonia (Siphonocladales, Chlorophyta). Hydrobiologia 204/205:425–34.CrossRefGoogle Scholar
Bhattacharya, D., and Druehl, L. D. (1988). Phylogenetic comparison of the small-subunit ribosomal DNA sequence of Costaria costata (Phaeophyta) with those of other algae, vascular plants and oomycetes. J. Phycol. 24:539–43.CrossRefGoogle Scholar
Bidwell, J. R., Wheeler, K. W., and Burridge, T. R. (1998). Toxicant effects on the zoospore stage of the marine macroalga Ecklonia radiata (Phaeophyta: Laminariales). Mar. Ecol. Prog. Ser. 163:259–65.CrossRefGoogle Scholar
Bidwell, R. G. S. (1979). Plant Physiology, 2nd edn. New York: Macmillan.Google Scholar
Bidwell, R. G. S., and McLachlan, J. (1985). Carbon nutrition of seaweeds: photosynthesis, photorespiration and respiration. J. Exp. Mar. Biol. Ecol. 86:15–46.CrossRefGoogle Scholar
Biebl, R. (1962). Seaweeds. In Lewin, R. A. (ed.), Physiology and Biochemistry of the Algae (pp. 799–815). New York: Academic Press.Google Scholar
Biebl, R. (1970). Vergleichende Untersuchungen zur Temperaturresistenz von Meeresalgen entlang der pazifischen Küste Nordamerikas. Protoplasma 69: 61–83.CrossRefGoogle Scholar
Biedka, R. F., Gosline, J. M., and DeWreede, R. E. (1987). Biomechanical analysis of wave-induced mortality in the marine algaPterygophora californica. Mar. Ecol. Prog. Ser. 36: 163–70.CrossRefGoogle Scholar
Biggins, J., and Bruce, D. (1989). Regulation of excitation energy transfer in organisms containing phycobilins. Photosynth. Res. 20: 1–34.CrossRefGoogle ScholarPubMed
Billard, E., Serrão, E., Pearson, G., Destombe, C., and Valero, M. (2010). Fucus vesiculosus and spiralis species complex: a nested model of local adaptation at the shore level. Mar. Ecol. Prog. Ser. 405:163–74.CrossRefGoogle Scholar
Billoud, B., Le Bail, A., and Charrier, B. (2008). A stochastic 1D nearest-neighbour automaton models early development of the brown alga Ectocarpus siliculosus. Funct. Plant Biol. 35:1014–24.CrossRefGoogle Scholar
Bingham, S., and Schiff, J. A. (1979). Conditions for attachment and development of single cells released from mechanically disrupted thalli of Prasiola stipitata Suhr. Biol. Bull. 156:257–71.CrossRefGoogle Scholar
Bird, K. T., Habig, C., and DeBusk, T. (1982). Nitrogen allocation and storage patterns in Gracilaria tikvahiae (Rhodophyta). J. Phycol. 18:344–8.CrossRefGoogle Scholar
Birrell, C. L., McCook, L. J., Willis, B. L., and Diaz-Pulido, G. A. (2008). Effects of benthic algae on the replenishment of corals and the implications for the resilience of coral reefs. Oceanogr. Mar. Biol. 46:25–63.Google Scholar
Bischof, K., and Steinhoff, F. (2012). Impacts of stratospheric ozone depletion and solar UVB radiation on seaweeds. In Wiencke, C., and Bischof, K. (eds), Seaweed Biology: Novel Insights into Ecophysiology, Ecology and Utilization, Ecological Studies, Vol. 219 (pp. 433–8). New York: Springer.CrossRefGoogle Scholar
Bischof, K., Hanelt, D., and Wiencke, C. (1998). UV-radiation can affect depth-zonation of Antarctic macroalgae. Mar Biol 131: 597–605.CrossRefGoogle Scholar
Bischof, K., Hanelt, D., and Wiencke, C. (2000). Effects of ultraviolet radiation on photosynthesis and related enzyme reactions of marine macroalgae. Planta 211:555–62.CrossRefGoogle ScholarPubMed
Bischof, K., Kräbs, G., Wiencke, C., and Hanelt, D. (2002a). Solar ultraviolet radiation affects the activity of ribulose-1,5-bisphosphate carboxylase/oxygenase and the composition of photosynthetic and xanthophyll cycle pigments in the intertidal green alga Ulva lactuca L. Planta 215: 502–9.CrossRefGoogle Scholar
Bischof, K., Peralta, G., Kräbs, G. et al. (2002b). Effects of solar UV-B radiation on canopy formation of natural Ulva communities from Southern Spain. J. Exp. Bot. 53:2411–21. [, , , ]CrossRefGoogle Scholar
Bischof, K., Janknegt, P. J., Buma, A. G. J., et al. (2003). Oxidative stress and enzymatic scavenging of superoxide radicals induced by solar UV-B radiation in Ulva canopies from southern Spain. Sci. Mar. 67:353–9.CrossRefGoogle Scholar
Bischof, K., Gómez, I., Molis, M. et al. (2006a). Ultraviolet radiation shapes seaweed communities. Rev. Environ. Sci. Bio-Technol. 5:141–66. [, , , ]CrossRefGoogle Scholar
Bischof, K., Rautenberger, R.Brey, L., and Perez-Llorens, J. L. (2006b). Physiological acclimation along gradients of solar irradiance within mats of the filamentous green macroalga Chaetomorpha linum from southern Spain. Mar. Ecol. Prog. Ser. 306:165–75.CrossRefGoogle Scholar
Bischoff, B., and Wiencke, C. (1995). Temperature adaptation in strains of the amphi-equatorial green alga Urospora penicilliformis (Acrosiphoniales): biogeographical implications. Mar. Biol. 122:681–8.CrossRefGoogle Scholar
Bischoff, K., and Rautenberger, R. (2012). Seaweed responses to environmental stress: reactive oxygen and antioxidative strategies In Wienke, C. and Bischof, K. (eds) Seaweed Biology: Novel Insights into Ecophysiology, Ecology and Utilization (pp. 109–33). New York: Springer.CrossRefGoogle Scholar
Bischoff-Bäsmann, B. (1997). Temperature requirements and biogeography of marine macroalgae: adaptation of marine macroalgae to low temperatures. Berichte zur Polarforschung 245, 134 pp.Google Scholar
Bischoff-Bäsmann, B., and Wiencke, C. (1996). Temperature requirements for growth and survival of Antarctic Rhodophyta. J. Phycol. 32:525–35.CrossRefGoogle Scholar
Bisgrove, S. R. (2007). Cytoskeleton and early development in fucoid algae. J. Integr. Plant Biol. 49:1192–8.CrossRefGoogle Scholar
Bisgrove, S. R., and Kropf, D. L. (2001). Cell wall deposition during morphogenesis in fucoid algae. Planta 212:648–58.CrossRefGoogle ScholarPubMed
Bisgrove, S. R., and Kropf, D. L. (2007). Asymmetric cell divisions: zygotes of fucoid algae as a model system. Cell Division Control in Plants; Series Plant Cell Monogr. 9:323–41.CrossRefGoogle Scholar
Bixler, H. J., and Porse, H. (2011). A decade of change in the seaweed hydrocolloids industry. J. Appl. Phycol. 23:321–35. [, , , ]CrossRefGoogle Scholar
Björk, M., Axelsson, L., and Beer, S. (2004). Why is Ulva intestinalis the only macroalga inhabiting isolated rockpools along the Swedish Atlantic coast? Mar. Ecol. Prog. Ser. 284:109–16.CrossRefGoogle Scholar
Björk, M., Mohammed, S. M., Björklund, M., and Semesi, A. (1995). Coralline algae, important coral reef builders threatened by pollution. Ambio. Stockholm 24:502–5.Google Scholar
Björn, L. O., Callaghan, T. V., Gehrke, C., Johanson, U., and Sonesson, M. (1999). Ozone depletion, ultraviolet radiation and plant life. Chemosphere 1:449–54.Google Scholar
Björnsater, B. R., and Wheeler, P. A. (1990). Effect of nitrogen and phosphorus supply on growth and tissue composition of Ulva fenestrata and Enteromorpha intestinalis (Ulvales, Chlorophyta). J. Phycol. 26:603–11.CrossRefGoogle Scholar
Black, W. A. P. (1949). Seasonal variation in chemical composition of some of the littoral seaweeds common to Scotland. Part II. Fucus serratus, Fucus vesiculosus, Fucus spiralis and Pelvetia canaliculata. J. Soc. Chem. Ind. 68:183–9.CrossRefGoogle Scholar
Black, W. A. P. (1950). The seasonal variation in weight and chemical composition of the common British Laminariaceae. J. Mar. Biol. Assoc. UK 29:45–72.CrossRefGoogle Scholar
Blanchette, C. A. (1996). Seasonal patterns of disturbance influence recruitment of the sea palm, Postelsia palmaeformis. J. Exp. Mar. Biol. Ecol. 197:1–14.CrossRefGoogle Scholar
Blanchette, C. A. (1997). Size and survival of intertidal plants in response to wave action: a case study with Fucus gardneri. Ecology 78:1563–78.CrossRefGoogle Scholar
Blanchette, C. A., Miner, C. M., Raimondi, P. T., et al. (2008). Biogeographical patterns of rocky intertidal communities along the Pacific coast of North America. J. Biogeogr. 35:1593–607.CrossRefGoogle Scholar
Blouin, N., Xiugeng, F., Peng, J., Yarish, C., and Brawley, S. H. (2007). Seeding nets with neutral spores of the red alga Porphyra umbilicals (L.) Kützing for use in integrated multi-trophic aquaculture (IMTA). Aquaculture 27:77–91.CrossRefGoogle Scholar
Blunden, G., and Gordon, S. M. (1986). Betaines and their sulphonio analogues in marine algae. Prog. Phycol. Res. 4:39–80.Google Scholar
Blunt, J. W., Copp, B. R., Hu, W.–P., et al. (2007). Marine natural products. Nat. Prod. Rep. 23:31–86.CrossRefGoogle Scholar
Boaventura, D., Alexander, M., Della Santina, P., et al. (2002). The effects of grazing on the distribution and composition of low-shore algal communities on the central coast of Portugal and on the southern coast of Britain. J. Exp. Mar. Biol. Ecol. 267:185–206.CrossRefGoogle Scholar
Bogorad, L. (1975). Phycobiliproteins and complementary chromatic adaptation. Annu. Rev. Plant Physiol. 26: 369–401.CrossRefGoogle Scholar
Bokn, T. (1987). Effects of diesel oil and subsequent recovery of commercial benthic algae. Hydrobiologia 151/152:277–84.CrossRefGoogle Scholar
Bokn, T.L., Moy, F. E., and Murray, S. N. (1993). Long-term effects of the water-accommodated fraction (WAF) of diesel oil on rocky shore populations maintained in experimental mesocosms. Bot. Mar. 36:313–19.CrossRefGoogle Scholar
Bokn, T. L., Duarte, C. M., Pedersen, M. F., et al. (2003). The response of experimental rocky shore communities to nutrient additions. Ecosystems 6:577–94.CrossRefGoogle Scholar
Bold, H. C., and Wynne, M. J. (1985). Introduction to the Algae: Structure and Reproduction, 2nd edn. Englewood Cliffs, NJ: Prentice Hall, 706 pp.Google Scholar
Boller, M. L., and Carrington, E. (2006a). In situ measurements of hydrodynamic forces imposed on Chondrus crispus Stackhouse. J. Exp. Mar. Biol. Ecol. 337:159–70.CrossRefGoogle Scholar
Boller, M. L., and Carrington, E. (2006b). The hydrodynamic effects of shape and size during reconfiguration of a flexible macroalga. J. Exp. Biol. 209:1895–903.CrossRefGoogle ScholarPubMed
Boller, M. L., and Carrington, E. (2007). Interspecific comparison of hydrodynamic performance and structural properties among intertidal macroalgae. J. Exp. Biol. 210:1874–84.CrossRefGoogle ScholarPubMed
Bolton, J. J. (1983). Ecoclinal variation in Ectocarpus siliculosus (Phaeophyceae) with respect to temperature growth optima and survival limits. Mar. Biol. 73:131–8.CrossRefGoogle Scholar
Bolwell, G. P, Callow, J. A., Callow, M. E., and Evans, L. V. (1979). Fertilization in brown algae. II. Evidence for lectinsensitive complementary receptors involved in gamete recognition in Fucus serratus. J. Cell Sci. 36:19–30.Google ScholarPubMed
Bolwell, G.P, Callow, J. A., and Evans, L. V. (1980). Fertilization in brown algae. III. Preliminary characterization of putative gamete receptors from eggs and sperm of Fucus serratus. J. Cell Sci. 43:209–24.Google ScholarPubMed
Bond, P. R., Brown, M. T., Moate, R. M., et al. (1999). Arrested development in Fucus spiralis (Phaeophyceae) germlings exposed to copper. Eur. J. Phycol. 34:513–21.CrossRefGoogle Scholar
Booth, D., Provan, J., and Maggs, C. A. (2007). Molecular approaches to the study of invasive seaweeds. Bot. Mar. 50:385–96.CrossRefGoogle Scholar
Borowitzka, L. M. (1986). Osmoregulation in blue-green algae. Prog. Phycol. Res. 4:243–56.Google Scholar
Borowitzka, M. A. (1977). Algal calcification. Oceanogr. Mar. Biol. Annu. Rev. 15:189–223.Google Scholar
Borowitzka, M. A. (1979). Calcium exchange and the measurement of calcification rates in the calcareous coralline red alga Amphiroa foliacea. Mar. Biol. 50:339–47.CrossRefGoogle Scholar
Borowitzka, M. A. (1982). Mechanisms in algal calcification. Prog. Phycol. Res. 1:137–77.Google Scholar
Borowitzka, M. A. (1987). Calcification in algae: mechanisms and the role of metabolism. CRC Crit. Rev. Plant Sci. 6:1–45.CrossRefGoogle Scholar
Borowitzka, M. A. (1989). Carbonate calcification in algae initiation and control. In Mann, S., Webb, J., and Williams, R. J. P. (eds), Biomineralization: Chemical and Biochemical Perspectives (pp. 63–94). Weinheim, Germany: VCH Publications.Google Scholar
Bothwell, J. H., Marie, D., Peters, A. F., Cock, J. M., and Coelho, S. M. (2010). Role of endoreduplication in the model brown alga Ectocarpus. New Phytol. 188:111–21. [, Fig. 2.4]CrossRefGoogle ScholarPubMed
Bouarab, K., Potin, P., Correa, J., and Kloareg, B. (1999). Sulfated oligosaccharides mediate the interaction between a marine red alga and its green algal pathogenic endophyte. Plant Cell 11:1635–50.CrossRefGoogle ScholarPubMed
Bouarab, K., Kloareg, B., Potin, P., and Correa, J. A. (2001). Ecological and biochemical aspects in algal infectious diseases. Cah. Biol. Mar. 42:91–100.Google Scholar
Bouarab, K., Adas, F., Gaquerel, E., B, et al. (2004). The innate immunity of a marine red alga involves oxylipins from both the eicosanoid and octadecanoid pathways. Plant Physiol. 135:1838–48.CrossRefGoogle ScholarPubMed
Bouzon, Z. L., and Ouriques, L. C. (2007). Characterization of Laurencia arbuscula spore mucilage and cell walls with stains and FITC-labelled lectins. Aquat. Bot. 86:301–8. [, , ]CrossRefGoogle Scholar
Bouzon, Z. L., Ouriques, L. C., and Oliveira, E. C. (2005). Ultrastructure of tetraspore germination in the agar-producing seaweed Gelidium floridanum (Gelidiales, Rhodophyta). Phycologia 44:409–15.CrossRefGoogle Scholar
Bouzon, Z. L., Ouriques, L. C., and Oliveira, E. C. (2006). Spore adhesion and cell wall formation in Gelidium floridanum (Rhodophyta, Gelidiales). J. Appl. Phycol. 18:287–94.CrossRefGoogle Scholar
Boyd, P. W., and Elwood, M. J. (2010). The biogeochemical cycle of iron in the ocean. Nature Geosci. 3:675–82.CrossRefGoogle Scholar
Boyen, C., Kloareg, B., and Vreeland, V. (1988). Comparison of protoplast wall regeneration and native wall deposition in zygotes of Fucus distichus by cell wall labelling with monoclonal antibodies. Plant Physiol. Biochem. 26:653–9.Google Scholar
Boyer, K. E., Fong, P., Armitage, A. R., and Cohen, R. A. (2004). Elevated nutrient content of tropical macroalgae increases rates of herbivory in coral, seagrass, and mangrove habitats. Coral Reefs 23:530–8.Google Scholar
Boyer, K. E., Kertesz, J. S., and Bruno, J. F. (2009). Biodiversity effects on productivity and stability of marine macroalgal communities: the role of environmental context. Oikos 118:1062–72.CrossRefGoogle Scholar
Boyle, T. P. (1985). Validation and Predictability of Laboratory Methods for Assessing the Fate and Effects of Contaminants in Aquatic Ecosystems. ASTM STP 865. Philadelphia: American Society for Testing and Materials.CrossRefGoogle Scholar
Bracken, M. E. S., and Nielsen, K. J. (2004). Diversity of intertidal macroalgae increases with nitrogen loading by invertebrates. Ecology 85:2828–36.CrossRefGoogle Scholar
Bracken, M. E. S., and Stachowicz, J. J. (2006). Seaweed diversity enhances nitrogen uptake via complementary use of nitrate and ammonium. Ecology 87:2397–403.CrossRefGoogle ScholarPubMed
Bracken, M. E. S., and Stachowicz, J. J. (2007). Top-down modification of bottom-up processes: selective grazing reduces macroalgal nitrogen uptake. Mar. Ecol. Prog. Ser. 330:75–82.CrossRefGoogle Scholar
Bracken, M. E. S., Gonzalez-Dorantes, C. A., and Stachowicz, J. J. (2007). Whole-community mutualism: associated invertebrates facilitate a dominant habitat-forming seaweed. Ecology 88:2211–19.CrossRefGoogle ScholarPubMed
Bracken, M. E. S., Jones, E., and Williams, S. L. (2011). Herbivores, tidal elevation, and species richness simultaneously mediate nitrate uptake by seaweed assemblages. Ecology 92:1083–93.CrossRefGoogle ScholarPubMed
Bradley, P M. (1991). Plant hormones do have a role in controlling growth and development of algae. J. Phycol. 27:317–21.CrossRefGoogle Scholar
Branch, G. M. (1975). Mechanisms reducing intraspecific competition in Patella spp.: migration, differentiation and territorial behaviour. J. Animal Ecol. 44:575–600.CrossRefGoogle Scholar
Branch, G. M., and Griffiths, C. L. (1988). The Benguela ecosystem. Part V. The coastal zone. Oceanogr. Mar. Bio. Annu. Rev. 26:395–486.Google Scholar
Brawley, S. H. (1987). A sodium-dependent, fast block to polyspermy occurs in eggs of fucoid algae. Devel. Biol. 124:390–7.CrossRefGoogle ScholarPubMed
Brawley, S. H. (1991). The fast block against polyspermy in fucoid algae is an electrical block. Devel. Biol. 144: 94–106.CrossRefGoogle ScholarPubMed
Brawley, S. H. (1992a). Fertilization in natural populations of the dioecious brown alga Fucus ceranoides and the importance of the polyspermy block. Mar. Biol. 113:145–57.CrossRefGoogle Scholar
Brawley, S. H. (1992b). Mesoherbivores. In John, D. M., Hawkins, S. J., and Price, J. H. (eds), Plant–Animal Interactions in the Marine Benthos (pp. 235–63). Oxford: Oxford University Press.Google Scholar
Brawley, S. H., and Adey, W. H. (1977). Territorial behavior of three-spot damselfish (Eupomacentrus planifrons) increases reef algal biomass and productivity. Env. Biol. Fish. 2:45–51.CrossRefGoogle Scholar
Brawley, S. H., and Adey, W. H. (1981). The effect of micrograzers on algal community structure in a coral reef microcosm. Mar. Biol. 61:167–77.CrossRefGoogle Scholar
Brawley, S. H., and Fei, X. G. (1987). Studies of mesoherbivory in aquaria and in an unbarricaded mariculture farm on the Chinese coast. J. Phycol. 23:614–23.CrossRefGoogle Scholar
Brawley, S. H., and Johnson, L. E. (1992). Gametogenesis, gametes, and zygotes – an ecological perspective on sexual reproduction in the algae. Brit. Phycol. J. 27:233–52.CrossRefGoogle Scholar
Breeman, A. M. (1988). Relative importance of temperature and other factors in determining geographic boundaries of seaweeds: experimental and phenological evidence. Helgol Meeresunters 42: 199–241.CrossRefGoogle Scholar
Breeman, A. M. (1990). Expected effects of changing seawater temperatures on the geographic distribution of seaweed species. In Beukema, J. J., Wolff, W. J., and Brouns, J. W. M. (eds) Expected Effects of Climate Change on Coastal Ecosystems (pp. 69–76). Amsterdam: Kluwer Academic Publishers.CrossRefGoogle Scholar
Breeman, A. M., and Guiry, M. D. (1989). Tidal influences on the photoperiodic induction of tetrasporogenesis in Bonnemaisonia hamifera (Rhodophyta). Mar. Biol. 102:5–14.CrossRefGoogle Scholar
Breeman, A. M., Bos, S., van Essen, S., and van Mulekom, L. L. (1984). Light-dark regimes in the intertidal zone and tetrasporangial periodicity in the red alga Rhodochorton purpureum (Lightf.) Rosenv. Helgol. Meeresunters. 37:365–87.CrossRefGoogle Scholar
Breeman, A. M., Meulenhoff, E. J. S., and Guiry, M. D. (1988). Life history regulation and phenology of the red alga Bonnemaisonia hamifera. Helgol. Meeresunters. 42:535–51.CrossRefGoogle Scholar
Breen, P. A., and Mann, K. H. (1976). Changing lobster abundance and the destruction of kelp beds by sea urchins. Mar. Biol. 34:137–42.CrossRefGoogle Scholar
Brekke, C., and Solberg, A. H. S. (2005). Oil detection by satellite remote sensing. Rem. Sens. Environ. 95:1–13.CrossRefGoogle Scholar
Brenchley, J. L., Raven, J. A., and Johnston, A. M. (1996). A comparison of reproductive allocation and reproductive effort between semelparous and iteroparous fucoids (Fucales, Phaeophyta). Hydrobiologia 326/327:185–90.CrossRefGoogle Scholar
Brinkhuis, B. H. (1985). Growth patterns and rates. In Littler, M. M and Littler, D. S. (eds), Handbook of Phycological Methods: Ecological Field Methods: Macroalgae (pp. 461–77). Cambridge: Cambridge University Press.Google Scholar
Brinkhuis, B. H., Levine, H. G., Schlenk, C. G., and Tobin, S. (1987). Laminaria cultivation in the Far East and North America. In Bird, K. T., and Benson, P. H. (eds), Seaweed Cultivation for Renewable Resources (pp. 107–46). Amsterdam: Elsevier.Google Scholar
Brinkhuis, B. H., Li, R., Wu, C., and Jiang, X. (1989). Nitrite uptake transients and consequences for in vivo algal nitrate reductase assays. J. Phycol. 25:539–45.CrossRefGoogle Scholar
Brinza, L., Nygård, C. A., Dring, M. J., Gavrilescu, M., and Benning, L. G. (2009). Cadmium tolerance and adsorption by the marine brown alga Fucus vesiculosus from the Irish Sea and Bothnian Sea. Bioresource Technol. 100:1727–33.CrossRefGoogle ScholarPubMed
Britz, S. J., and Briggs, W. R. (1976). Circadian rhythms of chloroplast orientation and photosynthetic capacity inUlva. Plant Physiol. 58:22–7.CrossRefGoogle Scholar
Broadwater, S. T., and Scott, J. (1982). Ultrastructure of early development in the female reproductive system of Polysiphonia harveyi Bailey (Ceramiales, Rhodophyta). J. Phycol. 18:427–41.CrossRefGoogle Scholar
Brodie, J. A., and Lewis, J. (2007). Unravelling the Algae: the Past, Present and Future of Algae Systematics. Boca Raton, FL: CRC Press Inc., 376 pp.CrossRefGoogle Scholar
Brodie, J., Andersen, R. A., Kawachi, M., and Millar, A. J. K.. (2009). Endangered algal species and how to protect them. Phycologia 48:423–38.CrossRefGoogle Scholar
Brodie, J. E., Kroon, F. J., Schaffelke, B., et al. (2012). Terrestrial pollutant runoff to the Great Barrier Reef: an update of issues, priorities and management responses. Mar. Poll. Bull. 65:81–100.CrossRefGoogle ScholarPubMed
Brokovich, E., Ayalon, I., Einbinder, S., et al. (2010). Grazing pressure on coral reefs across a wide depth gradient in the Gulf of Aqaba, Red Sea. Mar. Ecol. Prog. Ser. 399:69–80.CrossRefGoogle Scholar
Brown, J. H. (1995). Macroecology. Chicago, IL: University of Chicago Press.Google Scholar
Brown, M. T., and Newman, J. E. (2003). Physiological responses of Gracilariopsis longissima (S. G. Gmelin) Steentoft L. M. Irvine and Farnham (Rhodophyceae) to sublethal copper concentrations. Aquat. Toxic. 64:201–13.CrossRefGoogle Scholar
Brown, M. T., Newman, J. E., and Han, T. (2012). Inter-population comparisons of copper resistance and accumulation in the red seaweed, Gracilariopsis longissima. Ecotox. 21:591–600.CrossRefGoogle ScholarPubMed
Bruhn, J., and Gerard, V. A. (1996). Photoinhibition and recovery of the kelp Laminaria saccharina at optimal and superoptimal temperatures. Mar. Biol. 125: 639–48.CrossRefGoogle Scholar
Bruland, K. W. (1983). Trace elements in sea-water. In Riley, J. P. (ed.), Chemical Oceanography 8 (pp. 157–219). London: Academic Press.CrossRefGoogle Scholar
Bruland, K. W., and Lohan, M. C. (2004). Controls of trace metals in seawater. Treatise on Geochemistry 6:23–47.Google Scholar
Bruno, E., and Eklund, B. (2003). Two new growth inhibition tests with the filamentous algae Ceramium strictum and C. tenuicorne (Rhodophyta). Envir. Poll. 125:287–93.CrossRefGoogle Scholar
Bruno, J. F., and Bertness, M. D. (2001). Habitat modification and facilitation in benthic marine communities. In Bertness, M. D., Gaines, S. D., and Hay, M. E. (eds), Marine Community Ecology (pp. 201–18). Sunderland, MA: Sinauer Associates.Google Scholar
Bruno, J. F., Sweatman, H., Precht, W. F., Selig, E. R., and Schutte, V. G. W. (2009). Assessing evidence of phase shifts from coral to macroalgal dominance on coral reefs. Ecology 90:1478–84.CrossRefGoogle ScholarPubMed
Brzezinski, M. A., Reed, D. C., and Amsler, C. D. (1993). Neutral lipids as major storage products in zoospores of the giant-kelp Macrocystis pyrifera (Phaeophyceae). J. Phycol. 29:16–23.CrossRefGoogle Scholar
Buchanan, B. B., Gruissem, W., and Jones, R. L. (eds) (2000). Biochemistry and Molecular Biology of Plants. Rockville, MD: ASPP, 1367 pp. [, , , , 5, , , ]Google Scholar
Büchel, C., and Wilhelm, C. (1993). In vivo analysis of slow chlorophyll fluorescence induction kinetics in algae: progress, problems and perspectives. Photochem. Photobiol. 58:137–48.CrossRefGoogle Scholar
Buchholz, C. M., Krause, G., and Buck, B. H. (2012). Seaweed and man. In Wiencke, C. and Bischof, K. (eds) Seaweed Biology: Novel Insights into Ecophysiology, Ecology and Utilization (pp. 471–93). New York: Springer. [, , , ]CrossRefGoogle Scholar
Buck, B. H., and Buchholz, C. M. (2004). The offshore ring: a new system design for the open ocean aquaculture of macroalgae. J. Appl. Phycol. 16:355–68.CrossRefGoogle Scholar
Buck, B. H., and Buchholz, C. M. (2005). Response of offshore cultivated Laminaria saccharina to hydrodynamic forcing in the North Sea. Aquaculture 250:674–91.CrossRefGoogle Scholar
Buesseler, K., Aoyama, M., and Fukasawa, M. (2011). Impacts of the Fukushima nuclear power plants on marine radioactivity. Envir. Sci. Technol. 45:9931–5.CrossRefGoogle ScholarPubMed
Buesseler, K., Jayne, S. R., Fisher, N. S., et al. (2012). Fukushima-derived radionuclides in the ocean and biota off Japan. PNAS 109:5984–8.CrossRefGoogle ScholarPubMed
Buggeln, R. G. (1974). Negative phototropism of the haptera of Alaria esculenta (Laminariales). J. Phycol. 10:80–2.Google Scholar
Buggeln, R. G. (1981). Morphogenesis and growth substances. In Lobban, C. S. and Wynne, M. J. (eds), The Biology of Seaweeds (pp. 627–60). Oxford: Blackwell Scientific.Google Scholar
Buggeln, R. G. (1983). Photoassimilate translocation in brown algae. Prog Phycol Res 2:283–332.Google Scholar
Buggeln, R. G., Fensom, D. S., and Emerson, C. J. (1985). Translocation of 11C-photoassimilate in the blade of Macrocystis pyrifera (Phaeophyceae). J Phycol. 21:35–40.CrossRefGoogle Scholar
Bulleri, F. (2006). Duration of overgrowth affects survival of encrusting coralline algae. Mar. Ecol. Prog. Ser. 321:79–85.CrossRefGoogle Scholar
Bulleri, F. (2009). Facilitation research in marine systems: state of the art, emerging patterns and insights for future developments. J. Ecol. 97:1121–30.CrossRefGoogle Scholar
Bulleri, F., and Benedetti-Cecchi, L. (2008). Facilitation of the introduced green alga Caulerpa racemosa by resident algal turfs: experimental evaluation of underlying mechanisms. Mar. Ecol. Prog. Ser. 364:77–86CrossRefGoogle Scholar
Bulleri, F., Balata, D., Bertocci, I., Tamburello, L., and Benedetti–Cecchi, L. (2010). The seaweed Caulerpa racemosa on Mediterranean rocky reefs: from passenger to driver of ecological change. Ecology 91:2205–12.CrossRefGoogle ScholarPubMed
Burgess, R. M., Pelletier, M. C., Ho, K. T., et al. (2003). Removal of ammonia toxicity in marine sediment TIEs: a comparison of Ulva lactuca, xeolite and aeration methods. Mar. Poll. Bull. 46:607–18.CrossRefGoogle ScholarPubMed
Burgman, M. A., and Gerard, V. A. (1990). A stage-structured, stochastic population model for the giant kelp, Macrocystis pyrifera. Mar. Biol. 105: 15–23.CrossRefGoogle Scholar
Burke, L., Reytar, K., Spalding, M., and Perry, A. (2011). Reefs at Risk Revisited. Washington, DC: World Resources Institute.Google Scholar
Burkhardt, E., and Peters, A. F. (1998). Molecular evidence from nrDNA its sequences that Laminariocolax (Phaeophyceae, Ectocarpales sensu lato) is a worldwide clade of closely related kelp endophytes. J. Phycol. 34:682–91.CrossRefGoogle Scholar
Burki, F., Okamoto, N., Pombert, J.-F., and Keeling, P. J. (2012). The evolutionary history of haptophytes and cryptophytes: phylogenomic evidence for separate origins. Proc. R. Soc. B. 179:2246–54.CrossRefGoogle Scholar
Burridge, T. R., and Bidwell, J. (2002). Review of the potential use of brown algal ecotoxicological assays in monitoring effluent discharge and pollution in southern Australia. Mar. Poll. Bull. 45:140–7.CrossRefGoogle ScholarPubMed
Burridge, T. R., and Shir, M. A. (1995). The comparative effects of oil dispersants and oil/dispersant conjugates on germination of the marine macroalga Phyllospora comosa (Fucales: Phaeophyta). Mar. Poll. Bull. 31:446–52.CrossRefGoogle Scholar
Burritt, D. J., Larkindale, J., and Hurd, C. L. (2002). Antioxidant metabolism in the intertidal red seaweed Stictosiphonia arbuscula following desiccation. Planta 215: 829–38.CrossRefGoogle ScholarPubMed
Burrows, M. T., Harvey, R., and Robb, L. (2008). Wave exposure indices from digital coastlines and the prediction of rocky shore community structure. Mar. Ecol. Prog. Ser. 353:1–12.CrossRefGoogle Scholar
Burrows, M. T., Harvey, R., Robb, L., et al. (2009). Spatial scales of variance in abundance of intertidal species: effects of region, dispersal mode, and trophic level. Ecology 90:1242–54.CrossRefGoogle ScholarPubMed
Burtin, P. (2003). Nutritional value of seaweeds. J. Environmental Agric. Food Chem. 2:1–9.Google Scholar
Buschmann, A. H., and Bravo, A. (1990). Intertidal amphipods as potential dispersal agents of carpospores of Iridaea laminarioides (Gigartinales, Rhodophyta). J. Phycol. 26:417–20.CrossRefGoogle Scholar
Buschmann, A. H., Westermeier, R., and Retamales, C. A. (1995). Cultivation of Gracilaria on the sea-bottom in southern Chile: a review. J. Appl. Phycol. 7:291–301.CrossRefGoogle Scholar
Buschmann, A. H., Correa, J. A., Westermeier, R., Hernández-Gonález, M. C., and Norambuena, R. (2001). Red algal farming in Chile: a review. Aquaculture 194:203–20.CrossRefGoogle Scholar
Buschmann, A. H., Vásquez, J. A., Osorio, P., et al. (2004). The effect of water movement, temperature and salinity on abundance and reproductive patterns of Macrocystis spp. (Phaeophyta) at different latitudes in Chile. Mar. Biol. 145:849–62.CrossRefGoogle Scholar
Buschmann, A. H., Moreno, C., Vásquez, J. A., and Hernández-González, M. C. (2006). Reproduction strategies of Macrocystis pyrifera (Phaeophyta) in Southern Chile: the importance of population dynamics. J. Appl. Phycol. 18:575–82.CrossRefGoogle Scholar
Buschmann, A. H., Hernández-Gonález, M. C., and Varela, D. (2008). Seaweed future cultivation in Chile: perspectives and challenges. Internat. J. Envir. Poll. 33:432–55. [, , , ]CrossRefGoogle Scholar
Butler, D. M., Ostgaard, K., C. et al. (1989). Isolation conditions for high yields of protoplasts from Laminaria saccharaina and L. digitata (Phaeophyceae). J. Exp. Bot. 40: 1237–46.CrossRefGoogle Scholar
Butler, J. N., Morris, B. F. and Sleeter, T. D. (1976). The fate of petroleum in the open ocean. In Sources, Effects, and Sinks of Hydrocarbons in the Aquatic Environment (pp. 287–97). Washington DC: American Institute of Biological Sciences.Google Scholar
Butterworth, J., Lester, P., and Nickless, G. (1972). Distribution of heavy metals in the Severn estuary. Mar. Pollut. Bull. 3:72–4. [8.3.5]CrossRefGoogle Scholar
Cabello–Pasini, A., and Alberte, R. S. (1997). Seasonal patterns of photosynthesis and light-independent carbon fixation in marine macrophytes. J. Phycol. 33: 321–9.CrossRefGoogle Scholar
Cabello–Pasinsi, A., and Alberte, R. S. (2001). Enzymatic regulation of photosynthetic and light-independent carbon fixation in Laminaria setchelli (Phaeophyta), Ulva lactuca (Chlorophyta) and Iridea cordata (Rhodophyta). Rev. Chil. Hist. Nat. 74:226–36.Google Scholar
Cabioch, J. (1988). Morphogenesis and generic concepts in coralline algae-a reappraisal. Helgol. Meeresunters. 42:493–509.CrossRefGoogle Scholar
Cabioch, J., and Giraud, G. (1986). Structural aspects of biomineralization in the coralline algae (calcified Rhodophyceae). In Leadbeater, B. S. C. and Riding, R. (eds), Biomineralization in Lower Plants and Animals (pp. 141–56). Oxford: Clarendon Press.Google Scholar
Cáceres, E. J., and Parodi, E. R. (1998). Fine structure of zoosporogenesis, zoospore germination, and early gametophyte development in Cladophora surera (Cladophorales, Chlorophyta). J. Phycol. 34:825–34.CrossRefGoogle Scholar
Cairrão, E., Couderchet, M., Soares, A. M., and Guilhermino, L. (2004). Glutathione-S-transferase activity of Fucus spp. as a biomarker of environmental contamination. Aquat. Toxicol. 70:277–86.CrossRefGoogle ScholarPubMed
Caldeira, K., and Wickett, M. E. (2003). Anthropogenic carbon and ocean pH. Nature 425:365–7.CrossRefGoogle ScholarPubMed
Callaway, R. M. (2007). Positive Interactions and Interdependence in Plant Communities. Dordrecht, The Netherlands: Springer, 415 pp.Google Scholar
Callow, J. A., and Callow, M. E. (2006). The Ulva spore adhesive system. In Smith, A. M., and Callow, J. A. (eds), Biological Adhesives (pp. 63–78). Berlin and Heidelberg: Springer-Verlag.CrossRefGoogle Scholar
Callow, J. A., Callow, M. E., and Evans, L. V. (1979). Nutritional studies on the parasitic red algaChoreocolax polysiponiae. New Phytol. 83:451–62.CrossRefGoogle Scholar
Callow, M. E., Evans, L. V., Bolwell, G. P., and Callow, J. A. (1978). Fertilization in brown algae. I. SEM and other observations on Fucus serratus. J. Cell. Sci. 32:45–54.Google Scholar
Callow, M. E., Callow, J. A., Pickett-Heaps, J. D., and Wetherbee, R. (1997). Primary adhesion of Enteromorpha (Chlorophyta, Ulvales) propagules: quantitative settlement studies and video microscopy. J. Phycol. 33:938–47.CrossRefGoogle Scholar
Camargo, J. A., and Alonso, A. (2006). Ecological and toxicological effects of inorganic nitrogen pollution in aquatic ecosystems: a global assessment. Environ. Internat. 32:832–49.CrossRefGoogle ScholarPubMed
Cambridge, M., Breeman, A. M., and van den Hoek, C. (1984). Temperature responses of some North Atlantic Cladophora species (Chlorophyceae) in relation to their geographic distribution. Helgol Meeresunters 38: 349–63.CrossRefGoogle Scholar
Camilli, R., Reddy, C. M., Yoerger, D., et al. (2010). Tracking hydrocarbon plume transport and biodegradation at Deepwater Horizon. Science 330:201–4.CrossRefGoogle ScholarPubMed
Campbell, J. W., and Aarup, T. (1989). Photosynthetically available radiation at high latitudes. Limnol. Oceanogr. 34:1490–9.CrossRefGoogle Scholar
Cancino, J. M., Muñoz, J., Muñoz, M., and Orellana, M. C. (1987). Effects of the Bryozoan Membranipora tuberculata (Bosc.) on the photosynthesis and growth of Gelidium rex Santelices et Abbott. J. Exp. Mar. Biol. Ecol. 113:105–12.CrossRefGoogle Scholar
Canuel, E. A., Cloern, J. N., Ringelberg, D. B., Guckert, J. B., and Rau, G. H. (1995). Molecular and isotopic tracers used to examine source of organic matter and its incorporation into the food webs of San Francisco Bay. Limnol. Oceanogr. 40(1):67–81CrossRefGoogle Scholar
Capone, D. G., and Bautista, M. F. (1985). A groundwater source of nitrate in nearshore marine sediments. Nature 313:214–16.CrossRefGoogle Scholar
Carballo, J. L., Olabarria, C., and Osuna, T. G. (2002). Analysis of four macroalgal assemblages along the Pacific Mexican Coast during and after the 1997–1998 El Niño. Ecosystems 5:749–60.CrossRefGoogle Scholar
Carillo, S., López, E., Casas, M. M., et al. (2008). Potential use of seaweeds in the laying hen ration to improve the quality of n-3 fatty acids enriched eggs. J. Appl. Phycol. 20:721–8.CrossRefGoogle Scholar
Carney, L. T. (2011). A multispecies laboratory assessment of rapid sporophyte recruitment from delayed kelp gametophytes. J. Phycol. 47:244–51.CrossRefGoogle ScholarPubMed
Caron, L., Dubacq, J. P., Berkaloff, C., and Jupin, H. (1985). Subchloroplast fractions from the brown alga Fucus serratus: phosphatidylglycerol contents. Plant Cell Physiol 26:131–9.Google Scholar
Caron, L., Remy, R., and Berkaloff, C. (1988). Polypeptide composition of light harvesting complexes from some brown algae and diatoms. FEBS Letters 229:11–15.CrossRefGoogle Scholar
Carpenter, R. C. (1988). Mass mortality of a Caribbean sea urchin: immediate effects on community metabolism and other herbivores. Proc. Natl. Acad. Sci. USA 85:511–14.CrossRefGoogle ScholarPubMed
Carpenter, R. C. (1990). Competition among marine macroalgae: a physiological perspective. J. Phycol. 26:6–12. [4.0, , , , ]CrossRefGoogle Scholar
Carpenter, R. C., and Edmunds, P. J. (2006). Local and regional-scale recovery of Diadema promotes recruitment of scleractinian corals. Ecol. Lett. 9:268–77.CrossRefGoogle ScholarPubMed
Carpenter, R. C., Hackney, J. M., and Adey, W. H. (1991). Measurements of primary productivity and nitrogenase activity of coral reef algae in a chamber incorporating oscillatory flow. Limnol. Oceanogr. 36:40–9.CrossRefGoogle Scholar
Carrington, E. (1990). Drag and dislodgement of an intertidal macroalga: consequences of morphological variation in Mastocarpus papillatus Kiitzing. J. Exp. Mar. Biol. Ecol. 139:185–200.CrossRefGoogle Scholar
Carrington, E., Grace, S. P., and Chopin, T. (2001). Life history phases and the biomechanical properties of the red alga Chondrus crispus (Rhodophyta). J. Phycol. 37:699–704.CrossRefGoogle Scholar
Cashmore, A. R. (2005). Cryptochrome overview. In: Wada, M., Shimazaki, K., and Lino, M. (eds), Light Sensing in Plants (pp.121–30). Tokyo: Springer-Verlag (Yamada Science Foundation and Springer-Verlag Tokyo).CrossRefGoogle Scholar
Cassab, G. I. (1998). Plant cell wall proteins. Annu Rev. Plant Phys. 49:281–309.CrossRefGoogle ScholarPubMed
Castilla, J. C. (1988). Earthquake-caused coastal uplift and its effects on rocky intertidal kelp communities. Science 242:440–3.CrossRefGoogle ScholarPubMed
Castilla, J. C., Manriquez, P. H., and Camaño, A. (2010). Effects of rocky shore coseismic uplift and the 2010 Chilean mega-earthquake on intertidal biomarker species. Mar. Ecol. Prog. Ser. 418:17–23.CrossRefGoogle Scholar
Caut, S., Angulo, E., and Courchamp, F. (2009). Variation in discrimination factors (Delta N-15 and Delta C-13): the effect of diet isotopic values and applications for diet reconstruction. Journal of Applied Ecology 46(2):443–53CrossRefGoogle Scholar
Cavalier-Smith, T. (1999). Principles of protein and lipid targeting in secondary symbiogenesis: euglenoid, dinoflagellate, and sporozoan plastid origins and the eukaryote family tree. J. Eukaryot. Microbiol. 46:347–66.CrossRefGoogle ScholarPubMed
Cavanaugh, K. C., Siegel, D. A., Kinlan, B. P., and Reed, D. C. (2010). Scaling giant kelp field measurements to regional scales using satellite observations. Mar. Ecol. Prog. Ser. 403:13–27.CrossRefGoogle Scholar
Ceccarelli, D. M., Jones, G. P., and McCook, L. J. (2001). Territorial damselfishes as determinants of the structure of benthic communities on coral reefs. Oceanogr. Mar. Biol. 39:355–89.Google Scholar
Cecere, E., Petrocelli, A., and Verlaque, M. (2011). Vegetative reproduction by multicellular Rhodophyta: an overview. Mar. Ecol. 32:419–37.CrossRefGoogle Scholar
Cembella, A., Antia, N. J., and Harrison, P. J. (1984). The utilization of inorganic and organic phosphorus compounds as nutrients by eukaryotic microalgae: a multidisciplinary perspective. CRC Crit. Rev. Microbiol. 10:317–91.CrossRefGoogle ScholarPubMed
Cerda, O., Hinojosa, I. A., and Thiel, M. (2010). Nest-building by the amphipod Peramphithoe femorata (Kroyer) on the kelp Macrocystis pyrifera (Linnaeus) C. Agardh from northern-central Chile. Biol. Bull. 218:248–58.CrossRefGoogle ScholarPubMed
Cetrulo, G. L., and Hay, M. E. (2000). Activated chemical defenses in tropical versus temperate seaweeds. Mar. Ecol. Prog. Ser. 207:243–53.CrossRefGoogle Scholar
Chamberlain, Y. M. (1984). Spore size and germination in Fosliella, Pneophyllum and Melobesia (Rhodophyta, Corallinaceae). Phycologia 23:433–42.CrossRefGoogle Scholar
Chan, C. X., Blouin, N. A., Zhuang, Y., et al. (2012a). Porphyra (Bangiophyceae) transcriptomes provide insights into red algal development and metabolism. J. Phycol. 48:1328–42. [, , ]CrossRefGoogle ScholarPubMed
Chan, C. X., Zauner, S., Wheeler, G. L., et al. (2012b). Analysis of Porphyra membrane transporters demonstrates gene transfer among photosynthetic eukaryotes and numerous sodium-coupled transport systems. Plant Physiol. 158:2001–12.CrossRefGoogle ScholarPubMed
Chan, K., Wong, P. K., and Ng, S. L. (1982). Growth of Enteromorpha linza in sewage effluent and sewage effluent-seawater mixtures. Hydrobiologia 97:9–13CrossRefGoogle Scholar
Chapman, A. R. O. (l973). Methods for macroscopic algae. In Stein, J. R. (ed.), Handbook of Phycological Methods: Culture Methods and Growth Measurements (pp. 87–104). Cambridge: Cambridge University Press.Google Scholar
Chapman, A. R. O. (1984). Reproduction, recruitment and mortality in two species of Laminaria in south-west Nova Scotia. J. Exp. Mar. Biol. Ecol. 78:99–108.CrossRefGoogle Scholar
Chapman, A. R. O. (1985). Demography. In Littler, M. M. and Littler, D. S. (eds), Handbook of Phycological Methods: Ecological Field Methods: Macroalgae (pp. 251–68). Cambridge: Cambridge University Press.Google Scholar
Chapman, A. R. O. (1986). Population and community ecology of seaweeds. In Blaxter, J. H. S. and Southward, A. J. (eds), Advances in Marine Biology (vol. 23, pp. 1–161). London: Academic Press. [, , , ]Google Scholar
Chapman, A. R. O. (1987). The wild harvest and culture of Laminaria longicruris in eastern Canada. In Doty, M. S., Caddy, J. F., and Santelices, B. (eds), Case Studies of Seven Commercial Seaweed Resources (pp. 193–237). FAO Fish. Tech. Pap. 281.Google Scholar
Chapman, A. R. O. (1993). Hard data for matrix modeling of Laminaria digitata (Laminariales, Phaeophyta) populations. Hydrobiologia 261:263–7.CrossRefGoogle Scholar
Chapman, A. R. O. (1995). Functional ecology of fucoid algae: twenty-three years of progress. Phycologia 34:1–32.CrossRefGoogle Scholar
Chapman, A. R. O., and Burrows, E. M. (1970). Experimental investigations into the controlling effects of light conditions on the development and growth of Desmarestia aculeata (L.) Lamour. Phycologia 9:103–8.CrossRefGoogle Scholar
Chapman, A. R. O., and Craigie, J. S. (1977). Seasonal growth in Laminaria longicruris: relations with dissolved inorganic nutrients and internal reserves of nitrogen. Mar. Biol. 40:197–205.CrossRefGoogle Scholar
Chapman, A. R. O., and Goudey, C. L. (1983). Demographic study of the macrothallus of Leathesia difformis (Phaeophyta) in Nova Scotia. Can J. Bot. 61:319–23.CrossRefGoogle Scholar
Chapman, A. R. O., Markham, J. W., and Lüning, K. (1978). Effects of nitrate concentration on the growth and physiology of Laminaria saccharina (Phaeophyta) in culture. J. Phycol. 14:195–8.CrossRefGoogle Scholar
Chapman, A. S. (1999). From introduced species to invader: what determines variation in the success of Codium fragile ssp. tomentosoides (Chlorophyta) in the North Atlantic Ocean? Helgolander. Meeresun. 52:277–89.CrossRefGoogle Scholar
Chapman, M. G., and Underwood, A. J. (1998). Inconsistency and varation in the development of rocky intertidal algal assemblages. J. Exp. Mar. Biol. Ecol. 224: 265–89.CrossRefGoogle Scholar
Chapman, M. S., Suh, S. W., Curmi, P. M. G., et al. (1988). Tertiary structure of plant RuBisCO: domains and their contacts. Science 241: 71–4.CrossRefGoogle ScholarPubMed
Charpy, L., Casareto, B. E., Langlade, M. J., and Suzuki, Y. (2012). Cyanobacteria in coral reef ecosystems: a review. J. Mar. Biol. 2012 Article ID 259571.CrossRef
Charrier, B., Coelho, S. M., Le Bail, A., et al. (2008). Development and physiology of the brown alga Ectocarpus siliculosus: two centuries of research. New Phytol. 177:319–32. [, , ]CrossRefGoogle Scholar
Charters, A. C., Neushul, M., and Barilotti, C. (1969). The functional morphology of Eisenia arborea. Proc. Intl. Seaweed Symp. 6:89–105.Google Scholar
Chavez, F. P., Strutton, P. G., Friederich, G. E., et al. (1999). Biological and chemical response of the Equatorial Pacific Ocean to the 1997–1998 El Niño. Science 286: 2126–31.CrossRefGoogle Scholar
Chen, L. C.–M. (1986). Cell development of Porphyra miniata (Rhodophyceae) under axenic culture. Bot. Mar. 29:435–9.CrossRefGoogle Scholar
Chen, L. C.–M., Edelstein, T., Ogata, E., and McLachlan, J. (1970). The life history of Porphyra miniata. Can. J. Bot. 48:385–9.CrossRefGoogle Scholar
Cheshire, A. C., and Hallam, N. D. (1985). The environmental role of alginates in Durvillaea potatorum (Fucales, Phaeophyta). Phycologia 24:147–53.CrossRefGoogle Scholar
Cheshire, A. C., and Hallam, N. D. (1989). Methods for assessing the age composition of native stands of subtidal macroalgae: a case study on Durvillaea potatorum. Bot. Mar. 32:199–204.CrossRefGoogle Scholar
Chisolm, J. R. M. (2000). Calcification by crustose coralline algae on the northern Great Barrier Reef, Australia. Limnol. Oceanogr. 45:1476–84.CrossRefGoogle Scholar
Chisholm, J. R. M., Marchioretti, M., and Jaubert, J. M. (2000). Effect of low water temperature on metabolism and growth of a subtropical strain of Caulerpa taxifolia (Chlorophyta). Mar. Ecol. Prog. Ser. 201:189–98.CrossRefGoogle Scholar
Choi, D., Kim, J. H., and Lee, Y. (2008). Expansins in plant development. Adv. Bot. Res. 47:47–97.CrossRefGoogle Scholar
Choi, H.–G., and Lee, I. K. (1996). Mixed-phase reproduction in Dasysiphonia chejuensis (Rhodophyta) from Korea. Phycologia 35:9–18.CrossRefGoogle Scholar
Choi, H. G., and Norton, T. A. (2005a). Competition and facilitation between germlings of Ascophyllum nodosum and Fucus vesiculosus. Mar. Biol. 147:525–32.CrossRefGoogle Scholar
Choi, H. G., and Norton, T. A. (2005b). Competitive interactions between two fucoid algae with different growth forms, Fucus serratus and Himanthalia elongata. Mar. Biol. 146:283–91.CrossRefGoogle Scholar
Choo, K. S., Nilsson, J., Pedersen, M., and Snoeijs, P. (2005). Photosynthesis, carbon uptake and antioxidant defense in two coexisting filamentous green algae under different stress conditions. Mar. Ecol. Prog. Ser. 292:127–38.CrossRefGoogle Scholar
Chopin, T. (2011) Progression of the Integrated Multi-Trophic Aquaculture (IMTA) concept and upscaling of IMTA systems towards commercialization. Aquacult. Europe 36:5–12. [, , ]Google Scholar
Chopin, T. (2012) Aquaculture, Integrated Multi-Trophic (IMTA) In Meyers, R.A. (ed.), Encyclopedia of Sustainability Science and Technology (pp. 542–64). Dordrecht, The Netherlands: Springer.CrossRefGoogle Scholar
Chopin, T., and Sawhney, M. (2009). Seaweeds and their mariculture. In Steele, J. H, Thorpe, S. A., and Turekian, K. K. (eds), The Encyclopedia of Ocean Sciences (pp. 4477–87). Oxford: Elsevier.Google Scholar
Chopin, T. and Wagey, B. T. (1999). Factorial study of the effects of phosphorus and nitrogen enrichments on nutrient and carrageenan content in Chondrus crispus (Rhodophyceae) and on residual nutrient concentration in seawater. Bot. Mar. 42:23–31.CrossRefGoogle Scholar
Chopin, T., Hourmat, A., Floc’h, J.–Y., and Penot, M. (1990). Seasonal variations in the red alga Chondrus crispus on the Atlantic French coast. II. Relations with phosphorus concentration in seawater and intertidal phosphorylated fractions. Can. J. Bot. 68:512–17.CrossRefGoogle Scholar
Chopin, T., Lehmal, H., and Halcrow, K. (1997). Polyphosphates in the red macroalga Chondrus crispus (Rhodophyta). New Phytol. 135:587–94.CrossRefGoogle Scholar
Chopin, T., Kerin, B. F., and Mazerolle, R. (1999) Phycocolloid chemistry as a taxonomic indicator of phylogeny in the Gigartinales, Rhodophyceae: a review and current developments using Fourier transform infrared diffuse reflectance spectroscopy. Phycol. Res. 47:167–88.CrossRefGoogle Scholar
Chopin, T., Buschmann, A. H., Halling, C., et al. (2001). Integrating seaweeds into marine aquaculture systems: a key towards sustainability. J. Phycol. 37:975–86.CrossRefGoogle Scholar
Chopin, T., Robinson, S. M. C., Troell, M., et al. (2008). Multitrophic integration for sustainable marine aquaculture. In Jørgensen, S. E. and Fath, B. D. (eds), The Encyclopedia of Ecology, Ecological Engineering (Vol. 3, pp. 2463–75). Oxford: Elsevier.CrossRefGoogle Scholar
Chopin, T., Cooper, J. A., Reid, G., Cross, S. and Moore, C. (2012) Open-water integrated multi-trophic aquaculture: environmental biomitigation and economic diversification of fed aquaculture by extractive aquaculture. Rev. Aquaculture 4:209–20.CrossRefGoogle Scholar
Chow, F., and de Oliveira, M. C. (2008). Rapid and slow modulation of nitrate reductase activity in the red macroalga Gracilaria chilensis (Gracilariales, Rhodophyta): influence of different nitrogen sources. J. Appl. Phycol. 20:775–82.CrossRefGoogle Scholar
Christensen, T. (1988). Salinity preferences of twenty species of Vaucheria (Tribophyceae). J Mar Biol Assoc UK 68:531–45.CrossRefGoogle Scholar
Christie, A. O., and Shaw, M. (1968). Settlement experiments with zoospores of Enteromorpha intestinalis (L.) Link. Br. Phycol. Bull. 3:529–34.CrossRefGoogle Scholar
Christie, H., Norderhaug, K. M., and Fredriksen, S.. (2009). Macrophytes as habitat for fauna. Mar. Ecol. Prog. Ser. 396:221–33.CrossRefGoogle Scholar
Chu, Z. X., and Anderson, J. M. (1985). Isolation and characterization of a siphonoxanthin-chlorophyll a/b protein complex of photosystem I from a Codium species (Siphonales). Biochim Biophys Acta 806:154–60.CrossRefGoogle Scholar
Chung, I. K., and Brinkhuis, B. (1986). Copper effects in early stages of the kelp, Laminaria saccharina. Mar. Poll. Bull. 17:213–18.CrossRefGoogle Scholar
Chung, I. K., Beardall, J., Mehta, S., Sahoo, D., and Stojkovic, S. (2011). Using marine macroalgae for carbon sequestration: a critical appraisal. J. Appl. Phycol. 23:877–86.CrossRefGoogle Scholar
Clark, M. S., Tanguy, A., Jollivet, D. et al. (2010). Populations and pathways: genomic approaches to understanding population structure and environmental adaptation. In Cock, J. M., Tessmar–Raible, K., Boyen, C., and Viard, F. (eds), Introduction to Marine Genomics (pp. 73–118). Dordrecht, The Netherlands: Springer.CrossRefGoogle Scholar
Clark, R.B. (2001). Marine Pollution (5th edn). Oxford: Oxford University Press. 237 pp.Google Scholar
Claudet, J., Osenberg, C. W., Benedetti-Cecchi, L., et al. (2008). Marine reserves: size and age do matter. Ecol. Lett. 11:481–9.CrossRefGoogle ScholarPubMed
Clayton, M. N. (1981). Correlated studies on seasonal changes in the sexuality, growth rate and longevity of complanate Scytosiphon (Scytosiphonaceae, Phaeophyta) from southern Australia growing in situ. J. Exp. Mar. Biol. Ecol. 51:87–96.CrossRefGoogle Scholar
Clayton, M. N. (1984). An electron microscope study of gamete release and settling in the complanate form of Scytosiphon (Scytosiphonaceae, Phaeophyta). J. Phycol. 20:276–85.CrossRefGoogle Scholar
Clayton, M. N. (1988). Evolution and life histories of brown algae. Bot. Mar. 31:379–87.CrossRefGoogle Scholar
Clayton, M. N. (1992). Propagules of marine macroalgae: structure and development. Brit. Phycol. J. 27:219–32.CrossRefGoogle Scholar
Clayton, M. N., and Ashburner, C. M. (1990). The anatomy and ultrastructure of “conducting channels” in Ascoseira mirabilis (Ascoseirales, Phaeophyceae). Bot Mar 33: 63–70.CrossRefGoogle Scholar
Clayton, M. N., Kevekordes, K., Schoenwaelder, M. E. A., Schmid, C. E., and Ashburner, C. M. (1998). Parthenogenesis in Hormosira banksii (Fucales, Phaeophyceae). Bot. Mar. 41:23–30.CrossRefGoogle Scholar
Clements, K. D., Raubenheimer, D., and Choat, J. H. (2009). Nutritional ecology of marine herbivorous fishes: ten years on. Funct. Ecol. 23:79–92. [.]CrossRefGoogle Scholar
Clifton, K. E. (1997). Mass spawning by green algae on coral reefs. Science 275:1116–18.CrossRefGoogle ScholarPubMed
Clifton, K. E., and Clifton, L. M. (1999). The phenology of sexual reproduction by green algae (Bryopsidales) on Caribbean coral reefs. J. Phycol. 35:24–34.CrossRefGoogle Scholar
Cock, J. M., Sterck, L., Rouzé, P., et al. (2010a). The Ectocarpus genome and the independent evolution of multicellularity in brown algae. Nature 465:617–21. [, , , ]CrossRefGoogle ScholarPubMed
Cock, J.M., Tessmar-Raible, K., Boyen, C., and Viard, F. (eds) (2010b). Introduction to Marine Genomics, Advances in Marine Genomics 1. Dordrecht, The Netherlands: Springer.CrossRefGoogle Scholar
Cocquyt, E., Verbruggen, H., Leliaert, F., and De Clerck, O. (2010). Evolution and cytological diversification of the green seaweeds (Ulvophyceae). Mol. Biol. Evol. 27:2052–61.CrossRefGoogle Scholar
Coelho, L., Prince, J., and Nolen, T. G. (1998). Processing of defensive pigment in Aplysia californica: acquisition, modification and mobilization of the red algal pigment r-phycoerythrin by the digestive gland. J. Exp. Biol. 201:425–38.Google ScholarPubMed
Coehlo, S. M., Rijstenbil, J. W., and Brown, M. J. (2000). Impacts of anthropogenic stresses on the early development stages in seaweeds. J. Aquat. Ecosyst. Stress Recovery 7:317–33.Google Scholar
Coelho, S. M., Peters, A. F., Charrier, B., et al. (2007). Complex life cycles of multicellular eukaryotes: new approaches based on the use of model organisms. Gene 406:152–70.CrossRefGoogle ScholarPubMed
Coelho, S. M., Brownlee, C., and Bothwell, J. H. F..(2008). A tip-high, Ca2+-interdependent, reactive oxygen species gradient is associated with polarized growth in Fucus serratus zygotes. Planta 227:1037–46.CrossRefGoogle ScholarPubMed
Coelho, S. M., Heesch, S., Grimsley, N., Moreau, H., and Cock, M. J. (2010). Genomics of marine algae. In Cock, J. M., Tessmar-Raible, K., Boyen, C., and Viard, F. (eds), Introduction to Marine Genomics, Advances in Marine Genomics 1 (pp. 179–212). The Netherlands: Springer.CrossRefGoogle Scholar
Coelho, S. M., Godfroy, O., Arun, A., G, et al. (2011). OUROBOROS is a master regulator of the gametophyte to sporophyte life cycle transition in the brown alga Ectocarpus. PNAS 108:11, 518–23.CrossRefGoogle ScholarPubMed
Cohen, R. A., and Fong, P. (2005). Experimental evidence supports the use of δ15N content of the opportunistic green macroalga Enteromorpha intestinalis (Chlorophyta) to determine nitrogen sources to estuaries. J. Phycol. 41:149–56.CrossRefGoogle Scholar
Cohen-Bazire, G., and Bryant, D. A. (1982). Phycobilisomes: composition and structure. In Carr, N. G. and Whitton, B. A. (eds) The Biology of Cyanobacteria (pp. 143–190). Oxford: Blackwell Scientific.Google Scholar
Cole, D. G. (2003). The intraflagellar transport machinery of Chlamydomonas reinhardtii. Traffic 4:435–42.CrossRefGoogle ScholarPubMed
Coleman, M. A. (2002). Small-scale spatial variability in intertidal and subtidal turfing algal assemblages and the temporal generality of these patterns. J. Exp. Mar. Biol. Ecol. 267:53–74.CrossRefGoogle Scholar
Coleman, M. A., and Brawley, S. H. (2005). Are life history characteristics good predictors of genetic diversity and structure? A case study of the intertidal alga Fucus spiralis (Heterokontophyta; Phaeophyceae). J. Phycol. 41:753–62.CrossRefGoogle Scholar
Coleman, R. A., Underwood, A. J., Benedetti-Cecchi, L., et al. (2006). A continental scale evaluation of the role of limpet grazing on rocky shores. Community Ecology 147:556–64.Google ScholarPubMed
Coleman, R. A., Ramchunder, S. J., Moody, A. J., and Foggo, A. (2007). An enzyme in snail saliva induces herbivore-resistance in a marine alga. Funct. Ecol. 21:101–6.CrossRefGoogle Scholar
Collado-Vides, L. (2002a). Clonal architecture in marine macroalgae: ecological and evolutionary perspectives. Evol. Ecol. 15:531–45.CrossRefGoogle Scholar
Collado-Vides, L. (2002b). Morphological plasticity of Caulerpa prolifera (Caulerpales, Chlorophyta) in relation to growth form in a coral reef lagoon. Bot. Mar. 45:123–9.CrossRefGoogle Scholar
Collado-Vides, L., Rutten, L. M., and Fourqurean, J. W. (2005). Spatiotemporal variation of the abundance of calcareous green macroalgae in the Florida Keys: a study of synchrony within a macroalgal functional-form group. J. Phycol. 41:742–52.CrossRefGoogle Scholar
Collén, J. and Davison, I. R. (1999a). Reactive oxygen metabolism in intertidal Fucus spp. (Phaeophyceae). J Phycol 35:62–9. [, , , , , , ]CrossRefGoogle Scholar
Collén, J., and Davison, I. R. (1999b). Production and damage of reactive oxygen in intertidal Fucus spp. (Phaeophyceae). J Phycol 35: 54–61. [, , , , ]CrossRefGoogle Scholar
Collén, J., and Davison, I. R. (1999c). Stress tolerance and reactive oxygen metabolism in the intertidal red seaweeds Mastocarpus stellatus and Chondrus crispus. Plant Cell. Environ. 22:1143–51. [, , , ]CrossRefGoogle Scholar
Collén, J., and Davison, I. R. (2001). Seasonality and thermal acclimation of reactive oxygen metabolism in Fucus vesiculosus (Phaeophyceae). J. Phycol. 37:474–81.CrossRefGoogle Scholar
Collén, J., Del Río, M. J., García–Reina, G., and Pedersén, M. (1995). Photosynthetic H2O2 production by Ulva rigida. Planta 196:225–30.CrossRefGoogle Scholar
Collén, J., Roeder, V., Rousvoal, S., et al. (2006). An expressed sequence tag analysis of thallus and regenerating protoplasts of Chondrus crispus (Gigarinales, Rhodophyceae). J. Phycol. 42:104–12.CrossRefGoogle Scholar
Collén, J., Guisle–Marsollier, I., Léger, J., and Boyen, C. (2007). Response of the transcriptome of the intertidal red seaweed Chondrus crispus to controlled and natural stresses. New Phytol. 176:45–55. [, , ]CrossRefGoogle ScholarPubMed
Collén, J., Porcel, B., Carré, W., et al. (2013). Genome structure and metabolic features in the red seaweed Chondrus crispus shed light on evolution of the Archaeplastida. PNAS 110: 5247–52. [, , , ]CrossRefGoogle ScholarPubMed
Collins, C. J., Fraser, C. I., Ashcroft, A., and Waters, J. M. (2010). Asymmetric dispersal of southern bull-kelp (Durvillaea antarctica) adults in coastal New Zealand: testing an oceanographic hypothesis. Mol. Ecol. 19:4572–80.CrossRefGoogle ScholarPubMed
Collins, S., and Bell, G. (2004). Phenotypic consequences of 1,000 generations of selection at high CO2 in a green alga. Nature 431:566–69.CrossRefGoogle Scholar
Collos, Y., Siddiqi, M. Y., Yang, M. Y., Glass, A. D. M., and Harrison, P. J. (1992). Nitrate uptake kinetics by two marine diatoms using the radioactive tracer 13N. J. Exp. Mar. Biol. Ecol. 163:251–60.CrossRefGoogle Scholar
Colman, J. (1933). The nature of the intertidal zonation of plants and animals. J. Mar. Biol. Assoc. UK 18:435–76.CrossRefGoogle Scholar
Conklin, E. J., and Smith, J. E. (2005). Abundance and spread of the invasive red algae, Kappaphycus spp., in Kane’ohe Bay, Hawai’i and an experimental assessment of management options. Biological Invasions 7:1029–39.CrossRefGoogle Scholar
Connan, S., and Stengel, D. B. (2011). Impacts of ambient salinity and copper on brown algae: 1. Interactive effects on photosynthesis, growth and copper accumulation. Aquat. Toxicol. 104:94–107.CrossRefGoogle ScholarPubMed
Connell, J. H. (1961a). Effects of competition, predation by Thais lapillus, and other factors on natural populations of the barnacle Balanus balanoides. Ecol. Monogr. 31:61–104.CrossRefGoogle Scholar
Connell, J. H. (1961b). The influence of interspecific competition and other factors on the distribution of the barnacle Chthamalus stellatus. Ecology 42:710–23.CrossRefGoogle Scholar
Connell, J. H. (1978). Diversity in tropical rain forests and coral reefs. Science 199:1302–10.CrossRefGoogle ScholarPubMed
Connell, J. H., and Slatyer, R. O. (1977). Mechanisms of succession in natural communities and their role in community stability and organization. Am. Nat. 111:1119–44. [.]CrossRefGoogle Scholar
Connell, J. H., Hughes, T. P., and Wallace, C. C. (1997). A 30-year study of coral abundance, recruitment, and disturbance at several scales in space and time. Ecol. Monogr. 67:461–88.CrossRefGoogle Scholar
Connell, S. D. (2007a). Subtidal temperate rocky habitats: habitat heterogeneity at local to continental scales. In Connell, S. D. and Gillanders, B. M. (eds), Marine Ecology (pp. 378–401). Melbourne, Australia: Oxford University Press.Google Scholar
Connell, S. D. (2007b). Water quality and the loss of coral reefs and kelp forests: alternative states and the influence of fishing. In Connell, S. D. and Gillanders, B. M. (eds), Marine Ecology (pp. 556–68). Melbourne, Australia: Oxford University Press.Google Scholar
Connell, S. D., and Gillanders, B. M. (eds) (2007). Marine Ecology. Melbourne, Australia: Oxford University Press, 630 pp. [, , 4.0]Google Scholar
Connell, S. D., and Irving, A. D. (2009). The subtidal ecology of rocky coasts: local–regional biogeographic patterns and their experimental analysis. In Witman, J. D. and Roy, K. (eds), Marine Macroecology (pp. 392–417). Chicago, IL: University of Chicago Press. [, , , ]CrossRefGoogle Scholar
Connell, S. D., and Russell, B. D. (2010). The direct effects of increasing CO2 and temperature on non–calcifying organisms: increasing the potential for phase shifts in kelp forest. P. Roy. Soc. B Biol. Sci. 227:1409–15.CrossRefGoogle Scholar
Connolly, R. M., Guest, M. A., Melville, A. J., and Oakes, J. M. (2004). Sulfur stable isotopes separate producers in marine food–web analysis. Oecologia 138(2): 161–7.CrossRefGoogle ScholarPubMed
Connolly, S. R., and Roughgarden, J. (1999). Theory of marine communities: competition, predation, and recruitment-dependent interaction strength. Ecol. Monogr. 69:277–96.CrossRefGoogle Scholar
Contreras, L., Mella, D., Moenne, A., and Correa, J. A. (2009). Differential responses to copper-induced oxidative stress in the marine macroalgal Lessonia migrescens and Scytosiphon lomentaria (Phaeophyceae). Aquat. Tox. 94:94–102.CrossRefGoogle Scholar
Contreras-Porcia, L., Dennett, G., González, A., et al. (2011). Identification of copper-induced genes in the marine alga Ulva compressa (Chlorophyta). Mar. Biotechnol. 13:544–56.CrossRefGoogle Scholar
Cook, P. L. M., Butler, E. V. C., and Eyre, B. D. (2004). Carbon and nitrogen cycling on intertidal mudflats of a temperate Australian estuary. I. Benthic metabolism. Mar. Ecol. Prog. Ser. 280:25–38CrossRefGoogle Scholar
Coomans, R. J., and Hommersand, M. H. (1990). Vegetative growth and organization. In Cole, K. M. and Sheath, R. G. (eds), Biology of the Red Algae (pp. 275–304). Cambridge: Cambridge University Press.Google Scholar
Coon, D. A., Neushul, M., and Charters, A. C. (1972). The settling behavior of marine algal spores. Proc. Intl. Seaweed Symp. 7:237–42.Google Scholar
Cooper, T. F., Gilmour, J. P., and Fabricus, K. E. (2009). Bioindicators of changes in water quality on coral reefs: review and recommendations for monitoring programmes. Coral Reefs 28:589–606.[]CrossRefGoogle Scholar
Corbit, J. D., and Garbary, D. J. (1993). Computer-simulation of the morphology and development of several species of seaweed using Lindenmayer-systems. Comput. Graph. 17:85–8.CrossRefGoogle Scholar
Cornwall, C. E., Hepburn, C. D., Pilditch, C. A., and Hurd, C. L. (2013). Concentration boundary layers around complex assemblages of macroalgae: implications for the effects of ocean acidification on understory coralline algae. Limnol. Oceanogr. 58: 121–30.CrossRefGoogle Scholar
Correa, J. A., and Flores, V. (1995). Whitening, thallus decay and fragmentation in Gracilaria chilensis associated with an endophytic amoeba. J. Appl. Phycol. 7:421–5.CrossRefGoogle Scholar
Correa, J. A., and Martinez, E. A. (1996). Factors associated with host specificity in Sporocladopsis novae-zelandiae (Chlorophyta). J. Phycol. 32:22–7.CrossRefGoogle Scholar
Correa, J. A., and McLachlan, J. (1991). Endophytic algae of Chondrus crispus (Rhodophyta). III. Host specificity. J. Phycol. 27:448–59.CrossRefGoogle Scholar
Correa, J. A., and Mclachlan, J. L. (1994). Endophytic algae of Chondrus crispus (Rhodophyta). V. Fine structure of the infection by Acrochaete operculata (Chlorophyta). Eur. J. Phycol. 29:33–47.CrossRefGoogle Scholar
Correa, J., Novaczek, I., and McLachlan, J. (1986). Effect of temperature and daylength on morphogenesis of Scytosiphon lomentaria (Scytosiphonales, Phaeophyta) from eastern Canada. Phycologia 25: 469–75.CrossRefGoogle Scholar
Correa, J.A., González, P., Sánchez, P., Muñoz, J., and Orellana, M. C. (1996). Copper–algae interactions: inheritance or adaptation. Environ. Monitor. Assess. 40:41–54.CrossRefGoogle ScholarPubMed
Corzo, A., and Niell, F. X. (1991). Determination of nitrate reductase activity in Ulva rigida C. Agardh by the in situ method. J. Exp. Mar. Biol. Ecol. 146:181–91.CrossRefGoogle Scholar
Corzo, A., and Niell, F. X. (1994). Nitrate reductase activity and in vivo nitrate reduction rate in Ulva rigida illuminated by blue light. Mar. Biol. 120:17–23.Google Scholar
Cosgrove, D. J. (1981). Analysis of the dynamic and steady-state responses of growth rate and turgor pressure to changes in cell parameters. Plant Physiol. 68:1439–46.CrossRefGoogle ScholarPubMed
Cosse, A., LeBlanc, C., and Potin, P. (2008). Dynamic defense of marine macroalgae against pathogens: from early activated to gene-regulated responses. Adv. Bot. Res. 46:221–66.CrossRefGoogle Scholar
Costa, S., Crespo, D., Henriques, B. M. G.. et al. (2011). Kinetics of mercury accumulation and its effects on Ulva lactuca growth rate at two salinities and exposure conditions. Water Air Soil Poll. 217:689–99.CrossRefGoogle Scholar
Cotton, A. D. (1912). Marine algae. In Praeger, R. L. (ed.), A Biological Survey of Clare Island in the County of Mayo, Ireland and the Adjoining District. Section 1, Part 15 (pp. 1–178). Dublin, Ireland: Hodges, Figgis and Co., Ltd.Google Scholar
Coughlan, S. (1977). Sulphate uptake in Fucus serratus. 1. Exp. Bot. 28:1207–15.CrossRefGoogle Scholar
Coughlan, S., and Evans, L. V. (1978). Isolation and characterization of Golgi bodies from vegetative tissue of the brown algaFucus serratus. J. Exp. Bot. 29:55–68.CrossRefGoogle Scholar
Court, G. J. (1980). Photosynthesis and translocation studies of Laurencia spectabilis and its symbiont Janczewskia gardneri (Rhodophyceae). J Phycol. 16:270–9.CrossRefGoogle Scholar
Coutinho, R., and Yoneshigue, Y. (1990). Diurnal variation in photosynthesis vs. irradiance curves from “sun” and “shade” plants of Pterocladia capillacea (Gmelin) Bornet et Thuret (Gelidiaciaceae [sic]: Rhodophyta) from Cabo Frio, Rio de Janeiro, Brazil. J. Exp. Mar. Biol. Ecol. 118:217–28.CrossRefGoogle Scholar
Coyer, J. A., Peters, A. F., Hoarau, G., Stam, W. T., and Olsen, J. L. (2002). Hybridization of the marine seaweeds, Fucus serratus and Fucus evanescens (Heterokontophyta: Phaeophyceae) in a 100-year-old zone of secondary contact. P. Roy. Soc. Lond. B. Biol. Sci. 269:1829–34.CrossRefGoogle Scholar
Coyer, J. A., Hoarau, G., Oudot–Le Secq, M.-P., Stam, W. T. and Olsen, J. L. (2006a). A mtDNA-based phylogeny of the brown algal genus Fucus (Heterokontophyta; Phaeophyta). Mol. Phylogenet. Evol. 39:209–22.CrossRefGoogle Scholar
Coyer, J. A., Hoarau, G., Pearson, G. A., et al. (2006b). Convergent adaptation to a marginal habitat by homoploid hybrids and polyploidy ecads in the seaweed genus Fucus. Biol. Lett. 2:405–8.CrossRefGoogle ScholarPubMed
Craigie, J. S. (1974). Storage products. In Stewart, W. D. P. (ed.) Algal Physiology and Biochemistry (pp. 206–35). Oxford: Blackwell Scientific.Google Scholar
Craigie, J. S. (1990). Cell walls. In Cole, K. M. and Sheath, R. G. (eds), Biology of the Red Algae (pp. 221–57). Cambridge: Cambridge University Pres.Google Scholar
Craigie, J. S. (2011). Seaweed extract stimuli in plant science and agriculture. J. Appl. Phycol. 23:371–93.CrossRefGoogle Scholar
Craigie, J. S., and Shacklock, P. F. (1995). Culture of Irish moss. In Boghen, A. D. (ed.), Cold-water Aquaculture in Atlantic Canada (pp. 363–90). Canadian Institute for Research and Regional Development, Moncton, NB, Canada.Google Scholar
Craigie, J. S., Morris, E. R., Rees, D. A., and Thorn, D. (1984). Alginate block structure in Phaeophyceae from Nova Scotia: variation with species, environment and tissue-type. Carbohydro Polymers 4:237–52.CrossRefGoogle Scholar
Craigie, J. S., Correa, J. A., and Gordon, M. E. (1992). Cuticles from Chondrus crispus (Rhodophyta). J. Phycol. 28:777–86.CrossRefGoogle Scholar
Crawford, N. M., and Glass, A. D. M. (1998). Molecular and physiological aspects of nitrate uptake in plants. Trends Plant Sci. 3:389–95.CrossRefGoogle Scholar
Crawley, K. R., and Hyndes, G. A. (2007). The role of different types of detached macrophytes in the food and habitat choice of a surf-zone inhabiting amphipod. Mar. Biol. 151(4):1433–43.CrossRefGoogle Scholar
Crawley, K. R., Hyndes, G. A., and Ayvazian, S. G. (2006). Influence of different volumes and types of detached macrophytes on fish community structure in surf zones of sandy beaches. Mar. Ecol. Prog. Ser. 307:233–46.CrossRefGoogle Scholar
Crawley, K.R., Hyndes, G. A., Vanderklift, M. A., Revill, A. T., and Nichols, P. D. (2009). Allochthonous brown algae are the primary food source for consumers in a temperate, coastal environment. Mar. Ecol. Prog. Ser. 376:33–44.CrossRefGoogle Scholar
Crayton, M. A., Wilson, E., and Quatrano, R. S. (1974). Sulfation of fucoidan in Fucus embryos. II. Separation from initiation of polar growth. Devel. Biol. 39:134–7.CrossRefGoogle ScholarPubMed
Creed, J. C., Norton, T. A., and Harding, S. P. (1996). The development of size structure in a young Fucus serratus population. Eur. J. Phycol. 31:203–9.CrossRefGoogle Scholar
Creed, J. C., Norton, T. A., and Kain (Jones), J. M. (1997). Intraspecific competition in Fucus serratus germlings: the interaction of light, nutrients and density. J. Exp. Mar. Biol. Ecol. 212:211–23.CrossRefGoogle Scholar
Creed, J. C., Kain (Jones), J. M., and Norton, T. (1998). An experimental evaluation of density and plant size in two large brown seaweeds. J. Phycol. 34(1):39–52.CrossRefGoogle Scholar
Critchley, A. T., and Ohno, M. (eds) (1998). Seaweed Resources of the World. Yokosuka, Japan: Japan International Cooperation Agency, 429 pp.Google Scholar
Critchley, A. T., Ohno, M., and Largo, D. B. (2006). World Seaweed Resources: An Authoritative Reference System. DVD–ROM. Wokingham, UK: ETI Information Services.Google Scholar
Cronin, G., and Hay, M. E. (1996a). Susceptibility to herbivores depends on recent history of both the plant and the animal. Ecology 77:1531–43.CrossRefGoogle Scholar
Cronin, G., and Hay, M. E. (1996b). Induction of seaweed chemical defenses by amphipod grazing. Ecology 77:2287–301.CrossRefGoogle Scholar
Crothers, J. H. (1983). Field experiments on the effects of crude oil and dispersant on the common animals and plants of rocky seashores. Mar. Environ. Res. 8:215–39.CrossRefGoogle Scholar
Crowe, T.P., Thompson, R. C., Bray, S., and Hawkins, S. J. (2000). Impacts of anthropogenic stress on rocky intertidal communities. J. Aquat. Ecosy. Stress Recov. 7:273–97.CrossRefGoogle Scholar
Cruces, E., Huovinen, P., and Gómez, I. (2012). Phlorotannin and antioxidant responses upon short-term exposure to UV–radiation and elevated temperature in three South Pacific kelps. Photochem. Photobiol. 88:58–66.CrossRefGoogle ScholarPubMed
Cruz-Rivera, E., and Hay, M. E. (2000). Can quantity replace quality? Food choice, compensatory feeding, and fitness of marine mesograzers. Ecology 81:201–19.CrossRefGoogle Scholar
Cunningham, E. M., Guiry, M. D., and Breeman, A. M. (1993). Environmental regulation of development, life history and biogeography of Helminthora stackhousei (Rhodophyta) by daylength and temperature. J. Exp. Mar. Biol. 171:1–21.CrossRefGoogle Scholar
Dalby, D. H. (1980). Monitoring and exposure scales. In Price, J. H., Irvine, D. E. G., and Farnham, W. F. (eds), The Shore Environment. Vol. 1: Methods (pp. 117–36). New York: Academic Press.Google Scholar
Daly, M. A., and Mathieson, A. C (1977). The effects of sand movement on intertidal seaweeds and selected invertebrates at Bound Rock, New Hampshire, USA. Mar. Biol. 43:45–55.CrossRefGoogle Scholar
Davenport, A. C., and Anderson, T. W. (2007). Positive indirect effects of reef fishes on kelp performance: the importance of mesograzers. Ecology 88:1548–61.CrossRefGoogle ScholarPubMed
Davis, T. A., Volesky, B., and Mucci, A. (2003). A review of the biochemistry of heavy metal biosorption by brown algae. Water Res. 37:4311–30.CrossRefGoogle ScholarPubMed
Davison, I. R. (1991). Environmental effects on algal photosynthesis: temperature. J. Phycol. 27:2–8.CrossRefGoogle Scholar
Davison, I. R., and Davison, J. O. (1987). The effect of growth temperature on enzyme activities in the brown alga Laminaria saccharina. Br. Phycol. J. 22:77–87.CrossRefGoogle Scholar
Davison, I. R., and Pearson, G. A. (1996). Stress tolerance in intertidal seaweeds. J. Phycol. 32:197–222. [, , ]CrossRefGoogle Scholar
Davison, I. R., and Reed, R. H. (l985). Osmotic adjustment in Laminaria digitata (Phaeophyta) with particular reference to seasonal changes in internal solute concentrations. J. Phycol. 21:41–50.CrossRefGoogle Scholar
Davison, I. R., and Stewart, W. D. P. (1983). Occurrence and significance of nitrogen transport in the brown alga Laminaria digitata. Mar. Biol. 77:107–12.CrossRefGoogle Scholar
Davison, I. R., and Stewart, W. D. P. (1984). Studies on nitrate reductase activity in Laminaria digitata (Huds.) Lamour. I. Longitudinal and transverse profiles of nitrate reductase activity within the thallus. J. Exp. Mar. Biol. Ecol. 74:201–10.CrossRefGoogle Scholar
Davison, I. R., Andrews, M., and Stewart, W. D. P. (1984). Regulation of growth in Laminaria digitata: use of in vivo nitrate reductase activities as an indicator of nitrogen limitation in field populations of Laminaria spp. Mar. Biol. 84:207–17.CrossRefGoogle Scholar
Davison, I. R., Dudgeon, S. R., and Ruan, H–M. (1989). Effect of freezing on seaweed photosynthesis. Mar Ecol Prog Ser 58: 123–31.CrossRefGoogle Scholar
Davison, I. R., Greene, R. M., and Podolak, E. J. (1991). Temperature acclimation of respiration and photosynthesis in the brown alga Laminaria saccharina. Mar Biol 110: 449–54.CrossRefGoogle Scholar
Dawes, C. J. (1979). Physiological and biochemical comparisons of Eucheuma spp. (F1orideophyceae) yielding iotacarrageenan. Proc. Intl. Seaweed Symp. 9:188–207.Google Scholar
Dawes, C. J. (1987). The biology of commercially important tropical marine algae. In Bird, K. and Benson, P. H. (eds), Seaweed Cultivation for Renewable Resources (pp. 155–90). Amsterdam: Elsevier.Google Scholar
Dawes, C. J. (1989). Temperature acclimation in cultured Eucheuma isiforme from Florida and E. alvarezii from the Phillipines. J. Appl. Phycol. 1: 59–65.CrossRefGoogle Scholar
Dawes, C. J., and Barilotti, C. (1969). Cytoplasmic organization and rhythmic streaming in growing blades of Caulerpa prolifera. Am. J. Bot. 56:8–15.CrossRefGoogle Scholar
Dawes, C. J., and McIntosh, R. P. (1981). The effect of organic material and inorganic ions on the photosynthetic rate of the red alga Bostrychia binderi from a Florida estuary. Mar Biol 64:213–18.CrossRefGoogle Scholar
Dawes, C. J., Lawrence, J. M., Cheney, D. P., and Mathieson, A. C. (1974). Ecological studies of Floridian Eucheuma (Rhodophyta, Gigartinales). Ill. Seasonal variation of carrageenan, total carbohydrate, protein and lipid. Bull. Mar. Sci. 24:286–99.Google Scholar
Dawes, C. J., Stanley, N. F., and Stanicoff, O. J. (1977) Seasonal and reproductive aspects of plant chemistry, and iota-carrageenan from Florida Eucheuma (Rhodophyta, Gigartinales). Bot Mar 20: 137–47.CrossRefGoogle Scholar
Dawson, E. Y. (1950). A giant new Codium from Pacific Baja California. B. Torrey Bot. Club 77:298–300.CrossRefGoogle Scholar
Dawson, E. Y. (1951). A further study of upwelling and associated vegetation along Pacific Baja California, Mexico. J. Mar. Res. 10:39–58.Google Scholar
Dayton, P. K. (1971). Competition, disturbance, and community organization: the provision and subsequent utilization of space in a rocky intertidal community. Ecolog. Monogr. 41:351–89.CrossRefGoogle Scholar
Dayton, P K. (1973). Dispersion, dispersal, and persistence of the annual intertidal alga, Postelsia palmaeformis Ruprecht. Ecology 54:433–8.CrossRefGoogle Scholar
Dayton, P. K. (1975). Experimental evaluation of ecological dominance in a rocky intertidal algal community. Ecol. Monogr. 45:137–59.CrossRefGoogle Scholar
Dayton, P. K. (1985). The structure and regulation of some South American kelp communities. Ecol. Monogr. 55:447–68.CrossRefGoogle Scholar
Dayton, P. K., and Tegner, M. J. (1984a). The importance of scale in community ecology: a kelp forest example with terrestrial analogs. In Price, P. W., Slobodchikoff, C. M., and Gaud, W. S. (eds), A New Ecology: Novel Approaches to Interactive Systems (pp. 457–81). New York: John Wiley and Sons.Google Scholar
Dayton, P. K., and Tegner, M. J. (1984b). Catastrophic storms, El Niño and patch stability in southern California kelp community. Science 224:283–5.CrossRefGoogle ScholarPubMed
De Angelis, L., Rise, P., Giavarini, F., et al. (2005). Marine macroalgae analyzed by mass spectrometry are rich sources of polyunsaturated fatty acids. J.f Mass Spect. 40(12):1605–8CrossRefGoogle ScholarPubMed
De Beer, D., and Larkum, A. W. D. (2001). Photosynthesis and calcification in the calcifying algae Halimeda discoidea studied with microsensors. Plant Cell Environ. 24:1209–17. [, , ]CrossRefGoogle Scholar
De Burgh, M. E., and Fankboner, P. V. (1978). A nutritional association between the bull kelp Nereocystis luetkeana and its epizoic bryozoan Membranipora membranacea. Oikos 31:69–72.CrossRefGoogle Scholar
de Lestang–Bremond, G., and Quillet, M. (1981). The turnover of sulphates on the lambda-carrageenan of the cell-walls of the red seaweed Gigartinale: Catenella opuntia (Grev.). Proc. Intl Seaweed Symp. 10:449–54.Google Scholar
de los Santos, C. B., Pérez-Lloréns, J. L., and Vergara, J. J. (2009). Photosynthesis and growth in macroalgae: linking functional-form and power-scaling approaches. Mar. Ecol. Prog. Ser. 377:113–22.CrossRefGoogle Scholar
De Martino, A., Douady, D., Quinet–Szely, M., et al. (2000). The light-harvesting antenna of brown algae. Eur. J. Biochem. 267:5540–9.CrossRefGoogle ScholarPubMed
de Nys, R., Jameson, P. E., Chin, N., et al. (1990). The cytokinins as endogenous growth-regulators in Macrocystis pyrifera (L) C. Agardh (Phaeophyceae). Bot. Mar. 33:467–75.CrossRefGoogle Scholar
de Nys, R., Jameson, P. E., and Brown, M. T., (1991). The influence of cytokinins on the growth of Macrocystis pyrifera. Bot. Mar. 34:465–7.CrossRefGoogle Scholar
de Nys, R., Steinberg, P. D., Willemsen, P., et al. (1995). Broad-spectrum effects of secondary metabolites from the red alga Delisea pulchra in antifouling assays. Biofouling 8:259–71.CrossRefGoogle Scholar
de Ruyter van Steveninck, E. D., and Bak, R. P. M. (1986). Changes in abundance of coral reef bottom components related to mass mortaility of the sea urchin Diadema antillarum. Mar. Ecol. Prog. Ser. 34:87–94.CrossRefGoogle Scholar
de Wit, C. T. (1960). On competition. Versl. Landbouwkd Onderz (Agric. Res. Rep.) 66:1–82.Google Scholar
De Wreede, R. E. (1985). Destructive (harvest) sampling. In Littler, M. M. and Littler, D. S. (eds), Handbook of Phycological Methods: Ecological Field Methods: Macroalgae (pp. 147–60). Cambridge: Cambridge University Press.Google Scholar
De Wreede, R. E., and Klinger, T. (1988). Reproductive strategies in algae. In Lovett–Doust, J. and Lovett–Doust, L. (eds), Plant Reproductive Ecology: Patterns and Strategies (pp. 267–84). Oxford: Oxford University Press.Google Scholar
De Wreede, R E., Scrosati, R., and Servière–Zaragoza, E. (2000). Ramet dynamics for the clonal seaweed Pterocladiella capillacea (Rhodophyta): a comparison with Chondrus crispus and with Mazzaella cornucopiae (Gigartinales). J. Phycol. 36:1061–8Google Scholar
Deal, M. S., Hay, M. E., Wilson, D., and Fenical, W. (2003). Galactolipids rather than phlorotannins as herbivore deterrents in the brown seaweed Fucus vesiculosus. Oecologia 136:107–14.CrossRefGoogle ScholarPubMed
Dean, T. A., and Jacobsen, F. R. (1986). Nutrient-limited growth of juvenile kelp, Macrocystis pyrifera, during the 1982–1983 “El Niño” in southern California (USA). Mar. Biol. 90:597–602.CrossRefGoogle Scholar
DeBoer, J. A. (1981). Nutrients. In Lobban, C. S. and Wynne, M. J. (eds), The Biology of Seaweeds (pp. 356–91). Oxford: Blackwell Scientific. [, , , , , , , , , ]Google Scholar
DeBoer, J. A., and Whoriskey, F. G. (1983). Production and role of hyaline hairs of Ceramium rubrum. Mar. Biol. 77:229–34.CrossRefGoogle Scholar
DeBoer, J. A., Guigli, H.J., Israel, T. L., and D’Elia, C. F. (1978). Nutritional studies of two red algae. I. Growth rate as a function of nitrogen source and concentration. J. Phycol. 14:261–6. [, , , , ]CrossRefGoogle Scholar
DeCew, T. C., and West, J. A. (1982). A sexual life history in Rhodophysema (Rhodophyceae): a re-interpretation. Phycologia 21:67–74.CrossRefGoogle Scholar
Deckert, R. J., and Garbary, D. J. (2005). Ascophyllum and its symbionts. VI. Microscopic characterization of the Ascophyllum nodosum (Phaeophyceae), Mycophycias ascophylli (Ascomycetes) symbiotum. Algae 20:225–32.CrossRefGoogle Scholar
Deegan, L. A., Wright, A., Ayvazian, S. G., et al. (2002). Nitrogen loading alters seagrass ecosystem structure and support of higher trophic levels. Aquat. Conserv. 12(2):193–212CrossRefGoogle Scholar
Delage, L., Leblanc, C., Collén, P. N., et al. (2011). In silico survey of the mitochondrial protein uptake and maturation systems in the brown alga Ectocarpus siliculosus. PloS ONE 6:e19540. .CrossRefGoogle ScholarPubMed
de León–Chavira, F., Huerta–Diaz, M. A., and Chee–Barragán, A. (2003). New methodology for extraction of total metal from macroalgae and its application to selected samples collected in pristine zones from Baja California, Mexico. Bull. Environ. Contam. Toxicol. 70:809–16.CrossRefGoogle ScholarPubMed
Delgado, O., and Lapointe, B. E. (1994) Nutrient-limited productivity of calcareous versus fleshy macroalgae in a eutrophic, carbonate-rich tropical marine environment. Coral Reefs 13:151–9.CrossRefGoogle Scholar
D’Elia, C. F., and DeBoer, J. A. (1978). Nutritional studies of two red algae. II. Kinetics of ammonium and nitrate uptake. J. Phycol. 14:266–72. [, , ]CrossRefGoogle Scholar
Demes, K. W., Graham, M. H., and Suskiewicz, T. S. (2009a). Phenotypic plasticity reconciles incongruous molecular and morphological taxonomies: the giant kelp, Macrocystis (Laminariales, Phaeophyceae), is a monospecific genus. J. Phycol. 45:1266–9.CrossRefGoogle Scholar
Demes, K. W., Bell, S. S., and Dawes, C. J. (2009b). The effects of phosphate on the biomineralization of the green alga, Halimeda incrassata (Ellis) Lam. J. Exp. Mar. Biol. Ecol. 374:123–7.CrossRefGoogle Scholar
Demes, K. W., Carrington, E., Gosline, J., and Martone, P. T. (2011). Variation in anatomical and material properties explains differences in hydrodynamic performances of foliose red macroalgae (Rhodophyta). J. Phycol. 47:1360–7.CrossRefGoogle Scholar
Demmig-Adams, B., and Adams, W. W. (1992). Photoprotection and other responses of plants to high light stress. Annu. Rev. Plant Physiol. Plant Mol. Biol. 43:599–626.CrossRefGoogle Scholar
Demmig–Adams, B., Adams, W. W., and Mattoo, A. K. (2008). Photoprotection, photoinhibition, gene regulation, and environment. Advances in Photosynthesis and Respiration, Volume 21. Dordrecht, The Netherlands: Springer.Google Scholar
den Hartog, C. (1972). Substratum. Multicellular plants. In Kinne, O. (ed.), Marine Ecology (vol. I pt. 3, pp. 1277–89). New York: Wiley.Google Scholar
Deniaud, E., Fleurence, J., and Lahaye, M. (2003). Preparation and chemical characterization of cell wall fractions enriched in structural proteins from Palmaria palmate (Rhodophyta). Bot. Mar. 46:366–77.CrossRefGoogle Scholar
DeNiro, M. J., and Epstein, S. (1978). Influence of diet on the distribution of carbon isotopes in animals. Geochimica Cosmochim. Ac. 42:495–506CrossRefGoogle Scholar
Denley, E. J., and Dayton, P K. (1985). Competition among macroalgae. In Littler, M. M. and Littler, D. S. (eds), Handbook of Phycological Methods: Ecological Field Methods: Macroalgae (pp. 511–30). Cambridge: Cambridge University Press.Google Scholar
Denny, M. W. (1982). Forces on intertidal organisms due to breaking ocean waves: design and application of a telemetry system. Limnol. Oceanogr. 27:178–83.CrossRefGoogle Scholar
Denny, M. W. (1988). Biology and the Mechanics of the Wave Swept Environment. Princeton, NJ: Princeton University Press. [, , 8.0, , , ]CrossRefGoogle Scholar
Denny, M. W. (1993). Air and Water: The Biology and Physics of Life’s Media. Princeton, NJ: Princeton University Press, 341 pp. [8.0, ]Google Scholar
Denny, M. W. (2006). Ocean waves, nearshore ecology, and natural selection. Aquat. Ecol. 40:439–61. [8.0, , ]CrossRefGoogle Scholar
Denny, M. W., and Gaylord, B. (2010). Marine ecomechanics. Annu. Rev. Mar. Sci. 2:89–114. [8.0]CrossRefGoogle ScholarPubMed
Denny, M., and Helmuth, B. (2009). Confronting the physiological bottleneck: a challenge from ecomechanics. Integr. Comp. Biol. 49:197–201. [8.0]CrossRefGoogle ScholarPubMed
Denny, M., and Wethey, D. (2001). Physical processes that generate patterns in marine communities. In Bertness, M. D., Gaines, S. D., and Hay, M. E. (eds), Marine Community Ecology (pp. 3–37). Sunderland, MA: Sinauer Assocs. [, 8.0, , , , ]Google Scholar
Denny, M. W., Daniel, T. L., and Koehl, M. A. R. (1985). Mechanical limits to size in wave-swept organisms. Ecol. Monogr. 55:69–102.CrossRefGoogle Scholar
Denny, M. [W.], Brown, V., Carrington, E., Kraemer, G., and Miller, A. (1989). Fracture mechanics and the survival of waveswept macroalgae. J. Exp. Mar. Biol. Ecol. 127:211–28.CrossRefGoogle Scholar
Denny, M. W., Gaylord, B. P., and Cowen, E. A. (1997). Flow and flexibility. II. The roles of size and shape in determining wave forces on the bull kelp Nereocystis luetkeana. J. Exp. Biol. 200:3165–83.Google ScholarPubMed
Derenbach, J. B., and Gerneck, M. V. (1980). Interference of petroleum hydrocarbons with the sex pheromone reaction of Fucus vesiculosus (L.)J. Exp. Mar. Bioi. Ecol. 44:61–5.CrossRefGoogle Scholar
Destombe, C., and Oppliger, L. V. (2011). Male gametophyte fragmentation in Laminaria digitata: a life history strategy to enhance reproductive success. Cah. Biol. Mar. 52 (4SI):385–94.Google Scholar
Dethier, M. N. (1982). Pattern and process in tidepool algae: factors influencing seasonality and distribution. Bot. Mar. 25:55–66. [,4, ]CrossRefGoogle Scholar
Dethier, M. N. (1984). Disturbance and recovery in intertidal pools: maintenance of mosaic patterns. Ecol. Monogr. 54:99–118.CrossRefGoogle Scholar
Dethier, M. N., and Steneck, R. S. (2001). Growth and persistence of diverse intertidal crusts: survival of the slow in a fast-paced world. Mar. Ecol. Prog. Ser. 223:89–100.CrossRefGoogle Scholar
Dethier, M. N., and Williams, S. L. (2009). Seasonal stresses shift optimal intertidal algal habitats. Mar. Biol. 156:555–67.CrossRefGoogle Scholar
Devinny, J. S., and Volse, L. A. (1978). Effects of sediments on the development of Macrocystis pyrifera gametophytes. Mar. Biol. 48:343–8.CrossRefGoogle Scholar
Deysher, L., and Norton, T. A. (1982). Dispersal and colonization in Sargassum muticum (Yendo) Fensholt. J. Exp. Mar. Biol. Ecol. 56:179–95.CrossRefGoogle Scholar
Diamond, J. (1986). Overview: laboratory experiments, field experiments, and natural experiments. In Diamond, J. and Case, T. J., (ed.), Community Ecology (pp. 3–22). New York, NY: Harper and Row Publishers, Inc.Google Scholar
Dias, P. F., Siqueira, J. M., Vendruscolo, L. F., et al. (2005). Antiangiogenic and antitumoral properties of a polysaccharide isolated from the seaweed Sargassum stenophyllum. Cancer Chemother. Pharmacol. 56: 436–46.CrossRefGoogle ScholarPubMed
Diaz–Pulido, G., and McCook, L. J. (2002). The fate of bleached corals: patterns and dynamics of algal recruitment. Mar. Ecol. Prog. Ser. 232:115–28.CrossRefGoogle Scholar
Diaz-Pulido, G., and McCook, L. J. (2004). Effects of live coral, epilithic algal communities and substrate type on algal recruitment. Coral Reefs 23:225–33.CrossRefGoogle Scholar
Diaz–Pulido, G., Villamil, L., and Almanza, V. (2007a). Herbivory effects on the morphology of the brown alga Padina boergesenii (Phaeophyta). Phycologia 46:131–6.CrossRefGoogle Scholar
Diaz–Pulido, G., McCook, L. J., Larkum, A. W. D., et al. (2007b). Chapter 7: Vulnerability of macroalgae of the Great Barrier Reef to climate change. In Johnson, J. E. and Marshall, P. A., (eds), Climate Change and the Great Barrier Reef (pp. 153–92) Great Barrier Reef Marine Park Authority and Australian Greenhouse Office, Townsville.Google Scholar
Diaz–Pulido, G., Anthony, K. N. R., Kline, D. L., Dove, S., and Hoegh–Guldberg, O. (2012). Interactions between ocean acidification and warming on the mortality and dissolution of coralline algae. J. Phycol. 48:32–9.CrossRefGoogle ScholarPubMed
Dickson, L. G., and Waaland, J. R. (1985). Porphyra nerepcystis: a dual-daylength seaweed. Planta 165:548–53.CrossRefGoogle ScholarPubMed
Dieckmann, G. S. (1980). Aspects of the ecology of Laminaria pallida (Grev.) 1. Ag. off the Cape Pennisula (South Africa). I. Seasonal growth. Bot. Mar. 23: 579–85.Google Scholar
Dierssen, H. M., Zimmerman, R. C., Drake, L. A., and Burdige, D. (2010). Benthic ecology from space: optics and net primary production in seagrass and benthic algae across the Great Bahama Bank. Mar. Ecol. Prog. Ser. 411:1–15.CrossRefGoogle Scholar
Digby, P. S. B. (1977). Photosynthesis and respiration in the coralline algae Clathromorphum circumscriptum and Corallina officinalis and the metabolic basis of calcification. J. Mar. Biol. Assoc. UK 57:1111–24.CrossRefGoogle Scholar
Dillon, P S., Maki, J. S., and Mitchell, R. (1989). Adhesion of Enteromorpha swarmers to microbial films. Micro. Ecol. 17:39–47.CrossRefGoogle Scholar
Ding, H., and Ma, J. (2005). Simultaneous infection by red rot and chytrid diseases in Porphyra yezoensis Ueda. J. Appl. Phycol. 17:51–6.CrossRefGoogle Scholar
Dittami, S. M., Proux, C., Rousvoal, S., et al. (2011). Microarray estimation of genomic inter-strain variability in the genus Ectocarpus (Phaeophyceae). BMC Mol. Biol. 12:1–12.CrossRefGoogle Scholar
Dixon, P S., and Richardson, W. N. (1970). Growth and reproduction in red algae in relation to light and dark cycles. Ann. N.Y. Acad. Sci. 175:764–77.CrossRefGoogle Scholar
Doblin, M. S., Kurek, I., Jacob–Wilk, D., and Delmer, D. P. (2002). Cellulose biosynthesis in plants: from genes to rosettes. Plant Cell Physiol. 43:1407–20.CrossRefGoogle ScholarPubMed
Done, T. J. (1992). Phase shifts in coral reef communities and their ecological significance. Hydrobiologia 247:121–32.CrossRefGoogle Scholar
Done, T. J., Dayton, P. K., Dayton, A. E., and Steger, R. (1996). Regional and local variability in recovery of shallow coral communities: Moorea, French Polynesia and central Great Barrier Reef. Coral Reefs 9:183–92.CrossRefGoogle Scholar
Doney, S.C., Fabry, V. J., Feely, R.A., and Kleypas, J. A. (2009). Ocean acidification. The other CO2 problem. Ann. Rev. Mar. Sci. 1:169–92.CrossRefGoogle ScholarPubMed
Doropoulos, C., Hyndes, G. A., Lavery, P. S., and Tuya, F. (2009). Dietary preferences of two seagrass inhabiting gastropods: Allochthonous vs. autochthonous resources. Estuar. Coast. Shelf S. 83(1):13–18.CrossRefGoogle Scholar
Dortch, Q. (1982). Effect of growth conditions on accumulation of internal nitrate, ammonium, and protein in three marine diatoms. J. Exp. Mar. Biol. Ecol. 61:243–64.CrossRefGoogle Scholar
Dortch, Q. (1990). The interaction between ammonium and nitrate uptake in phytoplankton. Mar. Ecol. Prog. Ser. 61:183–201.CrossRefGoogle Scholar
Doty, M. S. (1946). Critical tide factors that are correlated with the vertical distribution of marine algae and other organisms along the Pacific Coast. Ecology 27:315–28.CrossRefGoogle Scholar
Doty, M. S. (1971). Measurement of water movement in reference to benthic algal growth. Bot. Mar. 14:32–5.CrossRefGoogle Scholar
Drechsler, Z., Sharkia, R., Cabantchik, Z. L., and Beer, S. (1994). The relationship of arginine groups to photosynthetic HCO3- uptake in Ulva sp. mediated by a putative anion exchanger. Planta 194:250–5.CrossRefGoogle Scholar
Drew, E. A. (1977). The physiology of photosynthesis and respiration in some Antarctic marine algae. Br. Antarct. Survey Bull. 46:59–76.Google Scholar
Drew, E. A. (1983). Light. In Earll, R. and Erwin, D. G. (eds), Sublittoral Ecology. The Ecology of the Shallow Sublittoral Benthos (pp. 10–57). Oxford: Clarendon Press.Google Scholar
Drew, E. A., and Abel, K. M. (1990). Studies on Halimeda. III. A daily cycle of chloroplast migration within segments. Bot. Mar. 33:31–45.CrossRefGoogle Scholar
Drew, E. A., and Hastings, R. M. (1992). A year-round ecophysiological study of Himantothallus grandifolius (Desmarestiales, Phaeophyta) at Signy Island, Antarctica. Phycologia 31:262–77.CrossRefGoogle Scholar
Drew, K. M. (1949). Conchocelis phase in the life history of Porphyra umbilicalis (L.) Kiitz. Nature 164:748 [, 10.21]CrossRefGoogle Scholar
Dring, M. J. (1967). Phytochrome in red alga, Porphyra tenera. Nature 215:1411–12.CrossRefGoogle Scholar
Dring, M. J. (1974). Reproduction. In Stewart, W. D. P. (ed.), Algal Physiology and Biochemistry (pp. 814–37). Oxford: Blackwell Scientific.Google Scholar
Dring, M. J. (1981). Photosynthesis and development of marine macrophytes in natural light spectra. In Smith, H. (ed), Plants and the Daylight Spectrum (pp. 297–314). London: Academic Press.Google Scholar
Dring, M. J. (1982). The Biology of Marine Plants. London: Arnold.Google Scholar
Dring, M. J. (1984a). Photoperiodism and phycology. Prog. Phycol. Res. 3:159–92.Google Scholar
Dring, M. J. (1984b). Blue light effects in marine macroalgae. In Senger, H. (ed.), Blue Light Effects in Biological Systems (pp. 509–16). Berlin: Springer-Verlag.CrossRefGoogle Scholar
Dring, M. J. (1987). Light climate in intertidal and subtidal zones in relation to photosynthesis and growth of benthic algae: a theoretical model. In Crawford, R. M. M. (ed.), Plant Life in Aquatic and Amphibious Habitats (pp. 23–34). Oxford: Blackwell Scientific.Google Scholar
Dring, M. J. (1988). Photocontrol of development in algae. Annu. Rev. Plant Physiol. Plant Molec. Biol. 39:157–74.CrossRefGoogle Scholar
Dring, M. J. (1989). Stimulation of light-saturated photosynthesis in Laminaria (Phaeophyta) by blue light. J. Phycol. 25:254–8.CrossRefGoogle Scholar
Dring, M. J. (1990). Light harvesting and pigment composition in marine phytoplankton and macroalgae. In Herring, P J., Campbell, A. K., Whitfield, M., and Maddock, L. (eds), Light and Life in the Sea (pp. 89–103). Cambridge: Cambridge University Press.Google Scholar
Dring, M. J. (2005). Stress resistance and disease resistance in seaweeds: the role of reactive oxygen metabolism. Adv. Bot. Res. 43:175–207. [, , ]CrossRefGoogle Scholar
Dring, M. J., and Brown, F. A. (1982). Photosynthesis of intertidal brown algae during and after periods of emersion: a renewed search for physiological causes of zonation. Mar. Ecol. Prog. Ser. 8:301–8.CrossRefGoogle Scholar
Dring, M. J., and Lüning, K. (1975). A photoperiodic response mediated by blue light in the brown a1gaScytosiphon lomentaria. Planta 125:25–32.Google Scholar
Dring, M. J., and Lüning, K. (1983). Photomorphogenesis in marine macroalgae. In Shropshire, Jr. W., and Mohr, H. (eds), Encyclopaedia of Plant Physiology, Vol. 16B: Photomorphogenesis (pp. 545–68). Berlin: Springer–Verlag.Google Scholar
Dring, M. J., and West, J. A. (1983). Photoperiodic control of tetrasporangium formation in the red algaRhodochorton purpureum. Planta 159:143–50.Google ScholarPubMed
Driskell, W. B., Ruesink, J. L., Lees, D. C., Houghton, J. P., and Lindstrom, S. C. (2001). Long-term signal of disturbance: Fucus gardneri after the Exxon Valdez oil spill. Ecol. Applic. 11:815–27.CrossRefGoogle Scholar
Dromgoole, F. I. (1978). The effects of oxygen on dark respiration and apparent photosynthesis of marine macroalgae. Aquat. Bot. 4:281–97.CrossRefGoogle Scholar
Dromgoole, F. I. (1982). The buoyant properties of Codium. Bot. Mar. 25:391–7.CrossRefGoogle Scholar
Dromgoole, F. I. (1990). Gas-filled structures, buoyancy and support in marine macro-algae. Prog. Phycol. Res. 7: 169–211.Google Scholar
Druehl, L. D. (1978) The distribution of Macrocystis integrifolia in British Columbia as related to environmental parameters. Can. J. Bot. 56: 69–79.CrossRefGoogle Scholar
Druehl, L. D. (1981). Geographic distribution. In Lobban, C. S. and Wynne, M. J. (eds), The Biology of Seaweeds (pp. 306–25). Oxford: Blackwell Scientific.Google Scholar
Druehl, L. D. (1988). Cultivated edible kelp. In Lembi, C. A. and Waaland, J. R. (eds), Algae and Human Affairs (pp. 119–34). Cambridge: Cambridge University Press.Google Scholar
Druehl, L. D. (2001). Pacific Seaweeds: A Guide to Common Seaweeds of the West Coast. Madeira Park, Canada: Harbour Publishing.Google Scholar
Druehl, L. D., and Footit, R. G. (1985). Biogeographical analysis. In Littler, M. M. and Littler, D. S. (eds) Handbook of Phycological Methods: Ecological Field Methods (pp. 315–25). Cambridge: Cambridge University Press.Google Scholar
Druehl, L. D., and Green, J. M. (1982). Vertical distribution of intertidal seaweeds as related to patterns of submersion and emersion. Mar. Ecol. Prog. Ser. 9:163–70.CrossRefGoogle Scholar
Druehl, L. D., and Saunders, G. W. (1992). Molecular explorations of kelp evolution. Prog. Phycol. Res. 8:47–83.CrossRefGoogle Scholar
Druehl, L. D., Baird, R., Lindwall, A., Lloyd, K. E., and Pakula, S. (1988). Longline cultivation of some Laminariaceae in British Columbia. Aquacult. Fish. Management 19:253–63.Google Scholar
Druehl, L. D., Collins, J. D., Lane, C. E., and Saunders, G. W. (2005a). An evaluation of methods used to assess intergeneric hybridization in kelp using Pacific Laminariales (Phaeophyceae). J. Phycol. 41:250–62.CrossRefGoogle Scholar
Druehl, L.D., Collins, J. D., Lane, C. E., and Saunders, G. W. (2005b). A critique of intergeneric kelp hybridization protocol, employing Pacific Laminariales (Phaeophyceae). J. Phycol. 41: 250–62.CrossRefGoogle Scholar
Duarte, C. M. (1992). Nutrient concentration of aquatic plants patterns across species. Limnol. Oceanogr. 37:882–9. [6, ]CrossRefGoogle Scholar
Duarte, P., and Ferreira, J. G. (1993). A methodology for parameter estimation in seaweed productivity modelling. Hydrobiologia 260/262:183–9.CrossRefGoogle Scholar
Duarte, P., and Ferreira, J. G. (1997). A model for the simulation of macroalgal population dynamics and productivity. Ecol. Model. 98:199–214.CrossRefGoogle Scholar
Dube, M. A., and Ball, E. (1971). Desmarestia sp. associated with the seapen Ptilosarcus gurneyi (Gray). J. Phycol. 7:218–20.Google Scholar
Ducreux, G. (1984). Experimental modification of the morphogenetic behavior of the isolated sub-apical cell of the apex of Sphacelaria cirrosa (Phaeophyceae). J. Phycol. 20:447–54.CrossRefGoogle Scholar
Ducreux, G., and Kloareg, B. (1988). Plant regeneration from protoplasts of Sphacelaria (Phaeophyceae). Planta 174:259.CrossRefGoogle Scholar
Dudgeon, S., and Petraitis, P. S. (2005). First year demography of the foundation species, Ascophyllum nodosum, and its community implications. Oikos 109:405–15.CrossRefGoogle Scholar
Dudgeon, S. R., Davison, I. R., and Vadas, R. L. (1989). Effect of freezing on photosynthesis of intertidal macroalgae: relative tolerance of Chondrus crispus and Mastocarpus stellatus (Rhodophyta). Mar. Biol. 101:107–14.CrossRefGoogle Scholar
Dudgeon, S. R., Davison, I. R., and Vadas, R. L. (1990). Freezing tolerance in the intertidal red algae Chondrus crispus and Mastocarpus stellatus: relative importance of acclimation and adaptation. Mar Biol 106:427–36.CrossRefGoogle Scholar
Dudgeon, S. R., Kubler, J. E., Vadas, R. L., and Davison, I. R. (1995). Physiological responses to environmental variation in intertidal red algae; does thallus morphology matter? Mar. Ecol. Prog. Ser. 117:193–206.CrossRefGoogle Scholar
Duffy, J. E. (1990). Amphipods on seaweeds: partners or pests? Oecologia 83:267–76.CrossRefGoogle Scholar
Duffy, J. E. (2009). Why biodiversity is important to the functioning of real-world ecosystems. Front. Ecol. Environ. 7:437–44.CrossRefGoogle Scholar
Duffy, J. E., and Hay, M. E. (1990). Seaweed adaptations to herbivory. BioScience 40:368–75. [3.2.2]CrossRefGoogle Scholar
Duffy, J. E., and Hay, M. E. (2000). Strong impacts of grazing amphipods on the organization of a benthic community. Ecological Monographs 70(2): 237–63.CrossRefGoogle Scholar
Duffy, J. E., and Hay, M. E. (2001). The ecology and evolution of marine consumer–prey interactions. In Bertness, M. D., Gaines, S. D., and Hay, M. E. (eds), Marine Community Ecology (pp. 131–57). Sunderland, MA: Sinauer Assocs.Google Scholar
Duggins, D. O. (1980). Kelp beds and sea otters: an experimental approach. Ecology 61:447–53.CrossRefGoogle Scholar
Duggins, D. O., Simenstad, C. A., and Estes, J. A. (1989). Magnification of secondary production by kelp detritus in coastal marine ecosystems. Science 245:170–3.CrossRefGoogle ScholarPubMed
Duke, C. S., Litaker, R. W., and Ramus, J. (1987). Seasonal variation in RuBPCase activity and N allocation in the chlorophyte seaweeds Viva curvata (Kiitz.) de Toni and Codium decorticatum (Woodw.) Howe. J. Exp. Mar. Biol. Ecol. 112:145–64.CrossRefGoogle Scholar
Duke, C. S., Litaker, W., and Ramus, J. (1989). Effects of temperature, nitrogen supply, and tissue nitrogen on ammonium uptake rates of the chlorophyte seaweeds Ulva curvata and Codium decorticatum. J. Phycol. 25:113–20.CrossRefGoogle Scholar
Duncan, M. J., and Foreman, R. E. (1980). Phytochromemediated stipe elongation in the kelp Nereocystis (Phaeophyceae). J. Phycol. 16:138–42.CrossRefGoogle Scholar
Dunn, E. K., Shoue, D. A., Huang, X., et al. (2007). Spectroscopic and biochemical analysis of regions of the cell wall of the unicellular “Mannan Weed”, Acetabularia acetabulum. Plant Cell Physiol. 48:122–33.CrossRefGoogle ScholarPubMed
Dunton, K. H., and Schell, D. M. (1986). Seasonal carbon budget and growth of Laminaria solidungula in the Alaskan high Arctic. Mar. Ecol. Prog. Ser. 31:57–66.CrossRefGoogle Scholar
Dunton, K. H., and Dayton, P. K. (1995). The biology of high latitude kelp. In Skjoldal, H. R., Hopkins, C., Erikstad, K. E., and Leinaas, H. P. (eds) Ecology of Fjords and Coastal Waters (pp. 499–507). Amsterdam:Elsevier Science.Google Scholar
Durako, M. J., and Dawes, C. J. (1980). A comparative seasonal study of two populations of Hypnea musciformis from the east and west coasts of Florida. USA II Photosynthetic and respiratory rates. Mar. Biol. 59:157–62.CrossRefGoogle Scholar
Durante, K. M., and Chia, F. S. (1991). Epiphytism on Agarum fimbriatum: can herbivore preferences explain distributions of epiphytic bryozoans? Mar. Ecol. Prog. Ser. 77:279–87.CrossRefGoogle Scholar
Dworjanyn, S. A., de Nys, R., and Steinberg, P. D. (1999). Localisation and surface quantification of secondary metabolites. Mar. Biol. 133:727–36.CrossRefGoogle Scholar
Dworjanyn, S. A., de Nys, R., and Steinberg, P. D.. (2006a). Chemically mediated antifouling in the red alga Delisea pulchra. Mar. Ecol. Prog. Ser. 318: 153–63.CrossRefGoogle Scholar
Dworjanyn, S. A., Wright, J. T., Paul, N. A., de Nys, R., and Steinberg, P. D. (2006b). Cost of chemical defence in the red alga Delisea pulchra. Oikos 113:13–22.CrossRefGoogle Scholar
Dyck, L. J., and DeWreede, R. E. (2006a). Seasonal and spatial patterns of population density in the marine macroalga Mazzaella splendens (Gigartinales, Rhodophyta). Phycol. Res. 54:21–31.CrossRefGoogle Scholar
Dyck, L. J., and DeWreede, R. E. (2006b). Reproduction and survival in Mazzaella splendens (Gigartinales, Rhodophyta). Phycologia 45:302–10.CrossRefGoogle Scholar
Eardley, D. D., Sutton, C. W., Hempel, W. M., Reed, D. C., and Ebeling, A. W. (1990). Monoclonal-antibodies specific for sulfated polysaccharides on the surface of Macrocystis pyrifera (Phaeophyceae). J. Phycol. 26:54–62.CrossRefGoogle Scholar
Edelstein, T., and McLachlan, J. (1975). Autecology of Fucus distichus ssp. distichus (Phaeophyceae: Fucales) in Nova Scotia, Canada. Mar Biol 30:305–24.CrossRefGoogle Scholar
Edwards, D. M., Reed, R. H., Chudek, J. A., Foster, R. M., and Stewart, W. D. P. (1987). Organic solute accumulation in osmotically-stressed Enteromorpha intestinalis. Mar. Biol. 95: 583–92.CrossRefGoogle Scholar
Edwards, M. S., and Connell, S. D. (2012). Competition, a major factor structuring seaweed communities. In Wiencke, C., and Bischof, K. (eds), Seaweed Biology: Novel Insights into Ecophysiology, Ecology and Utilization, Ecological Studies 219 (pp. 135–56). Berlin and Heidelberg: Springer-Verlag.CrossRefGoogle Scholar
Edwards, M. S., and Estes, J. A. (2006). Catastrophe, recovery and range limitation in NE Pacific kelp forests: a large-scale perspective. Mar. Ecol. Prog. Ser. 320:79–87.CrossRefGoogle Scholar
Edwards, M. S., and Hernández–Carmona, G. (2005). Delayed recovery of giant kelp near its southern range limit in the North Pacific following El Niño. Mar Biol 147:273–9.CrossRefGoogle Scholar
Eggert, A., and Karsten, U. (2010). Low molecular weight carbohydrates in red algae – an ecophysiological and biochemical perspective. In Seckbach, J., Chapman, D., and Weber, A. (eds) Cellular Origins, Life in Extreme Habitats and Astrobiology, Red Algae in the Genomic Age (pp. 445–56). Berlin and Heidelberg: Springer-Verlag.Google Scholar
Eggert, A., Burger, E. M., and Breeman, A. M. (2003). Ecotypic differentiation in thermal traits in the tropical to warm-temperate green macrophyte Valonia utricularis. Bot. Mar. 46:69–81.CrossRefGoogle Scholar
Eggert, A., Visser, R. J. W., van Hasselt, P. R., and Breeman, A. M. (2006). Differences in acclimation potential of photosynthesis in seven isolates of the tropical to warm temperate macrophyte Valonia utricularis (Chlorophyta). Phycologia 45:546–56.CrossRefGoogle Scholar
Eide, I., Myklestad, S., and Melson, S. (1980). Long-term uptake and release of heavy metals by Ascophyllum nodosum (L.) Le Jol. (Phaeophyceae) in situ. Environ. Pollut. 23:19–28.CrossRefGoogle Scholar
Eklund, B. (2005). Development of a growth inhibition test with the marine and brackish water red alga Ceramium tenuicorne. Mar. Poll. Bull. 50:921–30. [, , ]CrossRefGoogle ScholarPubMed
Eklund, B., and Kautsky, L. (2003). Review on toxicity testing with marine macroalgae and the need for method standardization – exemplified with copper and phenol. Mar. Poll. Bull. 46:171–81.CrossRefGoogle ScholarPubMed
Eklund, B., Elfström, M., and Borg, H. (2008). TBT originates from pleasure boats in Sweden in spite of firm restrictions. Open Environ Sci. 2:124–32.CrossRefGoogle Scholar
Eklund, B., Elfström, M., Gallego, I., Bengtsson, B-E., and Breitholtz, M. (2010). Biological and chemical characterization of harbour sediments from the Stockholm area. Soil Sed. Poll. 10:127–41.CrossRefGoogle Scholar
Elner, R. W., and Vadas, Sr R. L.. (1990). Inference in ecology: the sea urchin phenomenon in the northwest Atlantic. Am. Nat. 136:108–25.CrossRefGoogle Scholar
Emerson, C. J., Buggeln, R. G., and Bal, A. K. (1982). Translocation in Saccorhiza dermatodea (Laminariales, Phaeophyceae): anatomy and physiology. Can. J. Bot. 60: 2164–84.CrossRefGoogle Scholar
Engel, C. R., Wattier, R., Destombe, C., and Valero, M. (1999). Performance of non-motile male gametes in the sea: analysis of paternity and fertilization success in a natural population of a red seaweed, Gracilaria gracilis. Proc. Roy. Soc. Lond. B. Bio. 266:1879–86.CrossRefGoogle Scholar
Engel, C. R., Daguin, C., and Serrão, E. A. (2005). Genetic entities and mating system in hermaphroditic Fucus spiralis and its close dioecious relative F. vesiculosus (Fucaceae, Phaeophyceae). Mol. Ecol. 14:2033–46.CrossRefGoogle Scholar
Engelen, A., and Santos, R. (2009). Which demographic traits determine population growth in the invasive brown seaweed Sargassum muticum? J. Ecol. 97:675–84.CrossRefGoogle Scholar
Engelmann, T. W. (1883). Farbe und Assimilation. Bot Zeit 41:1–13.Google Scholar
Engelmann, T. W. (1884). Untersuchungen iiber die quantitativen Beziehungen zwischen Absorption des Lichtes und Assimilation in Pflanzenzellen. Bot Zeit 42:81–93.Google Scholar
Enright, C. T. (1979). Competitive interaction between Chondrus crispus (Florideophyceae) and Ulva lactuca (Chlorophyceae) in Chondrus aquaculture. Proc. Intl. Seaweed Symp. 9:209–18.Google Scholar
Enríquez, S., and Rodríquez-Román, A. (2006). Effect of water flow on the photosynthesis of three marine macrophytes from a fringing-reef lagoon. Mar. Ecol. Prog. Ser. 323:119–32.CrossRefGoogle Scholar
Enríquez, S., Duarte, C. M., Sand-Jensen, K., and Nielsen, S. L. (1996). Broad-scale comparison of photosynthetic rates across phototrophic organisms. Oecologia 108:197–206.CrossRefGoogle ScholarPubMed
Enríquez, S., Ávila, E., and Carballo, J. L. (2009). Phenotypic plasticity induced in transplant experiments in a mutualistic association between the red alga Jania adhaerens (Rhodophyta, Corallinales) and the sponge Haliclona caerulea (Porifera: Haplosclerida): morphological responses of the alga. J. Phycol. 45:81–90.CrossRefGoogle Scholar
Estes, J. A., and Steinberg, P. D., (1988). Predation, herbivory, and kelp evolution. Paleobiology 14:19–36.CrossRefGoogle Scholar
Estes, J. A., Danner, E. M., Doak, D. F., et al. (2004). Complex trophic interactions in kelp forest ecosystems. B. Mar. Sci. 74:621–38.Google Scholar
Estes, J. A., Tinker, M. T., Williams, T. M., et al. (1998). Killer whale predation on sea otters linking oceanic and nearshore ecosystems. Science 282:473–6.CrossRefGoogle ScholarPubMed
Estevez, J. M., and Cáceres, E. J. (2003). Fine structural study of the red seaweed Gymnogongrus torulosus (Phyllophoraceae, Rhodophyta). Biocell 27:181–7.Google Scholar
Evans, L. V., and Butler, D. M. (1988). Seaweed biotechnology: current status and future prospects. In Rogers, L. J. and Gallon, J. R. (eds.) Biochemistry of the Algae and Cyanobacteria (pp. 335–50). Oxford: Clarendon Press.Google Scholar
Evans, L. V., and Christie, A. O. (1970). Studies on the shipfouling alga Enteromorpha. I. Aspects of the fine-structure and biochemistry of swimming and newly settled zoospores. Ann. Bot. 34:451–66.CrossRefGoogle Scholar
Evans, L. V., Callow, J. A., and Callow, M. E. (1973). Structural and physiological studies on the parasitic red algaHolmsella. New Phytol. 72:393–402.CrossRefGoogle Scholar
Evans, L. V., Callow, J. A., and Callow, M. E. (1982). The biology and biochemistry of reproduction and early development in Fucus. Prog. Phycol. Res. 1:67–110.Google Scholar
Fabricius, K. E. (2011). Factors determining the resilience of coral reefs to eutrophication: a review and conceptual model. In Dubinsky, Z. and Stambler, N. (eds), Coral Reefs: An Ecosystem in Transition (pp. 493–506) Berlin and Heidelberg: Springer–Verlag.CrossRefGoogle Scholar
Fabricus, K., De’ath, G., McCook, L., Turak, E., and Williams, D. McB. (2005). Changes in algal, coral and fish assemblages along water quality gradients on the inshore Great Barrier Reef. Mar. Poll. Bull. 51:384–98.CrossRefGoogle Scholar
Fabricius, K. E., Langdon, C., Uthicke, S., et al. (2011). Losers and winners in coral reefs acclimatized to elevated carbon dioxide concentrations. Nature Clim. Change 1:165–9.CrossRefGoogle Scholar
Faes, V. A., and Viejo, R. M. (2003). Structure and dynamics of a population of Palmaria palmata (Rhodophyta) in northern Spain. J. Phycol. 39:1038–49.CrossRefGoogle Scholar
Fagerberg, W. R., and Dawes, C. J. (1977). Studies on Sargassum. II. Quantitative ultrastructural changes in differentiated stipe cells during wound regeneration and regrowth. Protoplasma 92:211–27.CrossRefGoogle Scholar
Fagerberg, W. R., (Lavoie) Hodges, E., and Dawes, C. J. (2010). The development and potential roles of cell wall trabeculae in Caulerpa mexicana (Chlorophyta). J. Phycol. 46:309–15.CrossRefGoogle Scholar
Fain, S. R., Druehl, L. D., and Baillie, D. L. (1988). Repeat and single copy sequences are differentially conserved in the evolution of kelp chloroplast DNA. J. Phycol. 24:292–302.Google Scholar
Fairhead, V. A., Amsler, C. D., McClintock, J. B., and Baker, B. J. (2005). Within-thallus variation in chemical and physical defences in two species of ecologically dominant brown macroalgae from the Antarctic Peninsula. J. Exp. Mar. Biol. Ecol. 322:1–12.[]CrossRefGoogle Scholar
Falcão, V. R., Oliviera, M. C., and Colepicolo, P. (2010). Molecular characterization of nitrate reductase gene and its expression in the marine red alga Gracilaria tenuistipitata (Rhodophyta). J. Appl. Phycol. 22:613–22.CrossRefGoogle Scholar
Falkowski, P. G., and LaRoche, J. (1991). Acclimation to spectral irradiance in algae. J. Phycol. 27:8–14.CrossRefGoogle Scholar
Falkowski, P. G., and Raven, J. A. (2007). Aquatic Photosynthesis, 2nd edn. Princeton, NJ: Princeton University Press. 484 pp. [5]Google Scholar
Falkowski, P. G., McClosky, L., Muscatine, L., and Dubinsky, Z. (1993). Population control in symbiotic corals. BioSci. 43:606–11.CrossRefGoogle Scholar
Fantle, M. S., Dittel, A. I., Schwalm, S. M., Epifanio, C. E., and Fogel, M. L. (1999). A food web analysis of the juvenile blue crab, Callinectes sapidus, using stable isotopes in whole animals and individual amino acids. Oecologia 120(3): 416–26.CrossRefGoogle ScholarPubMed
FAO (2005). Cultured aquaculture species information program: Porphyra spp. Rome: FAO. Available at: .
FAO (2006). Cultured aquaculture species information program: Laminaria japonica. Rome: FAO. Available at: .
FAO (2008). Cultured aquaculture species information program: Euchema spp. Rome: FAO. Available at: .
FAO (2009). World Review of Fisheries and Aquaculture. FAO Fisheries and Aquaculture Dept. Rome: FAO. 84 pp. [, , ].Google Scholar
FAO (2010). The State of World Fisheries and Aquaculture 2010. FAO Fisheries and Aquaculture Dept. Rome: FAO..Google Scholar
Fath, B. (eds), The Encyclopedia of Ecology, Ecological Engineering, Vol. 3 (pp. 2463–75). Oxford: Elsevier.
Fawley, M. W., Douglas, C. A., Stewart, K. D., and Mattox, K. R. (1990). Light-harvesting pigment-protein complexes of the Ulvophyceae (Chlorophyta): characterization and phylogenetic significance. J. Phycol. 26:186–95.CrossRefGoogle Scholar
Fei, X. G. (2004). Solving the coastal eutrophication problem by large scale seaweed cultivation. Hydrobiologia, 512(1–3):145–51.CrossRefGoogle Scholar
Fernández, P. V., Ciancia, M., Miravalles, A. B., and Estevez, J. M. (2010). Cell wall polymer mapping in the coennocytic macroalga Codium vermilara (Bryopsidales, Chlorophyta). J. Phycol. 46:456–65.CrossRefGoogle Scholar
Ferreira, J. G., and Ramos, L. (1989). A model for the estimation of annual production rates of macrophyte algae. Aquat. Bot. 33:53–70.CrossRefGoogle Scholar
Fetter, R., and Neushul, M. (1981). Studies in developing and released spermatia in the red alga Tiffaniella snyderae (Rhodophyta). J. Phycol. 17:141–59.CrossRefGoogle Scholar
Fielding, A. H., and Russell, G. (1976). The effect of copper on competition between marine algae. J. Appl. Ecol. 13:871–6.CrossRefGoogle Scholar
Fierst, J., terHorst, C., Kübler, J. E., and Dudgeon, S. (2005). Fertilization success can drive patterns of phase dominance in complex life histories. J. Phycol. 41:238–49.CrossRefGoogle Scholar
Figueroa, F. L., Aguilera, J., and Niell, F. X. (1994). End-of-day light control of growth and pigmentation in the red alga Porphyra umbilicalis (L.) Kützing. Z. Naturforsch. 49:593–600.Google Scholar
Figurski, J. D., Malone, D., Lacy, J. R., and Denny, M. (2011). An inexpensive instrument for measuring wave exposure and water velocity. Limnol. Oceanogr.: Methods 9:204–14.CrossRefGoogle Scholar
Filion-Myklebust, C., and Norton, T. A. (1981). Epidermis shedding in the brown seaweed Ascophyllum nodosum (L.) Le Jolis, and its ecological significance. Mar. Biol. Lett. 2:45–51.Google Scholar
Fischer, G., and Wiencke, C. (1992). Stable carbon isotope composition, depth distribution and fate of macroalgae from the Antarctic Peninsula region. Polar Biology 12:341–8.CrossRefGoogle Scholar
Fjeld, A. (1972). Genetic control of cellular differentiation in Ulva mutabilis. Gene effects in early development. Devel. Biol. 28:326–43.CrossRefGoogle ScholarPubMed
Fjeld, A., and Løvlie, A. (1976). Genetics of multicellular marine algae. In Lewin, R. A. (ed.), The Genetics of Algae (pp. 219–35). Oxford: Blackwell Scientific.Google Scholar
Fletcher, R. L., and Callow, M. E. (1992). The settlement, attachment and establishment of marine algal spores. Br. Phycol. J. 27:303–29. [, , , ]CrossRefGoogle Scholar
Fletcher, R. L., Baier, R. E., and Fornalik, M. S. (1985). The effects of surface energy on germling development of some marine macroalgae (abstract). Br. Phycol. J. 20: 184–5.Google Scholar
Floc’h, J.–Y. (1982). Uptake of inorganic ions and their long distance transport in Fucales and Laminariales. In Srivastava, L. M. (ed.), Synthetic and Degradative Processes in Marine Macrophytes (pp. 139–65). Berlin: WaIter de Gruyter.Google Scholar
Floeter, S. R., Behrens, M. D., Ferreira, C. E. L., Paddack, M. J., and Horn, M. H. (2005). Geographical gradients of marine herbivorous fishes: patterns and processes. Mar. Biol. 147:1435–47.CrossRefGoogle Scholar
Florence, T. M., Lumsden, B. G., and Fardy, J. J. (1984). Algae as indicators of copper speciation. In Kramer, C. J. M. and Duinker, J. C. (eds), Complexation of Trace Metals in Natural Waters (pp. 411–18). The Netherlands: Dr. W. Junk.CrossRefGoogle Scholar
Flores-Moya, A. (2012). Warm temperate seaweed communities: a case study of deep water kelp forests from the Alboran Sea (SW Mediterranean Sea) and the Strait of Gibraltar. In Wiencke, C. and Bischof, K. (eds), Seaweed Biology: Novel Insights Into Ecophysiology, Ecology and Utilization (pp. 315–27). Series: Ecological Studies, Volume 219. Berlin and Heidelberg: Springer.CrossRefGoogle Scholar
Flores–Moya, A., Fernández, J. A., and Niell, F. X. (1997). Growth pattern, reproduction and self-thinning in seaweeds: a re-evaluation in reply to Scrosati. J. Phycol. 33:1080–1.CrossRefGoogle Scholar
Fong, P., and Paul, V. J. (2011). Coral reef algae. In Dubinsky, Z. and Stambler, N. (eds), Coral Reefs: An Ecosystem in Transition (pp. 241–72). Dordrecht, The Netherlands: Springer.CrossRefGoogle Scholar
Fong, P., Donohoe, R. M., and Zedler, J. B. (1994). Nutrient concentration in the tissue of the macroalga Enteromorpha as a function of nutrient history: an experimental evaluation using field microcosms. Mar. Ecol. Prog. Ser. 106:273–81.CrossRefGoogle Scholar
Fork, D. C. (1963). Observations on the function of chlorophyll a and accessory pigments in photosynthesis. In Photosynthetic Mechanisms in Green Plants (pp. 352–61). Publ. no. 1145, NAS–NRC, Washington, DC.Google Scholar
Forrest, B. M., Brown, S. N., Taylor, M. D., Hurd, C. L., and Hay, C. H. (2000). The role of natural dispersal mechanisms in the spread of Undaria pinnatifida (Laminariales, Phaeophyta). Phycologia 39:547–53.CrossRefGoogle Scholar
Forsberg, A., Soderlund, S., Frank, A., Petersson, L. R., and Pedersen, M. (1988). Studies on metal content in the brown seaweed, Fucus vesiculosus, from the Archipelago of Stockholm. Environ. Pollut. 49:245–63.CrossRefGoogle ScholarPubMed
Forsberg, J., and Alen, J. F. (2001). Molecular recognition in thylakoid structure and function. Trends in Plant Sciences 6:317–26.Google Scholar
Förster, T. (1948). Zwischenmolekulare Energiewanderung und Fluoreszenz. Ann. Physik 437:5.CrossRefGoogle Scholar
Foster, M. S. (1972). The algal turf community in the nest of the ocean goldfish (Hypsypops rubicunda). Proc. Intl. Seaweed Symp. 7:55–60.Google Scholar
Foster, M. S. (1975). Regulation of algal community development in a Macrocystis pyrifera forest. Mar. Biol. 32:331–42.CrossRefGoogle Scholar
Foster, M. S., and Schiel, D. R. (2010). Loss of predators and the collapse of southern California kelp forests (?): alternatives, explanations and generalizations. J. Exp. Mar. Biol. Ecol. 393:59–70.CrossRefGoogle Scholar
Foster, M. S., and Sousa, W. P. (1985). Succession. In Littler, M. M. and Littler, D. S. (eds), Handbook of Phycological Methods: Ecological Field Methods: Macroalgae (pp. 269–90). Cambridge: Cambridge University Press.Google Scholar
Foster, M. S., Edwards, M. S., Reed, D. C., et al. (2006). Top-down vs. bottom-up effects in kelp forests. Science 313:1737–8.CrossRefGoogle ScholarPubMed
Foster, N. L., Box, S. J., and Mumby, P. J. (2008). Competitive effects of macroalgae on the fecundity of the reef-building coral Montastraea annularis. Mar. Ecol. Prog. Ser. 367:143–52.CrossRefGoogle Scholar
Foster, P. (1976). Concentrations and concentration factors of heavy metals in brown algae. Environ. Pollut. 10:45–54.CrossRefGoogle Scholar
Fowler, J. E., and Quantrano, R. S. (1997). Plant cell morphogenesis: plasma membrane interactions with the cytoskeleton and cell wall. Annu. Rev. Cell Dev. Biol. 13:697–743.CrossRefGoogle Scholar
Fowler-Walker, M. J., and Wernberg, T. (2006). Differences in kelp morphology between wave sheltered and exposed localities: morphologically plastic or fixed traits? Mar. Biol. 148:755–67.CrossRefGoogle Scholar
Fox, C.H., and Swanson, A. K. (2007). Nested PCR detection of microscopic life-stages of laminarian macroalgae and comparison with adult forms along intertidal height gradients. Mar. Ecol. Prog. Ser. 332: 1–10.CrossRefGoogle Scholar
Fralick, R. A., and Mathieson, A. C. (1975). Physiological ecology of four Polysiphonia species (Rhodophyta, Ceramiales). Mar. Biol. 29:29–36.CrossRefGoogle Scholar
Franklin, L. A., and Forster, R. M. (1997). The changing irradiance environment: consequences for marine macrophyte physiology, productivity and ecology. Eur. J. Phycol. 32:207–32.Google Scholar
Franklin, L. A., Osmond, B., and Larkum, A. W. D. (2003): Photoinhibition, UV-B and algal photosynthesis. In Larkum, A. W. D., Douglas, S. E., and Raven, J. A. (eds) Photosynthesis in Algae. Advances in Photosynthesis and Respiration (Vol. 14, pp. 11–28). Dordrecht, The Netherlands: Kluwer Academic Publishers.Google Scholar
Fraschetti, S., Terlizzi, A., and Benedetti-Cecchi, L. (2005). Patterns of distribution of marine assemblages from rocky shores: evidence of relevant scales of variation. Mar. Ecol. Prog. Ser. 296:13–29.CrossRefGoogle Scholar
Fraser, C. I., Hay, C. H., Spenser, H. G., and Waters, J. M. (2009). Genetic and morphological analyses of the southern bull kelp Durvillaea antarctica (Phaeophyceae: Durvillaeales) in New Zealand reveal cryptic species. J. Phycol. 45:436–43. [, , ]CrossRefGoogle ScholarPubMed
Frederick, J. E., Snell, H. E., and Haywood, E. K. (1989). Solar ultraviolet radiation at the earth´s surface. Photochem. Photobiol. 50:443–50.CrossRefGoogle Scholar
Fredersdorf, J., and Bischof, K. (2007). Irradiance of photosynthetically active radiation determines UV-susceptibility of photosynthesis in Ulva lactuca L. (Chlorophyta). Phycol. Res. 55:295–301.CrossRefGoogle Scholar
Fredersdorf, J., Müller, R., Becker, S., Wiencke, C., and Bischof, K. (2009). Interactive effects of radiation, temperature and salinity on different life history stages of the Arctic kelp Alaria esculenta (Phaeophyceae). Oecologia 160:483–92. [, , ]CrossRefGoogle Scholar
Freidenburg, T. L. (2002). Macroscale to Local Scale Variation in Rocky Intertidal Community Structure and Dynamics in Relation to Coastal Upwelling. PhD Dissertation, Oregon State University.
Fretwell, S. D. (1977). The regulation of plant communities by food chains exploiting them. Perspect. Biol. Med. 20:169–85.CrossRefGoogle Scholar
Fricke, A., Teichberg, M., Beilfuss, S., and Bischof, K. (2011a). Succession patterns in algal turf vegetation on a Caribbean coral reef. Bot. Mar. 54:111–26.CrossRefGoogle Scholar
Fricke, A., Titlyanova, T. V., Nugues, M. M., and Bischof, K. (2011b). Depth-related variation in epiphytic communities growing on the brown alga Lobophora variegata in a Caribbean coral reef. Coral Reefs 30:967–73.CrossRefGoogle Scholar
Friedlander, M. (2008). Advances in cultivation of Gelidiales. J. Appl. Phycol. 20:451–6.CrossRefGoogle Scholar
Friedlander, M., and Dawes, C. J. (1985). In situ uptake kinetics of ammonium and phosphate and chemical composition of the red seaweed Gracilaria tikvahiae. J. Phycol. 21:448–53.CrossRefGoogle Scholar
Friedlander, M., Krom, M. D., and Ben-Amotz, A. (1991). The effect of light and ammonium on growth, epiphytes and chemical constituents of Gracilaria conferta in outdoor cultures. Bot. Mar. 34:161–6.CrossRefGoogle Scholar
Friedmann, E. I. (1961). Cinemicrography of spermatozoids and fertilization in Fucales. Bull. Res. Counc. Israel 10D:73–83.Google Scholar
Friedrich, M. W. (2012). Bacterial communities on macroalgae. In Wiencke, C. and Bischof, K. (eds) (2012). Seaweed Biology: Novel Insights into Ecophysiology, Ecology and Utilization, Ecological Studies 219 (pp. 189–202). Berlin, Heidelberg: Springer-Verlag.CrossRefGoogle Scholar
Fries, L. (1966). Temperature optima of some red algae in axenic culture. Bot. Mar. 9:12–14.CrossRefGoogle Scholar
Fries, L. (1982). Selenium stimulates growth of marine macroalgae in axenic culture. J. Phycol. 18:328–31.CrossRefGoogle Scholar
Fries, N. (1979). Physiological characteristics of Mycosphaerella ascophylli, a fungal endophyte of the marine brown alga Ascophyllum nodosum. Physiol. Plant. 45:117–21.CrossRefGoogle Scholar
Fritsch, F. E. (1945). The Structure and Reproduction of the Algae, Vol. 2. Cambridge: Cambridge University Press.Google Scholar
Fry, B., Scalan, R. S., Winters, J. J., and Parker, P. L. (1982). Sulphur uptake by salt grasses, mangroves and seagrasses in anaerobic sediments. Geochim. Cosmochim. Acta 43:1121–4.CrossRefGoogle Scholar
Fu, W., Yao, J., Wang, X., et al. (2009). Molecular cloning and expression analysis of a cytosolic HSP70 gene from Laminaria japonica (Laminariaceae, Phaeophyta). Mar. Biotechnol. 11:738–47.CrossRefGoogle Scholar
Fujimura, T., Kawai, T., Shiga, M., Kajiwara, T., and Hatanaka, A. (1989). Regeneration of protoplasts into complete thalli in the marine green alga UIva pertusa. Nippon Suisan Gakkaishi 55:1353–9.CrossRefGoogle Scholar
Fujita, R. M. (1985). The role of nitrogen status in regulating transient ammonium uptake and nitrogen storage by macroalgae. J. Exp. Mar. Biol. Ecol. 99:283–301.CrossRefGoogle Scholar
Fujita, R. M., and Migata, S. (1986). Isolation of protoplasts from leaves of red algae Porphyra yezoensis. Jpn. J. Phycol. 34:63.Google Scholar
Fujita, R. M., Wheeler, P. A., and Edwards, R. L. (1988). Metabolic regulation of ammonium uptake by Ulva rigida (Chlorophyta): a compartmental analysis of the rate-limiting step for uptake. J. Phycol. 24:560–6.CrossRefGoogle Scholar
Fujita, R. M., Wheeler, P. A., and Edwards, R. L. (1989). Assessment of macroalgal nutrient limitation in a seasonal upwelling region. Mar. Ecol. Prog. Ser. 53:293–303.CrossRefGoogle Scholar
Fujita, S., Iseki, M., Yoshikawa, S., Makino, Y., et al. (2005). Identification and characterization of a fluorescent flagellar protein from the brown alga Scytosiphon lomentaria (Scytosiphonales, Phaeophyceae): a flavoprotein homologous to old yellow enzyme. Eur. J. Phycol. 40:159–67.CrossRefGoogle Scholar
Fukuhara, Y., Mizuta, H., and Yasui, H. (2002). Swimming activities of zoospores in Laminaria japonica (Phaeophyceae). Fisheries Sci. 68:1173–81.CrossRefGoogle Scholar
Fulcher, R. G., and McCully, M. E. (1969). Histological studies on the genus Fucus. IV. Regeneration and adventive embryony. Can. J. Bot. 47:1643–9.CrossRefGoogle Scholar
Fulcher, R. G., and McCully, M. E. (1971). Histological studies on the genus Fucus. V. An autoradiographic and electron microscopic study of the early stages of regeneration. Can. J. Bot. 49:161–5.CrossRefGoogle Scholar
Gacesa, P. (1988). Alginates. Carbohydr. Polymers 8: 161–82.CrossRefGoogle Scholar
Gachon, C. M. M., Sime-Ngando, T., Strittmatter, M., Chambouvet, A., and Kim, G.H. (2010). Algal diseases: spotlight on a black box. Trends Pl. Sci. 15:633–40.CrossRefGoogle ScholarPubMed
Gacia, E., Rodriguez-Prieto, C., Delgado, O., and Ballesteros, E. (1996). Seasonal light and temperature responses of Caulerpa taxifolia from the northwestern Mediterranean. Aquat. Bot. 53:215–25.CrossRefGoogle Scholar
Gagne, J. A., and Mann, K. H. (1987). Evaluation of four models used to estimate kelp productivity from growth measurements. Mar. Ecol. Prog. Ser. 37: 35–44.CrossRefGoogle Scholar
Gagne, J. A., Mann, K. H., and Chapman, A. R. O. (1982). Seasonal patterns of growth and storage in Laminaria longicruris in relation to differing patterns of availability of nitrogen in the water. Mar. Biol. 69:91–101. [, , ]CrossRefGoogle Scholar
Gagnon, P., Scheibling, R. E., Jones, W., and Tully, D. (2008). The role of digital bathmetry in mapping shallow marine vegetation from hyperspectral image data. Int. J. Remote Sens. 29:879–904.CrossRefGoogle Scholar
Gaillard, J., and L’Hardy-Halos, M. T. (1990). Morphogenèse du Dictyota dichotoma (Dictyotales, Phaeophyta). III. Ontogenese et croissance des frondes adventives. Phycologia 29:39–53.CrossRefGoogle Scholar
Gaines, S. D., and Lubchenco, J. (1982). A unified approach to marine plant-herbivore interactions. II. Biogeography. Annu. Rev. Ecol. Syst. 13: 111–38.CrossRefGoogle Scholar
Gaines, S. D., and Roughgarden, J. (1985). Larval settlement rate: a leading determinant of structure in an ecological community of the marine intertidal zone. Proc. Nat. Acad. Sci., USA 82:3707–11.CrossRefGoogle Scholar
Gaines, S. D., Lester, S. E., Eckert, G., et al. (2009). Dispersal and geographic ranges in the sea. In Witman, J. D., and Roy, K. (eds), Marine Macroecology (pp. 227–49). Chicago: University of Chicago Press.CrossRefGoogle Scholar
Gall, E. (Le Gall, Y.), Asensi, A., Marie, D., and Kloareg, B. (1996). Parthenogenesis and apospory in the Laminariales: a flow cytometry analysis. Eur. J. Phycol. 31:369–80.CrossRefGoogle Scholar
Galloway, J. N., Levy, H., and Kasibhatla, P.S. (1994). Year 2020: Consequences of population growth and development of deposition of oxidized nitrogen. Ambio 23:120–3.Google Scholar
Gansert, D., and Blossfeld, S. (2008). The application of novel optical sensors (optodes) in experimental plant ecology. Progr. Bot. 69:333–58.CrossRefGoogle Scholar
Gantt, E. (1975). Phycobilisomes: light harvesting pigment complexes. BioScience 25:781–8.CrossRefGoogle Scholar
Gantt, E., Berg, G. M., Bhattacharya, D., et al. (2010). Porphyra: complex life histories in a harsh environment: P. umbilicalis, an intertidal red alga for genomic analysis. In Seckbach, J. and Chapman, D. J. (eds), Red Algae in the Genomic Age (Cellular Origin, Life in Extreme Habitats and Astrobiology 13) (pp. 129–48). Dordrecht, The Netherlands: Springer.[, , , ]Google Scholar
Gao, K., and Umezaki, I. (l989a). Studies on diurnal photosynthetic performance of Sargassum thunbergii. I. Changes in photosynthesis under natural light. Sorui (Jpn J. Phycol.) 37:89–98.Google Scholar
Gao, K., and Umezaki, I. (l989b). Studies on diurnal photosynthetic performance of Sargassum thunbergii. II. Explanation of diurnal photosynthesis patterns from examinations in the laboratory. Sorui (Jpn J. Phycol.) 37:99–104.Google Scholar
Garbary, D. J., and Belliveau, D. J. (1990). Diffuse growth, a new pattern of cell wall deposition for the Rhodophyta. Phycologia 29:98–102.CrossRefGoogle Scholar
Garbary, D. J., and Clarke, B. (2002). Intraplant variation in nuclear DNA content in Laminaria saccharina and Alaria esculenta (Phaeophyceae). Bot. Mar. 45:211–16.CrossRefGoogle Scholar
Garbary, D. J., and Gautam, A. (1989). The Ascophyllum, Polysiphonia, Mycosphaerella symbiosis. I. Population ecology of Mycosphaerella from Nova Scotia. Bot. Mar. 32:181–6.CrossRefGoogle Scholar
Garbary, D. J., and London, J. F. (1995). The Ascophyllum / Polysiphonia / Mycosphaerella symbiosis V. Fungal infection protects A. nodosum from desiccation. Bot. Mar. 38:529–33.CrossRefGoogle Scholar
Garbary, D. J., and MacDonald, K. A. (1995). The Ascophyllum/Polysiphonia/Mycosphaerella symbiosi IV. Mutualism in the Ascophyllum/Mycosphaerella interaction. Bot. Mar. 38:221–5.CrossRefGoogle Scholar
Garbary, D. J., and McDonald, A. R. (1996). Fluorescent labeling of the cytoskeleton in Ceramium strictum (Rhodophyta). J. Phycol. 32:85–93.CrossRefGoogle Scholar
Garbary, D., Belliveau, D., and Irwin, R. (1988). Apical control of band elongation in Antithamnion defectum (Ceramiaceaae, Rhodophyta). Can. J. Bot. 66: 1308–15.CrossRefGoogle Scholar
Garbary, D. J., Kim, K. Y., Klinger, T., and Duggins, D. (1999). Red algae as hosts for endophytic kelp gametophytes. Mar. Biol. 135:35–40.CrossRefGoogle Scholar
Garbary, D. J., Lawson, G., Clement, K., and Galway, M. E. (2009). Cell division in the absence of mitosis: the unusual case of the fucoid Ascophyllum nodosum (L.) Le Jolis (Phaeophyceae). Algae. 24:239–48.CrossRefGoogle Scholar
García-Jiménez, P., Just, P. M., Delgado, A. M., and Robaina, R. R. (2007). Transglutaminase activity decrease during acclimation to hyposaline conditions in marine seaweed Grateloupia doryphora (Rhodophyta, Halymeniaceae). J. Plant Physiol. 164:367–70.CrossRefGoogle Scholar
García-Jiménez, P., García-Maroto, F., Garrido-Cárdenas, J. A., Ferrandiz, C., and Robaina, R. R. (2009). Differential expression of the ornithine decarboxylase gene during carposporogenesis in the thallus of the red seaweed Grateloupia imbricata (Halymeniaceae). J. Plant. Physiol. 166:1745–54.CrossRefGoogle Scholar
Garreta, A. G., Siguan, M. A. R., Soler, N. S., Lluch, J. R., and Kapraun, D. F. (2010). Fucales (Phaeophyceae) from Spain characterized by large-scale discontinuous nuclear DNA contents consistent with ancestral cryptopolyploidy. Phycologia 49:64–72.CrossRefGoogle Scholar
Garske, L. E. (2002). Macroalgas Marinas. In Danulat, E. and Edgar, G. J. (eds), Reserva Marina de Gálapagos. Línea Base de la Biodiversidad (pp. 419–431). Santa Cruz, Gálapagos, Ecuador: Estación Científica Charles Darwin/Servicio del Parque Nacional.Google Scholar
Gartner, A., Lavery, P., and Smit, A. J. (2002). Use of δ15N signatures of different functional forms of macroalgae and filter-feeders to reveal temporal and spatial patterns in sewage dispersal. Mar. Ecol. Prog. Ser. 235:63–73.CrossRefGoogle Scholar
Gattuso, J. P., Gentili, B., Duarte, C. M., et al. (2006). Light availability in the coastal ocean: impact on the distribution of benthic photosynthetic organisms and contribution to primary production. Biogeosci. 3:489–513.CrossRefGoogle Scholar
Gaur, J. P., and Rai, L. C. (2001). Heavy metal tolerance in algae. In Rai, L. C. and Gaur, J. P (eds), Algal Adaptation to Environmental Stresses: Physical, Biochemical and Molecular Mechanisms (pp. 363–8). New York: Springer. [, , ]CrossRefGoogle Scholar
Gaylord, B. (2000). Biological implications of surf-zone flow complexity. Limnol. Oceanogr. 45:174–88.CrossRefGoogle Scholar
Gaylord, B., and Denny, M. W. (1997). Flow and flexibility. I. Effects of size, shape and stiffness in determining wave forces on the stipitate kelps Eisenia arborea and Pterygophora californica. J. Exp. Biol. 200:3141–64.Google ScholarPubMed
Gaylord, B., Reed, D. C., Raimondi, P. T., et al. (2002). A physically based model of macroalgal spore dispersal in the wave and current-dominated nearshore. Ecology 83:1239–51.CrossRefGoogle Scholar
Gaylord, B., Rosman, J. H., Reed, D. C., et al. (2007). Spatial patterns of flow and their modification within and around a giant kelp forest. Limnol. Oceanogr. 52:1838–52.CrossRefGoogle Scholar
Gaylord, B., Denny, M. W., and Koehl, M. A. R. (2008). Flow forces on seaweeds: field evidence for roles of wave impingement and organism inertia. Biol. Bull. 215:295–308.CrossRefGoogle ScholarPubMed
Gekeler, W., Grill, E., Winnacker, E.-L., and Zenk, M. H. (1988). Algae sequester heavy metals via synthesis of phytochelatin complexes. Arch. Microbiol. 150:197–202. [)CrossRefGoogle Scholar
Genty, B., Briantais, J. M., and Baker, N. R. (1989). The relationship bretween the quantum yield of photosynthetic electron transport and quenching of chlorophyll fluorescence. Biochim. Biophys. Acta 990:87–92.CrossRefGoogle Scholar
Gerard, V. A. (1982a). Growth and utilization of internal nitrogen reserves by the giant kelp Macrocystis pyrifera in a low-nitrogen environment. Mar. Biol. 66:27–35.CrossRefGoogle Scholar
Gerard, V. A. (l982b). In situ rates of nitrate uptake by giant kelp, Macrocystis pyrifera (L.) C. Agardh: tissue differences, environmental effects, and predictions of nitrogen limited growth. J. Exp. Mar. Biol. Ecol. 62:211–24.CrossRefGoogle Scholar
Gerard, V. A. (1987). Hydrodynamic stream lining of Laminaria saccharina Lamour, in response to mechanical stress. J. Exp. Mar. Biol. Ecol. 107:237–44.CrossRefGoogle Scholar
Gerard, V. (1997). The role of nitrogen nutrition in high temperature tolerance of the kelp, Laminaria saccharina (Chromophyta). J. Phycol. 33:800–10. [, , ]CrossRefGoogle Scholar
Gerard, V. A., and Du Bois, K. R. (1988). Temperature ecotypes near the southern boundary of the kelpLaminaria saccharina. Mar. Biol. 97:575–80.CrossRefGoogle Scholar
Gerard, V. A., and Mann, K. H. (1979). Growth and production of Laminaria longicruris (Phaeophyta) populations exposed to different intensities of water movement. J. Phycol. 15:33–41.CrossRefGoogle Scholar
Gerard, V. A., Dunham, S. E., and Rosenberg, G. (1990). Nitrogen-fixation by cyanobacteria associated with Codium fragile (Chlorophyta): environmental effects and transfer of fixed nitrogen. Mar. Biol. 105:1–8.CrossRefGoogle Scholar
Gerlach, S. A. (1982). Marine Pollution: Diagnoses and Therapy. Berlin: Springer-Verlag, 218 pp.Google Scholar
Germann, I. (1988). Effects of the 1983-EI Niño on growth and carbon and nitrogen metabolism of Pleurophycus gardneri (Phaeophyceae: Laminariales) in the northeastern Pacific. Mar. Biol. 99:445–55.CrossRefGoogle Scholar
Gerwick, W. H., and Lang, N. J. (1977). Structural, chemical and ecological studies on iridescence in Iridaea (Rhodophyta). J. Phycol. 13:121–7.CrossRefGoogle Scholar
Gessner, F. (1970). Temperature: plants. In Kinne, O. (ed.), Marine Ecology (vol. I, pt. I, pp. 363–406). New York: Wiley.Google Scholar
Gessner, F., and Hammer, L. (1968). Exosmosis and “free space” in marine benthic algae. Mar. Biol. 2:88–91.CrossRefGoogle Scholar
Gessner, F., and Schramm, W. (1971). Salinity: plants. In Kinne, O. (ed.), Marine Ecology (Vol. I, pt. 2, pp. 705–820). New York: Wiley. [, , , ]Google Scholar
Gevaert, F., Barr, N. G., and Rees, T. A. V. (2007). Diurnal cycle and kinetics of ammonium assimilation in the green alga Ulva pertusa. Mar. Biol. 151:1517–24.CrossRefGoogle Scholar
Giddings, T. H., Wassman, C., and Staehelin, L.A. (1983). Structure of the thylakoids and envelope membranes of the cyanelles of Cyanophora paradoxa. Plant Physiol. 71:409–19.CrossRefGoogle ScholarPubMed
Gilman, S. E., Harley, C. D. G., Strickland, D. C., et al. (2006). Evaluation of effective shore level as a method of characterizing intertidal wave exposure regimes. Limnol. Oceanogr.: Methods 4:448–57.CrossRefGoogle Scholar
Ginger, M. L., Portman, N., and McKean, P. G. (2008). Swimming with protests: perception, motility and flagellum assembly. Nature 6:838–50.Google Scholar
Giordano, M., Beardall, J., and Raven, J. A. (2005a). CO2 concentrating mechanisms in algae: mechanisms, environmental modulation, and evolution. Ann. Rev. Plant Biol. 56:99–131.CrossRefGoogle ScholarPubMed
Giordano, M., Norici, A., and Hell, R. (2005b). Sulfur and phytoplankton: acquisition, metabolism and impact on the environment. New Phytol. 166:371–82.CrossRefGoogle ScholarPubMed
Giordano, M., Norci, A., Ratti, S., and Raven, J. A. (2008). Role of sulfur in algae: acquisition, metabolism, ecology and evolution. In Heel, R., Dahl, C., Knaff, D. B., and Leustek, T. (eds), Sulfur Metabolism in Phototrophic Organisms (pp. 397–415). Dordrecht, The Netherlands: Springer.CrossRefGoogle Scholar
Givskov, M., de Nys, R., Manefield, M., et al. (1996). Eukaryotic interference with homoserine lactone-mediated prokaryotic signaling. J. Bacteriol. 178:6618–22.CrossRefGoogle Scholar
Glass, A. D. M. (1989). Plant Nutrition: An Introduction to Current Concepts. Boston, MA: Jones and Barttell, 234 pp. [, , ]Google Scholar
Glazer, A. N. (1985). Light harvesting by phycobilisomes. Annu. Rev. Biochem. 14:47–77.Google ScholarPubMed
Gledhill, M., Nimmo, M., Hill, S. J., and Brown, M. T. (1997). The toxicity of copper(II) species to marine algae, with particular reference to macroalgae. J. Phycol. 33:2–11. [, , ]CrossRefGoogle Scholar
Gledhill, M., Nimmo, M., Hill, S. J., and Brown, M. T. (1999). The release of copper-complexing ligands by the brown alga Fucus vesiculosis (Phaeophyceae) in response to increasing total copper levels. J. Phycol. 35:501–9. [, , ]CrossRefGoogle Scholar
Gledhill, M., Brown, M. T., Nimmo, M., Moate, R., and Hill, S. J. (1998). Comparison of techniques for the removal of particulate material from seaweed tissue. Mar. Envir. Res. 45:295–307.CrossRefGoogle Scholar
Glibert, P. M., and Capone, D. G. (1993). Mineralization and assimilation in aquatic, sediment, and wetland systems. In Knowles, R. and Blackburn, R. (eds). Nitrogen Isotope Techniques (pp. 243–72). San Diego: Academic Press.CrossRefGoogle Scholar
Glibert, P. M., Lipshultz, F., McCarthy, J. J., and Altabet, M. A. (1982). Isotope dilution models of uptake and remineralization of ammonium by marine plankton. Limnol. Oceanogr. 27:639–50.CrossRefGoogle Scholar
Glynn, P. W. (1965). Community composition, structure, and interrelationships in the marine intertidal Endocladia muricata-Balanus glandula association in Monterey Bay, California. Beaufortia 12(148):1–198.Google Scholar
Glynn, P. W. (1988). El Niño-Southern Oscillation 1982–1983: nearshore population, community and ecosystem responses. Annu. Rev. Ecol. Syst. 19:309–45.CrossRefGoogle Scholar
Goecke, F., Labes, A., Weise, J., et al. (2010). Chemical interactions between marine macroalgae and bacteria. Mar. Ecol. Prog. Ser. 409:267–99.CrossRefGoogle Scholar
Goff, L. J. (1982). Biology of parasitic red algae. Prog. Phycol. Res. 1:289–369.Google Scholar
Goff, L. J., and Coleman, A. W. (l988). The use of plastid DNA restriction endonuclease patterns in delimiting red algal species and populations. J. Phycol. 24:357–68.Google Scholar
Goff, L. J., and Coleman, A. W. (1990). Red algal plasmids. Curr. Genet. 18:557–65.CrossRefGoogle ScholarPubMed
Goff, L. J., and Coleman, A. W. (1995). Fate of parasite and host organelle DNA cellular-transformation of red algae by their parasites. Plant Cell 7:1899–911.CrossRefGoogle ScholarPubMed
Goff, L. J., and Zuccarello, G. (1994). The evolution of parasitism in red algae: cellular interactions of adelphoparasites and their hosts. J. Phycol. 30:695–720.CrossRefGoogle Scholar
Goff, L. J., Moon, D. A., Nyvall, P., et al. (1996). The evolution of parasitism in the red algae: molecular comparisons of adelphoparasites and their hosts. J. Phycol. 32:297–312.CrossRefGoogle Scholar
Goff, L. J., Ashen, J., and Moon, D. (1997). The evolution of parasites from their hosts: a case study in the parasitic red algae. Evolution 51:1068–78.CrossRefGoogle ScholarPubMed
Gómez, I., and Huovinen, P. (2011). Morpho-functional patterns and zonation of South Chilean seaweeds: the importance of photosynthetic and bio-optical traits. Mar. Ecol. Progr. Ser. 422:77–91.CrossRefGoogle Scholar
Gómez, I., and Huovinen, P. (2012). Morpho-functionality of carbon metabolism in seaweeds. In Wiencke, C. and Bischof, K. (eds) Seaweed Biology: Novel Insights into Ecophysiology, Ecology and Utilization, Ecological Studies (Vol. 219, pp. 25–45). Berlin: Springer-Verlag.CrossRefGoogle Scholar
Gómez, I., and Wiencke, C. (1998). Seasonal changes in C, N and major organic compounds and their significance to morpho-functional processes in the endemic Antarctic brown alga Ascoseira mirabilis. Polar Biol. 19:115–24.Google Scholar
Gómez, I., Figueroa, F. L., Sousa-Pinto, I., et al. (2001). Effects of UV radiation and temperature on photosynthesis as measured by PAM fluorescence in the red alga Gelidium pulchellum (Turner) Kützing. Bot. Mar. 44:9–16.CrossRefGoogle Scholar
Gómez, I., Ulloa, N., and Orostegui, M. (2005). Morpho-functional patterns of photosynthesis and UV sensitivity in the kelp Lessonia nigrescens (Laminariales, Phaeophyta). Mar. Biol. 148:231–40.CrossRefGoogle Scholar
Gómez, I., Orostegui, M., and Huovinen, P. (2007) Morpho-functional patterns of photosynthesis in the South Pacific kelp Lessonia nigrescens: effects of UV radiation on 14C fixation and primary photochemical reactions. J. Phycol. 43:55–64.CrossRefGoogle Scholar
Gómez, I., Wulff, A., Roleda, M. Y., et al. (2009). Light and temperature demands of marine benthic micro-algae and seaweeds in the polar regions. Bot. Mar. 52:593–608.CrossRefGoogle Scholar
Gonen, Y., Kimmel, E., and Friedlander, M. (1995). Diffusion boundary layer transport in Gracilaria conferta (Rhodophyta). J. Phycol. 31:768–73.CrossRefGoogle Scholar
Gonen, Y., Kimmel, E., Tel-Or, E., and Friedlander, M. (1996). Intercellular assimilate translocation in Gracilaria cornea (Gracilariaceae, Rhodophyta). Hydrobiologia 326/327:421–8.CrossRefGoogle Scholar
Goodwin, T. W., and Mercer, E. I. (1983). Introduction to Plant Biochemisty, 2nd edn. Oxford: Pergamon.Google Scholar
Gordillo, F. J. L. (2012). Environment and algal nutrition. In Weincke, C. and Bischoff, K. (eds) Seaweed Biology: Novel Insights into Ecophysiology and Utilization (pp. 67–86). Ecology Study Series. Berlin: Springer.CrossRefGoogle Scholar
Gordillo, F. J., Aguilera, J., and Jimenez, C. (2006). The response of nutrient assimilation and biochemical composition of Arctic seaweeds to a nutrient input in summer. J. Exp. Bot. 57:2661–71.CrossRefGoogle Scholar
Gordon, D. M., Birch, P. B., and McComb, A. J. (1980). The effects of light, temperature and salinity on photosynthetic rates of an estuarine Cladophora. Bot. Mar. 29:749–55.Google Scholar
Gordon, D. M., Birch, P. B., and McComb, A. J. (1981). Effects of inorganic phosphorus and nitrogen on the growth of an estuarine Cladophora in culture. Bot. Mar. 24:93–106.CrossRefGoogle Scholar
Gordon, R., and Brawley, S. H. (2004). Effects of water motion on propagule release from algae with complex life histories. Mar. Biol. 145:21–9.CrossRefGoogle Scholar
Gorman, D., and Connell, S. D. (2009). Recovering subtidal forests in human-dominated landscapes. J. Appl. Ecol. 46:1258–65.CrossRefGoogle Scholar
Gorospec, K. D., and Karl, S. A. (2011). Small-scale spatial analysis of in situ sea temperature throughout a single coral patch reef. J. Mar. Biol. Article ID 719580. 12 pp.
Graeve, M., Hagen, W., and Kattner, G. (1994). Herbivorous or omnivorous? On the significance of lipid compositions as trophic markers in Antarctic copepods. Deep Sea Res. I 41: 915–24.CrossRefGoogle Scholar
Graeve, M., Kattner, G., Wiencke, C., and Karsten, U. (2002). Fatty acid composition of Arctic and Antarctic macroalgae: indicator of phylogenetic and trophic relationships. Mar. Ecol. Prog. Ser. 231:67–74. [, , ]CrossRefGoogle Scholar
Graham, L. E., Graham, J. M., and Wilcox, L. W. (2009). Algae. London: Pearson College Division, 616 pp. [, , , , , , , , , ]Google Scholar
Graham, M. H. (1997). Factors determining the upper limit of giant kelp, Macrocystis pyrifera Agardh, along the Monterey Peninsula, central California, USA. J. Exp. Mar. Biol. Ecol. 218:127–49.CrossRefGoogle Scholar
Graham, M. H. (1999). Identification of kelp zoospores from in situ plankton samples. Mar. Biol. 135:709–20.CrossRefGoogle Scholar
Graham, M. H. (2002). Prolonged reproductive consequences of short-term biomass loss in seaweeds. Mar. Biol. 140:901–11.Google Scholar
Graham, M. H., Vásquez, J. A., and Buschmann, A. H. (2007a). Global ecology of the giant kelp Macrocystis from ecotypes to ecosystems. Oceanogr. Mar. Biol. 45:39–88. [, , , , , , ]Google Scholar
Graham, M. H., Kinlan, B. P., Druehl, L. D., Garske, L. E., and Banks, S. (2007b). Deep-water kelp refugia as potential hotspots of tropical marine diversity and productivity. PNAS 104:16,576–80.CrossRefGoogle ScholarPubMed
Graham, M. H., Kinlan, B. P., and Grosberg, R. K. (2010). Post-glacial redistribution and shifts in productivity of giant kelp forests. Proc. Royal Soc. B 277:399–406.CrossRefGoogle ScholarPubMed
Grall, J., and Chauvaud, L. (2002). Marine eutrophication and benthos: the need for new approaches and concepts. Global Change Biol. 8:813–30.CrossRefGoogle Scholar
Granbom, L. M., Pedersén, M., Kadel, P., and Lüning, K. (2001). Circadian rhythm of photosynthetic oxygen evolution in Kappaphycus alvarezii (Rhodophyta): dependence on light quantity and quality. J. Phycol. 37:1020–5.CrossRefGoogle Scholar
Granhag, M., Larsson, A. I., and Jonsson, P. R. (2007). Algal spore settlement and germling removal as a function of flow speed. Mar. Ecol. Prog. Ser. 344:63–9.CrossRefGoogle Scholar
Grant, A. J., Trautman, D. A., Menz, I., et al. (2006). Separation of two cell signaling molecules from a symbiotic sponge that modify algal carbon metabolism. Biochem. Bioph. Res. Co. 348:92–8.CrossRefGoogle ScholarPubMed
Grant, B. R., and Borowitzka, M. A. (1984). The chloroplasts of giant-celled and coenocytic algae: biochemistry and structure. Bot. Rev. 50:267–307.CrossRefGoogle Scholar
Gravel, D., Guichard, F., Loreau, M., and Mouquet, N. (2010a). Source and sink dynamics in meta-ecosystems. Ecology 91:2172–84.CrossRefGoogle ScholarPubMed
Gravel, D., Mouquet, N., Loreau, M., and Guichard, F. (2010b). Patch dynamics, persistence, and species coexistence in metaecosystems. Am. Nat. 176:289–302.CrossRefGoogle ScholarPubMed
Gravot, A., Dittami, S. M., Rousvoal, S., et al. (2010). Diurnal oscillations of metabolite abundances and gene analysis provide new insights into central metabolic processes of the brown alga Ectocarpus siliculosus. New Phytol. 188:98–110.CrossRefGoogle ScholarPubMed
Gray, J. S. (2002). Biomagnification in marine systems: the perspective of an ecologist. Mar. Poll. Bull. 45:46–52.CrossRefGoogle ScholarPubMed
Green, B. R., and Durnford, D. G. (1996). The chlorophyll-carotenoid proteins of oxygenic photosynthesis. Ann. Rev. Plant Physiol. Plant Mol. Biol. 47:685–714.CrossRefGoogle ScholarPubMed
Greene, R. M., and Gerard, V. A. (1990). Effects of high frequency light fluctuations on growth and photoacclimation of the red alga Chondrus crispus. Mar. Biol. 105:337–44.CrossRefGoogle Scholar
Greene, R. W. (1970). Symbiosis in sacoglossan opisthobranchs: functional capacity of symbiotic chloroplasts. Mar. Biol. 72:138–42.CrossRefGoogle Scholar
Greer, S. P., and Amsler, C. D. (2004). Clonal variation in phototaxis and settlement behaviors of Hincksia irregularis (Phaephyceae) spores. J. Phycol. 40:44–53.CrossRefGoogle Scholar
Gregory, T. R. (2005). The C-value enigma in plants and animals: a review of parallels and an appeal for partnership. Ann. Bot.-London 95:133–46.CrossRefGoogle ScholarPubMed
Gressler, V., Colepicolo, P., and Pinto, E. (2009). Useful strategies for algal volatile analysis. Curr. Analyt. Chem. 5:271–92.CrossRefGoogle Scholar
Grime, J.P (1979). Plant Strategies and Vegetation Processes. New York: Wiley.Google Scholar
Groen, P. (1980). Oceans and seas. I. Physical and chemical properties. In The New Encyclopaedia Britannica (vol. 13, pp. 484–97). London: MacropaediaGoogle Scholar
Gross, M. G. (1996). Oceanography. New York: Prentice Hall, 236 pp.Google Scholar
Gross, W. (1990). Occurrence of glycolate oxidase and hydropyruvate reductase in Egreggia menziesii (Phaeophyta). J. Phycol. 26:381–3.CrossRefGoogle Scholar
Grossman, A. (2005). Regeneration of a cell from protoplasm. J. Phycol. 42:1–5.CrossRefGoogle Scholar
Grossman, A. R., Anderson, L. K., Conley, P. B., and Lemaux, P. G. (1989). Molecular analyses of complementary chromatic adaptation and the biosynthesis of a phycobilisome. In Coleman, A. W., Goff, L. J., and Stein-Taylor, J. R. (eds) Algae as Experimental Systems (pp. 269–88). New York: Alan R. Liss.Google Scholar
Grzymski, J., Johnsen, G., and Sakshaug, E. (1997). The significance of intracellular self-shading on the bio-optical properties of brown, red and green macroalgae. J. Phycol. 33:408–14.CrossRefGoogle Scholar
Gualtieri, P., and Robinson, K. R. (2002). A rhodopsin-like protein in the plasma membrane of Silvetia compressa eggs. Photochem. Photobiol. 75:76–8.2.0.CO;2>CrossRefGoogle ScholarPubMed
Guerry, A. D. (2006). Grazing, nutrients, and marine benthic algae: insights into the drivers and protection of diversity. PhD Dissertation, Oregon State University.
Guerry, A. D., Menge, B. A., and Dunmore, R. A. (2009). Effects of consumers and enrichment on abundance and diversity of benthic algae in a rocky intertidal community. J. Exp. Mar. Biol. Ecol. 369:155–64.CrossRefGoogle Scholar
Guest, M. A., Frusher, S. D., Nichols, P. D., Johnson, C. R., and Wheatley, K. E. (2009). Trophic effects of fishing southern rock lobster Jasus edwardsii shown by combined fatty acid and stable isotope analyses. Mar. Ecol. Prog. Ser. 388:169–84.CrossRefGoogle Scholar
Guillemin, M. L., Faugeron, S., Destombe, C., et al. (2008). Genetic variation in wild and cultivated populations of the haploid-diploid red alga Gracilaria chilensis: how farming practices favor asexual reproduction and heterozygosity. Evolution 62:1500–19.CrossRefGoogle ScholarPubMed
Guimarães, S. M. P. B., Broga, M. R. A., Cordeiro-Marino, M., and Pedrini, A. G. (1986). Morphology and taxonomy of Jolyna laminarioides, a new member of the Scytosiphonales (Phaeophyceae) from Brazil. Phycologia 25:99–108.CrossRefGoogle Scholar
Guiry, M. (2011). Available online: .
Guiry, M.D. (1974). A preliminary consideration of the taxonomic position of Palmaria palmata (Linnaeus) Stackhouse = Rhodymenia palmata (Linnaeus) Greville. J. Mar. Biol. Assoc. UK 54:509–28.CrossRefGoogle Scholar
Guiry, M. D. (1978). The importance of sporangia in the classification of the Florideophycidae. In Irvine, D. E. G. and Price, J. H. (eds), Modern Approaches to the Taxonomy of Red and Brown Algae (pp. 111–44). London: Academic Press.Google Scholar
Guiry, M. D. (1990). Sporangia and spores. In Cole, K. M. and Sheath, R. G. (eds), Biology of the Red Algae (pp. 43–71). Cambridge: Cambridge University Press.Google Scholar
Gundlach, E., and Hayes, M. (1978). Vulnerability of coastal environments to oil spill impacts. Mar. Technol. Soc. J. 12:1827.Google Scholar
Gundlach, E. R., Boehm, P. D., Marchand, M., et al. (1983). The fate of Amoco. Cadiz oil. Science 221:122–9. [, , ]CrossRefGoogle ScholarPubMed
Haavisto, F., Välikangas, T., and Jormalainen, V. (2010). Induced resistance in a brown alga: phlorotannins, genotypic variation and fitness costs for the crustacean herbivore. Oecologia 162:685–95.CrossRefGoogle Scholar
Hable, W. E., and Kropf, D. L. (2000). Sperm entry induces polarity in fucoid zygotes. Development 127:493–501.Google ScholarPubMed
Hable, W. E., Miller, N. R., and Kropf, D. L. (2003). Polarity establishment requires dynamic actin in fucoid zygotes. Protoplasma 221:193–204.Google ScholarPubMed
Hackney, J. M., and Sze, P. (1988). Photorespiration and productivity rates of a coral reef algal turf assemblage. Mar. Biol. 98:483–92.CrossRefGoogle Scholar
Hackney, J. M., Carpenter, R. C., and Adey, W. H. (1989). Characteristic adaptations to grazing among algal turfs on a Caribbean coral reef. Phycologia 28: 109–19.CrossRefGoogle Scholar
Häder, D. P., Kumar, H. D., Smith, R. C., and Worrest, R. C. (1998). Effects on aquatic ecosystems. J. Photochem. Photobiol. B: Biol. 46:53–68.CrossRefGoogle Scholar
Hafting, J. F. (1999). A novel technique for propagation of Porphyra yezoensis Ueda blades in suspension cultures via monospores. J. Appl. Phycol. 11:361–7.CrossRefGoogle Scholar
Hagen, N. T. (1995). Recurrent destructive grazing of successionally immature kelp forests by green sea urchins in vestfjorden, Nothern Norway. Mar. Ecol. Prog. Ser. 123:95–106.CrossRefGoogle Scholar
Hagopian, J. C., Reis, M., Kitajima, J. P., et al. (2004). Comparative analysis of the complete plastid genome sequence of the red alga Gracilaria tenuistipitata var. liui provides insights into the evolution of rhodoplasts and their relationship to other plastids. J. Mol. Evol. 59:464–77.CrossRefGoogle ScholarPubMed
Haines, K. C., and Wheeler, P. A. (1978). Ammonium and nitrate uptake by the marine macrophyte Hypnea musciformis (Rhodophyta) and Macrocystis pyrifera (Phaeophyta). J. Phycol. 14:319–24.CrossRefGoogle Scholar
Hairston, N. G., Smith, F. E., and Slobodkin, L. B. (1960). Community structure, population control, and competition. Am. Nat. 94:421–5.CrossRefGoogle Scholar
Hall, J. D., and Murray, S. N. (1998). The life history of a Santa Catalina Island population of Liagora californica (Nemaliales, Rhodophyta) in the field and in laboratory culture. Phycologia 37:184–94.CrossRefGoogle Scholar
Hall, J. L. (2002). Cellular mechanisms for heavy metal detoxification and tolerance. J. Exp. Bot. 53(386):1–11.CrossRefGoogle ScholarPubMed
Hall, Jr. L.W., Giddings, J. M., Solomon, K. R., and Balcomb, R. (1999). An ecological assessment for the use of Irgarol 1051 an an algacide for antifouling paints. Crit. Rev. Toxicol. 29:367–437.Google Scholar
Hall, L. W., and Pinkney, A. E. (1984). Acute and sublethal effects of organotin compounds on aquatic biota: an interpretive literature evaluation. CRC Crit. Rev. Toxicol. 14:159–209.CrossRefGoogle Scholar
Halldal, P. (1964). Ultraviolet action spectra of photosynthesis and photosynthetic inhibition in a green alga and a red alga. Physiol. Plant 17:414–21.CrossRefGoogle Scholar
Hall-Spencer, J. M., Rodolfo-Metalpa, R., Martin, S., et al. (2008). Volcanic carbon dioxide vents show ecosystem effects of ocean acidification. Nature 354:96–9.CrossRefGoogle Scholar
Halpern, B. S., Walbridge, S., Selkoe, K. A., et al. (2008) A global map of human impact on marine ecosystems. Science 319:948–52.CrossRefGoogle ScholarPubMed
Hamilton, J. G., Zangerl, A. R., DeLucia, E. H., and Berenbaum, M. R. (2001). The carbon-nutrient balance hypothesis: its rise and fall. Ecol. Lett. 4:86–95.CrossRefGoogle Scholar
Han, T., and Choi, G-W. (2005). A novel marine algal toxicity bioassay based on sporulation inhibition in the green macroalga Ulva pertusa (Chlorophyta). Aquat. Toxicol. 75:202–12.CrossRefGoogle Scholar
Han, T., Han, Y.-S., Kain, J. M., and Häder, D.-P. (2003). Thallus differentiation of photosynthesis, growth, reproduction and UV-B sensitivity in the green alga Ulva pertusa (Chlorophyceae). J. Phycol. 39:712–21.CrossRefGoogle Scholar
Han, T., Han, Y. S, Park, C. Y, et al. (2008). Spore release by the green alga Ulva: A quantitative assay to evaluate aquatic toxicants. Environ. Poll. 153:699–705.CrossRefGoogle ScholarPubMed
Han, T., Kong, J. A., and Brown, M. T. (2009). Aquatic toxicity tests of Ulva pertusa Kjellman (Ulvales, Chlorophyta) using spore germination and gametophyte growth. Eur. J. Phycol. 44:357–63.CrossRefGoogle Scholar
Han, Y. S., Brown, M. T., Park, G. S., and Han, T. (2007). Evaluating aquatic toxicity by visual inspection of thallus color in the green macroalga Ulva: Testing a novel bioassay. Environ. Sci. Technol. 47:3667–71.CrossRefGoogle Scholar
Hanelt, D. (1998). Capability of dynamic photoinhibition in Arctic macroalgae is related to their depth distribution. Mar. Biol. 131:361–9.CrossRefGoogle Scholar
Hanelt, D., and Nultsch, W. (1990). Daily changes of the phaeoplast arrangement in the brown alga Dictyota dichotoma as studied in field experiments. Mar. Ecol. Prog. Ser. 61:273–9.CrossRefGoogle Scholar
Hanelt, D., and Nultsch, W. (1991). The role of chromatophore arrangement in protecting the chromatophores of the brown alga Dictyota dichotoma against photodamage. J. Plant Physiol. 138:470–5.CrossRefGoogle Scholar
Hanelt, D., and Nultsch, W. (1995). Field studies of photoinhibition show non-correlations between oxygen and fluorescence measurements in the Arctic red alga Palmaria palmata. J. Plant Physiol. 145:31–8.CrossRefGoogle Scholar
Hanelt, D., and Nultsch, W. (2003). Photoinhibition in seaweeds. In Heldmaier, G. and Werner, D. (eds) Environmental Signal Processing and Adaptation (pp. 1414–67). Heidelberg and Berlin: Springer Publishers.Google Scholar
Hanelt, D., and Roleda, M. Y. (2009). UVB radiation may ameliorate photoinhibition in specific shallow-water tropical marine macrophytes. Aquat. Bot. 91:6–12.CrossRefGoogle Scholar
Hanelt, D., Tüg, H., Bischof, K., et al. (2001) Light regime in an Arctic fjord: a study related to stratospheric ozone depletion as a basis for determination of UV effects on algal growth. Mar. Biol. 138:649–58.CrossRefGoogle Scholar
Hanelt, D., Wiencke, C., and Bischof, K. (2003). Photosynthesis in marine macroalgae. In Larkum, A. W. D., Douglas, S. E., and Raven, J. A. (eds) Photosynthesis in Algae. Advances in Photosynthesis and Respiration (Vol. 14, pp. 413–35). Dordrecht, The Netherlands: Kluwer Academic Publishers.CrossRefGoogle Scholar
Hanelt, D., Hawes, I., and Rae, R. (2006). Reduction of UV-B radiation causes an enhancement of photoinhibition in high light stressed aquatic plants from New Zealand lakes. J. Photochem. Photobiol. B: Biol. 84:89–102.CrossRefGoogle ScholarPubMed
Hanic, L. A., and Craigie, J. S. (1969). Studies on the algal cuticle. J. Phycol. 5:89–109.CrossRefGoogle ScholarPubMed
Hanisak, M. D. (1979). Nitrogen limitation of Codium fragile ssp. tomentosoides as determined by tissue analysis. Mar. Biol. 50:333–7.CrossRefGoogle Scholar
Hanisak, M. D. (1983). The nitrogen relationships of marine macroalgae. In Carpenter, E. J. and Capone, D. G. (eds). Nitrogen in the Marine Environment (pp. 699–730). New York: Academic Press.CrossRefGoogle Scholar
Hanisak, M. D. (1990). The use of Gracilaria tikvahiae (Gracilariales, Rhodophyta) as a model system to understand the nitrogen limitation of cultured seaweeds. Hydrobiologia 204/205:79–87.CrossRefGoogle Scholar
Hanisak, M. D., and Harlin, M. M. (1978). Uptake of inorganic nitrogen by Codium fragile subsp. tomentosoides (Chlorophyta). J. Phycol. 14:450–4.CrossRefGoogle Scholar
Hanisak, M. D., Littler, M. M., and Littler, D. S. (1988). Significance of macroalgal polymorphism: intraspecific tests of the functional-form model. Mar Biol 99:157–65.CrossRefGoogle Scholar
Hansen, A. T., Hondzo, M., and Hurd, C. L. (2011). Photosynthetic oxygen flux by Macrocystis pyrifera: a mass transfer model with experimental validation. Mar. Ecol. Prog. Ser. 434:45–55.CrossRefGoogle Scholar
Hanson, C. E., Hyndes, G. A., and Wang, S. F. (2010). Differentiation of benthic marine primary producers using stable isotopes and fatty acids: implications to food web studies. Aquat Bot 93: 114–22.CrossRefGoogle Scholar
Harder, D. L., Speck, O., Hurd, C. L., and Speck, T. (2004b). Reconfiguration as a pre-requisite for survival in highly unstable flow-dominated habitats. J. Plant Growth Regul. 23:98–107.CrossRefGoogle Scholar
Harder, T., Dobretsov, S., and Qian, P.-Y. (2004a). Waterborne polar macromolecules act as algal antifoulants in the seaweed Ulva reticulata. Mar. Ecol. Prog. Ser. 274:133–41.CrossRefGoogle Scholar
Hardy, F. G., and Moss, B. L. (1979). Attachment and development of the zygotes of Pelvetia canaliculata (L.) Dcne. et Thur. (Phaeophyceae, Fucales). Phycologia 18:203–12.CrossRefGoogle Scholar
Harley, C. D. G. (2003). Abiotic stress and herbivory interact to set range limits across a two-dimensional stress gradient. Ecology 84:1477–88.CrossRefGoogle Scholar
Harley, C. D., and Helmuth, B. S. T. (2003). Local- and regional-scale effects of wave exposure, thermal stress, and absolute versus effective shore level on patterns of intertidal zonation. Limnol. Oceanogr. 48:1498–508.CrossRefGoogle Scholar
Harlin, M. M. (1978). Nitrate uptake by Enteromorpha spp. (Chlorophyceae): applications to aquaculture systems. Aquaculture 15:373–6.CrossRefGoogle Scholar
Harlin, M. M. (1987). Allelochemistry in marine macroalgae. CRC Crit. Rev. Plant Sci. 5:237–49.CrossRefGoogle Scholar
Harlin, M. M., and Craigie, J. S. (1975). The distribution of photosynthate in Ascophyllum nodosum as it relates to epiphytic Polysiphonia lanosa. J. Phycol. 11:109–13.Google Scholar
Harlin, M. M., and Craigie, J. S. (1978). Nitrate uptake by Laminaria longicruris (Phaeophyceae). J. Phycol. 14:464–7.CrossRefGoogle Scholar
Harlin, M. M., and Wheeler, P. A. (1985). Nutrient uptake. In Littler, M. M. and Littler, D. S. (eds), Handbook of Phycological Methods: Ecological Field Methods: Macroalgae (pp. 493–508). Cambridge: Cambridge University Press.Google Scholar
Harmer, S. L. (2009). The circadian system in higher plants. Annu. Rev. Plant Biol. 60:357–77.CrossRefGoogle ScholarPubMed
Harper, J. L. (1977). Population Biology of Plants. New York: Academic Press, 892 pp.Google Scholar
Harpole, W. S., Ngai, J. T., Cleland, E. E., et al. (2011). Nutrient co-limitation of primary producers. Ecol. Lett. 14:852–62.CrossRefGoogle Scholar
Harrington, L., Fabricius, K., De’Ath, G., and Negri, A. (2004). Recognition and selection of settlement substrata determine post-settlement survival in corals. Ecology 85:3428–37.CrossRefGoogle Scholar
Harrington, L., Fabricius, K. E., and Negri, A. (2005). Synergistic effects of diuron and sedimentation on the photophysiology and survival of crustose coralline algae. Marine Pollution Bulletin 51:415–27.CrossRefGoogle ScholarPubMed
Harris, L. G., and Jones, A. C. (2005). Temperature, herbivory and epibiont acquisition as factors controlling the distribution and ecological role of an invasive seaweed. Biol. Invas. 7:913–24.CrossRefGoogle Scholar
Harrison, P J., and Druehl, L. D. (1982). Nutrient uptake and growth in the Laminariales and other macrophytes: a consideration of methods. In Srivastava, L. M. (ed.), Synthetic and Degradative Processes in Marine Macrophytes (pp. 99–120). Berlin: Walter de Gruyter.Google Scholar
Harrison, P. J., and Hurd, C. L. (2001). Nutrient physiology of seaweeds: application of concepts to aquaculture. Cah. Biol. Mar. 42:71–82.Google Scholar
Harrison, P. J., Druehl, L. D., Lloyd, K. E., and Thompson, P. A. (1986). Nitrogen uptake kinetics in three year-classes of Laminaria groenlandica (Laminariales: Phaeophyta). Mar. Biol. 93:29–35.CrossRefGoogle Scholar
Harrison, P. J., Parslow, J. S., and Conway, H. L. (1989). Determination of nutrient uptake kinetic parameters: a comparison of methods. Mar. Ecol. Prog. Ser. 52:301–12. [, , ]CrossRefGoogle Scholar
Harrison, P. J., Hu, M. H., Yang, Y. P., and Lu, X. (1990). Phosphate limitation in estuarine and coastal waters of China. J. Exp. Mar. Biol. Ecol. 140:79–87.CrossRefGoogle Scholar
Hartnoll, R. G., and Hawkins, S. J. (1982). The emersion curve in semidiurnal tidal regimes. Estuar. Coast. Shelf S. 15:365–71.CrossRefGoogle Scholar
Haslin, C., Lahaye, M., Pellegrini, M., and Chermann, J. C. (2001). In vitro anti-HIV activity of sulfated cell wall polysaccharides from gametic, carposporic and tetrasporic stages of the Mediterranean red alga Asparagopsis armata. Planta Med 67: 301–5.CrossRefGoogle ScholarPubMed
Hata, H., and Kato, M. (2004). Monoculture and mixed-species algal farms on a coral reef are maintained through intensive and extensive management by damselfishes. J. Exp. Mar. Biol. Ecol. 313:285–96.CrossRefGoogle Scholar
Hata, H., and Kato, M. (2006). A novel obligate cultivation mutualism between damselfish and Polysiphonia algae. Biol. Lett. 2:593–6.CrossRefGoogle ScholarPubMed
Hatcher, B. G. (1977). An apparatus for measuring photosynthesis and respiration of intact large marine algae and comparison of results with those from experiments with tissue segments. Mar. Biol. 43:381–5.CrossRefGoogle Scholar
Hatcher, B. G., Chapman, A. R. O., and Mann, K. H. (1977). An annual carbon budget for the kelp Laminaria longicruris. Mar Biol 44: 85–96. [, , ]CrossRefGoogle Scholar
Haug, A. (1976). The influence of borate and calcium on the gel formation of a sulfated polysaccharide from UIva lactuca. Acta Chem. Scand. B30:562–6.CrossRefGoogle Scholar
Haug, A., and Larsen, B. (1974). Biosynthesis of algal polysaccharides. In Pridham, J. B. (ed), Plant Carbohydrate Biochemistry (pp. 207–18). New York: Academic Press.Google Scholar
Hauri, C., Fabricius, K. E., Schaffelke, B., and Humphrey, C. (2010). Chemical and physical environmental conditions underneath mat- and canopy-forming macroalgae, and their effects on understorey corals. PLoS ONE 5:e12685.CrossRefGoogle ScholarPubMed
Hawkes, M. W. (1990). Reproductive strategies. In Cole, K. M. and Sheath, R. G. (eds), Biology of the Red Algae (pp. 455–76). Cambridge: Cambridge University Press.Google Scholar
Hawkins, S. J., and Southward, A. J. (1992). The Torrey Canyon oil spill: recovery of rock shore communities. In Thayer, G.W. (ed.). Restoring the Nation’s Environment (pp. 583–631). College Park, MD: Maryland Sea Grant Coll.Google Scholar
Hawkins, S. J., Sugden, H. E., Mieszkowskz, N., et al. (2009). Consequences of climate-driven biodiversity changes for ecosystem functioning of North European rocky shores. Mar. Ecol. Prog. Ser. 396:245–59.CrossRefGoogle Scholar
Haxen, P.G., and Lewis, O. A. M. (1981). Nitrate assimilation in the marine kelp, Macrocystis angustifolia (Phaeophyceae). Bot. Mar. 24:631–5.Google Scholar
Haxo, F. T., and Blinks, L. R. (1950). Photosynthetic action spectra of marine algae. J. Gen. Physiol. 33:389–425.CrossRefGoogle ScholarPubMed
Hay, M. E. (l981a). The functional morphology of turf-forming seaweeds: persistence in stressful marine habitats. Ecology 62:739–50.CrossRefGoogle Scholar
Hay, M. E. (l981b). Spatial patterns of grazing intensity on a Caribbean barrier reef: herbivory and algal distribution. Aquat. Bot. 11: 97–109.CrossRefGoogle Scholar
Hay, M. E. (1986). Functional geometry of seaweeds: ecological consequences of thallus layering and shape in contrasting light environments. In: Givnish, T. J. (ed) On the Economy of Plant Form and Function (pp. 635–66). Cambridge: Cambridge University Press.Google Scholar
Hay, M. E. (1988). Associational plant defenses and the maintenance of species diversity: turning competitors into accomplices. Am. Nat. 128:617–41.CrossRefGoogle Scholar
Hay, M. E. (2009). Marine chemical ecology: chemical signals and cues structure marine populations, communities, and ecosystems. Annu. Rev. Mar. Sci. 1:193–212.CrossRefGoogle ScholarPubMed
Hay, M. E., Parker, J. D., Burkepile, D. E., et al. (2004). Mutualisms and aquatic community structure: the enemy of my enemy is my friend. Annu. Rev. Ecol. Evol. Syst. 35:175–97.CrossRefGoogle Scholar
Hay, M. E., Stachowicz, J. J., Cruz-Rivera, E., et al. (1998). Bioassays with marine and freshwater macroorganisms. In Haynes, K. F. and Millar, J. G. (eds), Methods in Chemical Ecology, Vol. 2: Bioassay Methods (pp. 39–141). New York: Chapman and Hall.Google Scholar
Hay, M. E., Kappel, Q. E., and Fenical, W. (1994). Synergisms in plant defenses against herbivores: interactions of chemistry, calcification, and plant quality. Ecology 75:1714–26.CrossRefGoogle Scholar
Hay, M. E., Paul, V. J., Lewis, S. M., et al. (l988). Can tropical seaweeds reduce herbivory by growing at night? Diel patterns of growth, nitrogen content, herbivory, and chemical versus morphological defenses. Oecologia 75:233–45.CrossRefGoogle Scholar
Hays, C. G. (2007). Adaptive phenotypic differentiation across the intertidal gradient in the alga Silvetia compressa. Ecology 88:149–57.CrossRefGoogle ScholarPubMed
He, P., and Yarish, C. (2006). The developmental regulation of mass cultures of free-living conchocelis for commercial net seeding of Porphyra leucosticta from North America. Aquaculture 257:373–81.CrossRefGoogle Scholar
He, P., Xu, S., Zhang, H., et al. (2008). Bioremediation efficiency in the removal of dissolved inorganic nutrients by the red seaweed, Porphyra yezoensis, cultivated in the open sea. Water Res. 42:1281–9.CrossRefGoogle ScholarPubMed
Heaton, T. H. E. (1986). Isotopic studies of nitrogen pollution in the hydrosphere and atmosphere: a review. Chem Geol. 59:87–102.CrossRefGoogle Scholar
Hebert, P.D.N., Cywinska, A., Ball, S. L., and deWaard, J. R. (2003). Biological identifications through DNA barcodes. Proc. R. Soc. Lond. B 270: 313–22.CrossRefGoogle ScholarPubMed
Heesch, S., Peters, A. F., Broom, J. E., and Hurd, C. L. (2008). Affiliation of the parasite Herpodiscus durvillaeae (Phaeophyceae) with the Sphacelariales based on DNA sequence comparisons and morphological observations. Eur. J. Phycol. 43:283–95.CrossRefGoogle Scholar
Heesch, S., Cho, G. Y., Peters, A. F., et al. (2010). A sequence-tagged genetic map for the brown alga Ectocarpus siliculosus provides large-scale assembly of the genome sequence. New Phytol. 188:42–51.CrossRefGoogle ScholarPubMed
Hegemann, P. (2008). Algal sensory photoreceptors. Annu. Rev. Plant Biol. 59:167–89. [, , ]CrossRefGoogle ScholarPubMed
Hein, M., Pedersen, M. F., and Sand-Jensen, K. (1995). Size-dependent nitrogen uptake in micro- and macroalgae. Mar. Ecol. Prog. Ser. 118:247–53. [, , ]CrossRefGoogle Scholar
Hellebust, J. A. (1976). Osmoregulation. Annu. Rev. Plant Physiol. 27: 485–505.CrossRefGoogle Scholar
Helmuth, B., and Denny, M. W. (2003). Predicting wave exposure in the rocky intertidal zone: do bigger waves always lead to larger forces?. Limnol. Oceanogr. 48:1338–45.CrossRefGoogle Scholar
Helmuth, B., Mieszkowska, N., Moore, P., and Hawkins, S. J. (2006). Living on the edge of two changing worlds: forecasting the responses of rocky intertidal ecosystems to climate change. Annu. Rev. Ecol. Evol. Syst. 37:373–404.CrossRefGoogle Scholar
Hemmi, A., Honkanen, T., and Jormalainen, V. (2004). Inducible resistance to herbivory in Fucus vesiculosus: duration, spreading and variation with nutrient availability. Mar. Ecol. Prog. Ser. 273:109–20.CrossRefGoogle Scholar
Hemminga, M. A., and Duarte, C. M. (eds) (2000). Seagrass Ecology. Cambridge: Cambridge University Press, 298 pp.CrossRefGoogle Scholar
Henry, E. C. (1988). Regulation of reproduction in brown algae by light and temperature. Bot. Mar. 31:353–7.CrossRefGoogle Scholar
Henry, E. C., and Cole, K. M. (1982). Ultrastructure of swarmers in the Laminariales. I. Zoospores. J. Phycol. 18:550–69.CrossRefGoogle Scholar
Hepburn, C. D., and Hurd, C. L. (2005). Conditional mutualism between the giant kelp Macrocystis pyrifera and colonial epifauna. Mar. Ecol. Prog. Ser. 302:37–48.CrossRefGoogle Scholar
Hepburn, C. D., Hurd, C. L., and Frew, R. D. (2006). Colony structure and seasonal differences in light and nitrogen modify the impact of sessile epifauna on the giant kelp Macrocystis pyrifera (L.) C Agardh. Hydrobiologia 560:373–84.CrossRefGoogle Scholar
Hepburn, C. D., Holborrow, J. D., Wing, S. R., Frew, R. D., and Hurd, C. L. (2007). Exposure to waves enhances the growth rate and nitrogen status of the giant kelp Macrocystis pyrifera. Mar. Ecol. Prog. Ser. 339:99–108.CrossRefGoogle Scholar
Hepburn, C. D., Pritchard, D. W., Cornwall, C. E., et al. (2011). Diversity of carbon use strategies in a kelp forest community: implications for a high CO2 ocean. Glob. Change Biol. 17:2488–97.CrossRefGoogle Scholar
Hepburn, C. D., Frew, R. D., and Hurd, C. L. (2012). Uptake and transport of nitrogen derived from sessile epifauna in the giant kelp Macrocystis pyrifera. Aquat. Biol. 14:121–8.CrossRefGoogle Scholar
Herbert, R. A. (1999). Nitrogen cycling in coastal marine ecosystems. FEMS Microbiol. Rev. 23:563–90.CrossRefGoogle ScholarPubMed
Herbert, S. K. (1990). Photoinhibition resistance in the red alga Porphyra perforata. The role of photoinhibition repair. Plant Physiol. 92:514–19.CrossRefGoogle ScholarPubMed
Herbert, S. K., and Waaland, J. R. (1988). Photoinhibition of photosynthesis in a sun and a shade species of the red algal genusPorphyra. Mar. Biol. 97:1–7.CrossRefGoogle Scholar
Hernández, I., Christmas, M., Yelloly, J. M., and Whitton, B. A. (1997). Factors affecting surface alkaline phosphatase activity in the brown alga Fucus spiralis at a North Sea intertidal site (Tyne Sands, Scotland). J. Phycol. 33:569–75.CrossRefGoogle Scholar
Hernández, I., Niell, F. X., and Whitton, B. A. (2002). Phosphatase activity of benthic marine algae. An overview. J. Appl. Phycol. 14:475–87.CrossRefGoogle Scholar
Heyward, A. J., and Negri, A. P. (1999). Natural inducers for coral larval metamorphosis. Coral Reefs 18:273–9.CrossRefGoogle Scholar
Higgins, H. W., and Mackey, D. J. (1987). Role of Ecklonia radiata (C. Ag.) J. Agardh in determining trace metal availability in coastal waters. I. Total trace metals. Aust. J. Mar. Freshw. Res. 38:307–15.CrossRefGoogle Scholar
Hill, N. A., Pepper, A. R., Poutinen, M. L., et al. (2010). Quantifying wave exposure in shallow temperate reef systems: applicability of fetch models for predicting algal biodiversity. Mar. Ecol. Prog. Ser. 417:83–95.CrossRefGoogle Scholar
Hillebrand, H., Gruner, D. S., Borer, E. T., et al. (2007). Consumer versus resource control of producer diversity depends on ecosystem type and producer community structure. P. Natl. Acad. Sci. USA 104:10,904–9.CrossRefGoogle ScholarPubMed
Hillis, L. (1997). Coral reefs from a calcareous green alga perspective, and a first carbonate budget. Proc. 8th Int. Coral Reef Symp. 1:761–6.Google Scholar
Hillis-Colinveaux, L. (1985). Halimeda and other deep forereef algae at Enewetak Atoll. In Harmelin, V. M. and Salvat, B. (eds), Proceedings of the 5th International Coral Reef Congress, Tahiti (vol. 5, pp. 9–14). Moorea, French Polynesia: Antenne Museum-EPHE.Google Scholar
Hillman, W. S. (1976). Biological rhythms and physiological timing. Annu. Rev. Plant Physiol. 27:159–79.CrossRefGoogle Scholar
Hincha, D. K. (2002). Cryoprotectin: a plant lipid-transfer protein homolog that stabilizes membranes during freezing. Phil. Trans. R. Soc. London B 357:909–15.CrossRefGoogle Scholar
Hinds, P.A., and Ballentine, D. L. (1987). Effects of the Caribbean threespot damselfish, Stegastes planifrons (Cuvier), on algal lawn composition. Aquat. Bot. 27:299–308.CrossRefGoogle Scholar
Hinojosa, I. A., Pizarro, M., Ramos, M., and Thiel, M. (2010). Spatial and temporal distribution of floating kelp in the channels and fjords of southern Chile. Estuar. Coast. Shelf S. 87:367–77.CrossRefGoogle Scholar
Hiscock, K. (1983). Water movement. In Earll, R. and Erwin, D. G. (eds), Sublittoral Ecology (pp. 58–96). Oxford: Clarendon Press.Google Scholar
Ho, Y. B. (1990). Metals in Ulva lactuca in Hong Kong intertidal waters. Bull. Mar. Sci. 47:79–85.Google Scholar
Hoarau, G., Coyer, J. A., and Olsen, J. L. (2009). Paternal leakage of mitochondrial DNA in a Fucus (Phaeophyceae) hybrid zone. J. Phycol. 45:621–4.CrossRefGoogle Scholar
Hoegh-Guldberg, O., Mumby, P. J., Hooten, A. J., et al. (2007). Coral reefs under rapid climate change and ocean acidification. Science 318:1737–42.CrossRefGoogle ScholarPubMed
Hoffmann, A. J., and Camus, P. (1989). Sinking rates and viability of spores from benthic algae in central Chile. J. Exp. Mar. Biol. Ecol. 126:281–91.CrossRefGoogle Scholar
Hofmann, L. C., Yildiz, G., Hanelt, D., and Bischof, K. (2012a). Physiological responses of the calcifying rhodophyte, Corallina officinalis (L.), to future CO2 levels. Mar. Biol. 159:783–92.CrossRefGoogle Scholar
Hofmann, L. C., Straub, S., and Bischof, K. (2012b). Competitive interactions between calcifying and noncalcifying temperate marine macroalgae under elevated CO2 levels: a mesocosm study. Mar. Ecol. Prog. Ser. 464:89–105.CrossRefGoogle Scholar
Holbrook, N. M., Denny, M. W., and Koehl, M. A. R. (1991). Intertidal “trees”: consequences of aggregation on the mechanical and photosynthetic properties of sea-palms Postelsia palmaeformis Ruprecht. J. Exp. Mar. Biol. Ecol. 146:39–67.CrossRefGoogle Scholar
Holmes, M. A., Brown, M. T., Loutit, M. W., and Ryan, K. (1991). The involvement of epiphytic bacteria in zinc concentration by the red algaGracilaria sordida. Mar. Environ. Res. 31:56–67.Google Scholar
Hommersand, M. H., and Fredericq, S. (1990). Sexual reproduction and cystocarp development. In Cole, K. M. and Sheath, R. G. (eds) Biology of the Red Algae (pp. 305–45). Cambridge: Cambridge University Press.Google Scholar
Hong, Y., Sohn, C. H., Polne-Fuller, M., and Gibor, A. (1995). Differential display of tissue-specific messenger RNAs in Porphyra perforata (Rhodophyta) thallus. J. Phycol. 31:640–3.CrossRefGoogle Scholar
Honkanen, T., and Jormalainen, V. (2005). Genotypic variation in tolerance and resistance to fouling in the brown alga Fucus vesiculosus. Oecologia 144:196–205.CrossRefGoogle ScholarPubMed
Hood, D. W., Schoener, A., Park, P. K., and Duedall, I. W. (1989). Evolution of at-sea scientific monitoring strategies. In Hood, D. W., Schoener, A., and Park, P. K. (eds), Oceanic Processes in Marine Pollution. Vol. 4: Scientific Monitoring Strategies for Ocean Waste Disposal (pp. 4–28). Malabau, FL: E. W. Krieger Publishing.Google Scholar
Hooper, R. G., and South, G. R. (1977). Distribution and ecology of Papenfussiella callitricha (Rosenv.) Kylin (Phaeophyceae, Chordariaceae). Phycologia 16:153–7.CrossRefGoogle Scholar
Hooper, R. G., Henry, E. C., and Kuhlenkamp, R. (1988). Phaeosiphoniella cryophila gen. et sp. nov., a third member of the Tilopteridales (Phaeophyceae). Phycologia 27:395–404CrossRefGoogle Scholar
Hop, H., Wiencke, C., Vögele, B., and Kovaltchouk, N. A. (2012). Species composition, zonation, and biomass of marine benthic macroalgae in Kongsfjorden, Svalbard. Bot. Mar. 55:399–414.CrossRefGoogle Scholar
Hopkin, R., and Kain, J. M. (1978). The effects of some pollutants on the survival, growth and respiration of Laminaria hyperborea. Estu. Cstl. Mar. Sci. 7:531–53. [, , ]CrossRefGoogle Scholar
Hou, X., and Yan, X. (1998). Study on concentration and seasonal variation of inorganic elements in 35 species of marine algae. Sci. Total Environ. 222:141–56.CrossRefGoogle Scholar
Hou, X. L., Dahlgaard, H., and Nielsen, S. P. (2000). Iodine-129 time series in Danish, Norwegian and northwest Greenland coast and the Baltic Sea by seaweed. Est. Coast. Shelf Sci. 52:571–84.CrossRefGoogle Scholar
Howard, B. M., and Fenical, W. (1981). The scope and diversity of terpenoid biosynthesis by the marine algaLaurencia. Prog. Phytochem. 7:263–300.Google Scholar
Howard, R. J., Gayler, K. R., and Grant, B. R. (1975). Products of photosynthesis in Caulerpa simpliciuscula. J. Phycol. 11: 463–71.Google Scholar
Howe, C. J., Barbrook, A. C., Nisbet, R. E. R., Lockhart, P. J., and Larkum, A. W. D. (2008). The origin of plastids. Philos. T. Roy. Soc. B. 363:2675–85.CrossRefGoogle ScholarPubMed
Hoyer, K., Karsten, U., Sawall, T., and Wiencke, C. (2001). Photoprotective substances in Antarctic macroalgae and their variation with respect to depth distribution, different tissues and developmental stages. Mar. Ecol. Prog. Ser. 211:117–29.CrossRefGoogle Scholar
Hsiao, S. I.-C. (1969). Life history and iodine nutrition of the marine brown alga Petalonia fascia (O. F. Mull.) Kuntze. Can. J. Bot. 47:1611–16.CrossRefGoogle Scholar
Huang, I., Rominger, J., and Nepf, H. (2011). The motion of kelp blades and the surface renewal model. Limnol. Oceanogr. 56:1453–62.CrossRefGoogle Scholar
Hubbard, C. B., Garbary, D. J., Kim, K. Y., and Chiasson, D. M. (2004). Host specificity and growth of kelp gametophytes symbiotic with filamentous red algae (Ceramiales, Rhodophyta). Helgoland Mar. Res. 58:18–25.CrossRefGoogle Scholar
Hughes, J. A., and Otto, S. P. (1999). Ecology and evolution of biphasic life cycles. Am. Nat. 154:306–20.CrossRefGoogle Scholar
Hughes, T. P. (1994). Catastrophes, phase shifts, and large-scale degradation of a Caribbean coral reef. Science 265:1547–51.CrossRefGoogle ScholarPubMed
Hughes, T. P, Reed, D. C., and Boyle, M.-J. (1987). Herbivory on coral reefs: community structure following mass mortality of sea urchins. J. Exp. Mar. Biol. Ecol. 113:39–60.CrossRefGoogle Scholar
Hughes, T., Szmant, A. M., Steneck, R., Carpenter, R., and Miller, S. (1999). Algal blooms on coral reefs: what are the causes? Limnol. Oceanogr. 44:1583–6.CrossRefGoogle Scholar
Huisman, J.M. (2000). Marine Plants of Australia. Nedlands, Australia: University of Western Australia Press.Google Scholar
Hulatt, C. J., Thomas, D. N., Bowers, D. G., Norman, L., and Zhang, C. (2009). Exudation and decomposition of chromophoric dissolved organic matter (CDOM) from some temperate macroalgae. Estuar. Coast Shelf Sci. 84: 147–53.CrossRefGoogle Scholar
Hunter, K. A., and Strzepek, R. (2008). Iron cycle. In Jorgensen, S.E. and Fath, B.D. (eds), Global Ecology, Vol. 3 of Encyclopedia of Ecology (pp. 2028–33).Oxford: Elsevier.CrossRefGoogle Scholar
Huovinen, P., and Gómez, I. (2011). Spectral attenuation of solar radiation in Patagonian fjords and coastal waters and implications for algal photobiology. Cont. Shelf Res. 31:254–9. [, , ]CrossRefGoogle Scholar
Huovinen, P., and Gómez, I. (2012). Cold temperate seaweed communities of the southern hemisphere. In Wiencke, C., and Bischof, K. (eds), Seaweed Biology: Novel Insights Into Ecophysiology, Ecology and Utilization (Series: Ecological Studies, Volume 219, pp. 293–313). Berlin and Heidelberg: Springer. [, , , , ]CrossRefGoogle Scholar
Huovinen, P., Gómez, I., Figueroa, F. L., et al. (2004). UV absorbing mycosporine-like amino acids in red macroalgae from Chile. Bot. Mar. 47:21–9.CrossRefGoogle Scholar
Huovinen, P., Gómez, I., and Orostegui, M. (2007). Patterns and UV sensitivity of carbonic anhydrase and nitrate reductase activities in south Pacific macroalgae. Mar. Biol. 151:1813–21.CrossRefGoogle Scholar
Huovinen, P., Leal, P., and Gómez, I. (2010). Impact of interaction of copper, nitrogen and UV radiation on the physiology of three south Pacific kelps. Mar. Freshw. Res. 61:330–41.CrossRefGoogle Scholar
Huppertz, K., Hanelt, D., and Nultsch, W. (1990). Photoinhibition of photosynthesis in the marine brown alga Fucus serratus as studied in field experiments. Mar. Ecol. Prog. Ser. 66:175–82.CrossRefGoogle Scholar
Hurd, C. L. (2000). Water motion, marine macroalgal physiology, and production. J. Phycol. 36:453–72. [, , ]CrossRefGoogle Scholar
Hurd, C. L., and Dring, M. J. (1990). Phosphate uptake by intertidal fucoid algae in relation to zonation and season. Mar. Biol. 107:281–9.CrossRefGoogle Scholar
Hurd, C. L., and Dring, M. J. (1991). Desiccation and phosphate uptake by intertidal fucoid algae in relation to zonation. Br. Phycol. J. 26:327–33.CrossRefGoogle Scholar
Hurd, C. L., and Pilditch, C. A. (2011). Flow-induced morphological variations affect diffusion boundary-layer thickness of Macrocystis pyrifera (Heterokontophyta, Laminariales). J. Phycol. 47:341–51. [, , ]CrossRefGoogle Scholar
Hurd, C. L., and Stevens, C. L. (1997). Flow visualization around single- and multiple-bladed seaweeds with various morphologies. J. Phycol. 33:360–7.CrossRefGoogle Scholar
Hurd, C.L., Galvin, R. S., Norton, T. A., and Dring, M. J. (1993). Production of hyline hairs by intertidal species of Fucus (Fucales) and their role in phosphate uptake. J. Phycol. 29:160–5. [, , ]CrossRefGoogle Scholar
Hurd, C. L., Durante, K. M., Chia, F. S., and Harrison, P. J. (1994a). Effect of bryozoan colonization on inorganic nitrogen acquisition by the kelps Agarum fimbriatum and Macrocystis integrifolia. Mar. Biol. 121:167–73.CrossRefGoogle Scholar
Hurd, C. L., Quick, M., Stevens, C. L., et al. (1994b). A low-volume flow tank for measuring nutrient uptake by large macrophytes. J. Phycol. 30:892–6.CrossRefGoogle Scholar
Hurd, C.L., Berges, J. A., Osbourne, J., and Harrison, P. J. (1995). An in vitro nitrate reductase assay for marine macroalgae: optimization and characterization of the enzyme for Fucus gardneri (Phaeophyta). J. Phycol. 31:835–43.CrossRefGoogle Scholar
Hurd, C. L., Harrison, P. J., and Druehl, L. D. (1996). Effect of seawater velocity on inorganic nitrogen uptake by morphologically distinct forms of Macrocystis integrifolia from wave-sheltered and exposed sites. Mar. Biol. 126:205–14.CrossRefGoogle Scholar
Hurd, C. L., Stevens, C. L., Laval, B. E., et al. (1997). Visualization of seawater flow around morphologically distinct forms of the giant kelp Macrocystis integrifolia from wave-sheltered and exposed sites. Limnol. Oceanogr. 42:156–63.CrossRefGoogle Scholar
Hurd, C. L., Hepburn, C. D., Currie, K. I., Raven, J. A., and Hunter, K. A. (2009). Testing the effects of ocean acidification on algal metabolism: considerations for experimental designs. J. Phycol. 45:1236–51.CrossRefGoogle ScholarPubMed
Hurd, C. L., Cornwall, C. E., Currie, K., et al. (2011). Metabolically-induced pH fluctuations by some coastal calcifiers exceed projected 22nd century ocean acidification: a mechanism for differential susceptibility? Glob. Change Biol. 17:3254–62.CrossRefGoogle Scholar
Hurlbert, S. H. (1984). Pseudoreplication and the design of ecological field experiments. Ecol. Monogr. 54:187–211.CrossRefGoogle Scholar
Hutchins, L. W. (1947). The basis for temperature zonation in geographical distribution. Ecol. Monogr. 17:325–35.CrossRefGoogle Scholar
Hwang, E. K., and Dring, M. J. (2002). Quantitative photoperiodic control of erect thallus production in Sargassum muticum. Bot. Mar. 45:471–5.CrossRefGoogle Scholar
Hyndes, G. A., and Lavery, P. S. (2005). Does transported seagrass provide an important trophic link in unvegetated, nearshore areas? Est. Coast. Shelf Sci. 63(4):633–43.CrossRefGoogle Scholar
Hyndes, G. A., Lavery, P. S., and Doropoulos, C. (2012). Dual processes for cross-boundary subsidies: incorporation of nutrients from reef-derived kelp into a seagrass ecosystem. Mar. Ecol. Prog. Ser. 445:97–107.CrossRefGoogle Scholar
Ianora, A., Boersma, M., Casotti, R., et al. (2006). The H. T. Odum synthesis essay: new trends in marine chemical ecology. Estuar. Coast. 29:531–51.CrossRefGoogle Scholar
Ikawa, M., Watanabe, T, and Nisizawa, K. (1972). Enzymes involved in the last steps of the biosynthesis of mannitol in brown algae. Plant Cell Physiol. 3:1017–29.Google Scholar
Iken, K. (2012). Grazers on benthic seaweeds. In Wiencke, C. and Bischof, K. (eds), Seaweed Biology: Novel Insights Into Ecophysiology, Ecology and Utilization (Series: Ecological Studies, Volume 219, pp. 157–75). Berlin and Heidelberg: Springer.CrossRefGoogle Scholar
Ilvessalo, H., and Tuomi, J. (1989). Nutrient availability and accumulation of phenolic compounds in the brown alga Fucus vesiculosus. Mar. Biol. 101:115–19.CrossRefGoogle Scholar
Inderjit, , Chapman, D., Ranelletti, M., and Kaushik, S. (2006). Invasive marine algae: an ecological perspective. Bot. Rev. 72:153–78.Google Scholar
Innes, D. J. (1988). Genetic differentiation in the intertidal zone in populations of the alga Enteromorpha linza (Ulvales: Chlorophyta). Mar. Biol. 97:9–16.CrossRefGoogle Scholar
Inouye, R. S., and Schaffer, W. M. (1981). On the ecological meaning of ratio (de Wit) diagrams in plant ecology. Ecology 62: 1679–81.CrossRefGoogle Scholar
IPCC (2007). Climate Change 2007: The physical science basis. Summary for policymakers. Contribution of working group I to the fourth assessment report. The Intergovernmental Panel on Climate Change, .
Ireland, H. E., Harding, S. J., Bonwick, G. A., et al. (2004). Evaluation of heat shock protein 70 as a biomarker of environmental stress in Fucus serratus and Lemna minor. Biomarkers 9:139–55.CrossRefGoogle Scholar
Irving, A. D., and Connell, S. D. (2006a). Predicting understorey structure from the presence and composition of canopies: an assembly rule for marine algae. Oecologia 148:491–502.CrossRefGoogle ScholarPubMed
Irving, A. D., and Connell, S. D. (2006b). Physical disturbance by kelp abrades erect algae from the understorey. Mar. Ecol. Prog. Ser. 324:127–37.CrossRefGoogle Scholar
Irving, A. D., Balata, D., Colosio, F., Ferrando, G. A., and Airoldi, L. (2009). Light, sediment, temperature, and the early life-history of the habitat-forming alga Cystoseira barbata. Mar. Biol. 156:1223–31.CrossRefGoogle Scholar
Irwin, S., and Davenport, J. (2002). Hyperoxic boundary layers inhabited by the epiphytic meiofauna of Fucus serratus. Mar. Ecol. Prog. Ser. 244:73–9.CrossRefGoogle Scholar
Ishikawa, M., Takahashi, F., and Nozaki, H. (2009). Distribution and phylogeny of the blue light receptors aureochromes in eukaryotes. Planta 230:543–52.CrossRefGoogle ScholarPubMed
ISO (2006). Water quality – Marine algal growth inhibition test with Skeletonema costatum and Phaeodactylum tricornutum. ISO 10253:2006.
ISO (2010). Water quality — Growth inhibition test with the marine and brackish water macroalga Ceramium tenuicorne. ISO 10710:2010.
Israel, A., Levy, I., and Friedlander, M. (2006). Experimental tank cultivation of Porphyra in Israel. J. Appl. Phycol. 18:235–40.CrossRefGoogle Scholar
Iwamoto, K., and Ikawa, T. (1997). Glycolate metabolism and subcellular distribution of glycolate oxidase in Patoglossum pacificum (Phaeophyceae, Chromophyta). Phycol. Res. 45:77–83.CrossRefGoogle Scholar
Jackson, G. A. (1984). The physical and chemical environment of a kelp community. In Bascom, W. (ed.), The Effects of Waste Disposal on Kelp Communities (pp. 11–37). Long Beach, CA: So. Calif. Coastal Water Res Proj.Google Scholar
Jackson, G. A. (1987). Modelling the growth and harvest yield of the giant kelp Macrocystis pyrifera. Mar. Biol. 95:611–24.CrossRefGoogle Scholar
Jackson, G. A., James, D. E., and North, W. J. (1985). Morphological relationships among fronds of giant kelp Macrocystis pyrifera off La Jolla, California. Mar. Ecol. Prog. Ser. 26:261–70.CrossRefGoogle Scholar
Jacobs, J. P. (1993). A search for some angiosperm hormones and their metabolites in Caulerpa paspaloides (Chlorophyta). J. Phycol. 29:595–600.CrossRefGoogle Scholar
Jacobs, W. P. (1970). Develoment and regeneration of the algal giant coenocyteCaulerpa. Ann. N. Y. Acad. Sci. 175:732–48.CrossRefGoogle Scholar
Jacobs, W. P. (1994). Caulerpa. Scientific American, December 1994.CrossRef
Jacobs, W. P., and Olson, J. (1980). Developmental changes in the algal coenocyte Caulerpa prolifera (Siphonales) after inversion with respect to gravity. Am. J. Bot. 67: 141–6.CrossRefGoogle Scholar
Janouškovec, J., Liu, S.-L., Martone, P. T., et al. (2013). Evolution of red algal plastid genomes: ancient architectures, introns, horizontal gene transfer, and taxonomic utility of plastid markers. PLoS ONE 8(3), e59001.CrossRefGoogle ScholarPubMed
Jaschinski, S., Brepohl, D. C., and Sommer, U. (2008). Carbon sources and trophic structure in an eelgrass Zostera marina bed, based on stable isotope and fatty acid analyses. Mar. Ecol. Prog. Ser. 358:103–14.CrossRefGoogle Scholar
Jassby, A. D., and Platt, T. (1976). Mathematical formulation of the relationship between photosynthesis and light for phytoplankton. Limnol. Oceanogr. 21:540–7.CrossRefGoogle Scholar
Jékely, G. (2009). Evolution of phototaxis. Phil. Trans. R. Soc. B. 364:2795–808.CrossRefGoogle ScholarPubMed
Jelinek, R., and Kolusheva, S. (2004). Carbohydrate biosensors. Chem. Rev. 104:5987–6015.CrossRefGoogle ScholarPubMed
Jenkins, S. R., Moore, P., Burrows, M. T., et al. (2008). Comparative ecology of North Atlantic shores: do differences in players matter for process? Ecology 89:S3-S23.CrossRefGoogle ScholarPubMed
Jennings, J. G., and Steinberg, P. D. (1997). Phlorotannins versus other factors affecting epiphyte abundance on the kelp Ecklonia radiata. Oecologia 109:461–73.CrossRefGoogle ScholarPubMed
Jensen, A. (1995). Production of alginate. In Wiessner, W., Schnepf, E., and Star, R. C. (eds) Algae, Environment and Human Affairs (pp. 79–92). Bristol, UK: Biopress Ltd.Google Scholar
Jensen, A., and Haug, A. (1956). Geographical and seasonal variation in the chemical compositon of Laminaria hyperborea and Laminaria digitata from the Norwegian Coast. Norwegian Institute of Seaweed Research, Report 14.
Jensen, R. G., and Bahr, J. T. (1977). Ribulose 1,5-bisphosphate carboxylase-oxygenase. Annu. Rev. Plant Physiol. 28:379–400.CrossRefGoogle Scholar
Jerlov, N. G. (1970). Light: general introduction. In Kinne, O. (ed), Marine Ecology (vol. I, pt. I, pp. 95–102). New York: Wiley.Google Scholar
Jerlov, N. G. (1976). Marine Optics. Amsterdam: Elsevier, 231 pp.Google Scholar
Jiménez, C., Berl, T., Rivard, C. J., Edelstein, C. L., and Capasso, J. M. (2004). Phosphorylation of MAP kinase-like proteins mediate the response of the halotolerant alga Dunaliella viridis to hypertonic shock. Biochim. Biophys. Acta: Mol. Cell Res. 1644: 61–9.CrossRefGoogle ScholarPubMed
Jin, Q., and Dong, S. (2003). Comparative studies on the allelopathic effects of two different strains of Ulva pertusa on Heterosigma akashiwo and Alexandrium tamarense. J. Exp. Mar. Biol Ecol. 293:41–55.CrossRefGoogle Scholar
Jin, Q., Dong, S., and Wang, C. (2005). Allelopathic growth inhibition of Prorocentrum micans (Dinophyta) by Ulva pertusa and Ulva linza (Chlorophyta) in laboratory cultures. Eur. J. Phycol. 40:31–7.CrossRefGoogle Scholar
Johannesson, K. (1989). The bare zone of Swedish rocky shores: why is it there? Oikos 54:77–86.CrossRefGoogle Scholar
Johansson, G., and Snoeijs, P. (2002). Macroalgal photosynthetic responses to light in relation to thallus morphology and depth zonation. Mar. Ecol. Prog. Ser. 244:63–72.CrossRefGoogle Scholar
Johansson, G., Sosa, P. A., and Snoeijs, P. (2003). Genetic variability and level of differentiation in North Sea and Baltic Sea populations of the green alga Cladophora rupestris. Mar. Biol. 142:1019–27.CrossRefGoogle Scholar
Johnson, A. S. (2001). Drag, drafting, and mechanical interactions in canopies of the red alga Chondrus crispus. Biol. Bull. 201:126–35.CrossRefGoogle ScholarPubMed
Johnson, A., and Koehl, M. (1994). Maintenance of dynamic strain similarity and environmental stress factor in different flow habitats: thallus allometry and material properties of a giant kelp. J. Exp. Biol. 195:381–410.Google ScholarPubMed
Johnson, C. H. (2010). Circadian clocks and cell division. Cell Cycle 9:3864–73.CrossRefGoogle ScholarPubMed
Johnson, C. R., and Chapman, A.R.O. (2007). Seaweed invasions: introduction and scope. Bot. Mar. 50:321–5.CrossRefGoogle Scholar
Johnson, C. R., Banks, S. C., Barrett, N. S., et al. (2011). Climate change cascades: shifts in oceanography, species’ ranges and subtidal marine community dynamics in eastern Tasmania. J. Exp. Mar. Biol. Ecol. 400:17–32.CrossRefGoogle Scholar
Johnson, L. E., and Brawley, S. H. (1998). Dispersal and recruitment of a canopy-forming intertidal alga: the relative roles of propagule availability and post-settlement processes. Oecologia 117:517–26.CrossRefGoogle ScholarPubMed
Johnson, M. P., Hawkins, S. J., Hartnoll, R. G., and Norton, T. A. (1998). The establishment of fucoid zonation on algal-dominated rocky shores: hypotheses derived from a simulation model. Funct. Ecol. 12:259–69.CrossRefGoogle Scholar
Johnson, W. S., Gigon, A., Gulmon, S. L., and Mooney, H. A. (1974). Comparative photosynthetic capacities of intertidal algae under exposed and submerged conditions. Ecology 55:450–3.CrossRefGoogle Scholar
Johnston, A. M., and Raven, J. A. (1986). Dark carbon fixation studies on the intertidal macroalga Ascophyllum nodosum (Phaeophyta). J. Phycol. 22:78–83.CrossRefGoogle Scholar
Johnston, A. M., and Raven, J. A. (1990). Effects of culture in high CO2 on the photosynthetic physiology ofFucus serratus. Br. Phycol. J. 25:75–82.CrossRefGoogle Scholar
Johnston, A. M., Maberly, S. C., and Raven, J. A. (1992). The acquisition of inorganic carbon by four red macroalgae from different habitats. Oecologia 92:317–26.CrossRefGoogle Scholar
Joint, I., Callow, M. E., Callow, J. A., and Clarke, K. R. (2000). The attachment of Enteromorpha zoospores to a bacterial biofilm assemblage. Biofouling 16:151–8.CrossRefGoogle Scholar
Joint, I., Tait, K., and Callow, M. E. (2002). Cell-to-cell communication across the prokaryote–eukaryote boundary. Science 298:1207.CrossRefGoogle ScholarPubMed
Joly, A. B., and de Oliveira Filho, E. C. (1967). Two BrazilianLaminaria. Publ. Inst. Pesq. Mar. 4:1–13.Google Scholar
Jones, A. B., Dennison, W. C., and Stewart, G. R. (1996). Macroalgal responses to nitrogen source and availability: amino acid metabolic profiling as a bioindicator using Gracilaria edulis (Rhodophyta). J. Phycol. 32:757–66.CrossRefGoogle Scholar
Jones, G. P., Santana, L., McCook, L. J., and McCormick, M. I. (2006). Resource use and impact of three herbivorous damselfishes on coral reef communities. Mar. Ecol. Prog. Ser. 328:215–24.CrossRefGoogle Scholar
Jones, J. L., Callow, J. A., and Green, J. R. (1988). Monoclonal antibodies to sperm surface antigens of the brown alga Fucus serratus exhibit region-, gamete-, species- and genus preferential binding. Planta 176:298–306.CrossRefGoogle ScholarPubMed
Jónsson, S., Laur, M.-H., and Pham-Quang, L. (1985). Mise en evidence de differents types de glycoproteines dans un extrait inhibiteur de la gametogenese chez Enteromorpha prolifera, Chlorophyceae marine. Crypt. Algol. 6:253–64.Google Scholar
Jormalainen, V., and Honkanen, T. (2008). Macroalgal chemical defenses and their roles in structuring temperate marine communities. In Amsler, C. D. (ed.), Algal Chemical Ecology (pp. 57–90). London: Springer (Limited).CrossRefGoogle Scholar
Jormalainen, V., Honkanen, T., and Heikkila, N. (2001). Feeding preferences and performance of a marine isopod on seaweed hosts: cost of habitat specialization. Mar. Ecol. Prog. Ser. 220:219–30.CrossRefGoogle Scholar
Jormalainen, V., Honkanen, T., Koivikko, R., et al. (2003). Induction of phlorotannin production in a brown alga: defense or resource dynamics? Oikos 103:640–50.CrossRefGoogle Scholar
Joska, M. A. P., and Bolton, J. J. (1987). In situ measurement of zoospore release and seasonality of reproduction in Ecklonia maxima (Alariaceae, Laminariales). Br. Phycol. J. 22:209–14.CrossRefGoogle Scholar
Josselyn, M. N., and Mathieson, A. C. (1978). Contribution of receptacles from the fucoid Ascophyllum nodosum to the detrital pool of a north temperate estuary. Estuaries 1:258–61.CrossRefGoogle Scholar
Juanes, J. A., Guinda, X., Puente, A., and Revilla, J. A. (2008). Macroalgae, a suitable indicator of the ecological status of coastal rocky communities in the NE Atlantic. Ecol. Indicators 8:351–9.CrossRefGoogle Scholar
Jung, V., and Pohnert, G. (2001). Rapid wound-activated transformation of the green algal defensive metabolite caulerpenyne. Tetrahedron 57:7169–72.CrossRefGoogle Scholar
Kaczmarska, I., and Dowe, L. L. (1997). Reproductive biology of the red alga Polysiphonia lanosa (Ceramiales) in the Bay of Fundy, Canada. Mar. Biol. 128:695–703.CrossRefGoogle Scholar
Kagami, Y., Mogi, Y., Arai, T., et al. (2008). Sexuality and uniparental inheritance of chloroplast DNA in the isogamous green alga Ulva compressa (Ulvophyceae). J. Phycol. 44:691–702.CrossRefGoogle Scholar
Kai, T., Nimura, K., Yasui, H., and Mizuta, H. (2006). Regulation of sorus formation by auxin in Laminariales sporophyte. J. Appl. Phycol. 18:95–101.CrossRefGoogle Scholar
Kain, J. M. (1969). The biology of Laminaria hyperborea. V. Comparison with early stages of competitors. J. Mar. Biol. Assoc. UK 49:455–73.CrossRefGoogle Scholar
Kain, J. M. (1979). A view of the genusLaminaria. Oceanogr. Mar. Biol. Annu. Rev. 17:101–61.Google Scholar
Kain, J. M. (1989). The seasons in the subtidal. Brit. Phycol. J. 24:203–15.CrossRefGoogle Scholar
Kain (Jones), J. M. (2006). Photoperiodism in Delesseria sanguinea (Ceramiales, Rhodophyta) 2. Daylengths are shorter underwater. Phycologia 45:624–31.CrossRefGoogle Scholar
Kain (Jones), J. M. (2008). Winter favours growth and survival of Ralfsia verrucosa (Phaeophyta) in high intertidal rockpools in southeast Australia. Phycologia 47:498–509.CrossRefGoogle Scholar
Kain (Jones), J. M., and Destombe, C. (1995). A review of the life history, reproduction and phenology of Gracilaria. J. Appl. Phycol. 7:269–81.CrossRefGoogle Scholar
Kaiser, M. J., Attrill, M. J., Jennings, S., et al. (2011). Marine Ecology: Processes, Systems and Impacts. Oxford: Oxford University Press, 528 pp.Google Scholar
Kajiwara, T., Hatanaka, A., Tanaka, Y., et al. (1989). Volatile constitutents from marine brown algae of JapaneseDictyopteris. Phytochemistry 28:636–8.CrossRefGoogle Scholar
Kalle, K. (1945). Der Stoffhaushalt der Meere. Leipzig, Germany: Becker & Erler, 263 pp.Google Scholar
Kalle, K. (1971). Salinity: general introduction. In Kinne, O. (ed.), Marine Ecology (vol. I, pt. 2, pp. 683–8). New York: Wiley.Google Scholar
Kalle, K. (1972). Dissolved gases: general introduction. In Kinne, O. (ed.), Marine Ecology (vol. I, pt. 3, pp. 1451–7). NewYork: Wiley.Google Scholar
Kamiya, M., and West, J. A. (2010). Investigations on reproductive affinities in red algae. In Seckbach, J. and Chapman, D. J. (eds), Red Algae in the Genomic Age (Cellular Origin, Life in Extreme Habitats and Astrobiology 13) (pp. 77–109). Dordrecht, The Netherlands: Springer. [, , ]Google Scholar
Kapraun, D. F. (2005). Nuclear DNA content estimates in multicellular green, red and brown algae: phylogenetic considerations. Ann. Bot.-London. 95:7–44.CrossRefGoogle Scholar
Kapraun, D. E., and Boone, P W. (1987). Karyological studies of three species of Scytosiphonaceae (Phaeophyta) from coastal North Carolina. J. Phycol. 23:318–22.CrossRefGoogle Scholar
Kapraun, D. E., and Martin, D. J. (1987). Karyological studies of three species of Codium (Codiales, Chlorophyta) from coastal North Carolina. Phycologia 26:228–34.CrossRefGoogle Scholar
Karez, R., Engelbert, S., and Sommer, U. (2000). “Co-consumption” and “protective coating”: two new proposed effects of epiphytes on their macroalgal hosts in mesograzer–epiphyte–host interactions. Mar. Ecol. Prog. Ser. 205:85–93.CrossRefGoogle Scholar
Karlsson, J., and Eklund, B. (2004). New biocide-free antifouling paints are toxic. Mar. Poll. Bull. 49:456–64.CrossRefGoogle Scholar
Karlsson, J., Breitholtz, M., and Eklund, B. (2006). A practical ranking system to compare toxicity of antifouling paints. Mar. Poll. Bull. 52:1661–7.CrossRefGoogle Scholar
Karlsson, J., Ytreberg, E., and Eklund, B. (2010). Toxicity of antifouling paints for use on pleasure boats and vessels to non-target organisms representing three trophic levels. Environ. Poll. 158:681–7.CrossRefGoogle Scholar
Karsten, U. (2007). Salinity tolerance of Arctic kelps from Spitsbergen. Phycol. Res. 55:257–62.CrossRefGoogle Scholar
Karsten, U., and Kirst, G. O. (1989). Incomplete turgor pressure regulation in the “terrestrial” red alga, Bostrychia scorpioides (Huds.). Mont. Plant Sci. 61:29–36.CrossRefGoogle Scholar
Karsten, U., King, R. J., and Kirst, G. O. (1990). The distribution of D-sorbitol and D-dulcitol in the red algal genera Bostrychia and Stictosiphonia (Rhodomelaceae, Rhodophyta) a re-evalution. Br. Phycol. J. 25:363–6.CrossRefGoogle Scholar
Karsten, U., Barrow, K. D., Nixdorf, O., West, J. A., and King, R. J. (1997). Characterization of mannitol metabolism in the mangrove red alga Caloglossa leprieurii (Montagne)J. Agardh. Planta 201:173–8.CrossRefGoogle Scholar
Karsten, U., Michalik, D., Michalik, M., and West, J. A. (2005). A new unusual low molecular weight carbohydrate in the red algal genus Hypoglossum (Delesseriaceae, Ceramiales) and its possible function as osmolyte. Planta 222:319–26.CrossRefGoogle ScholarPubMed
Karsten, U., Görs, S., Eggert, A., and West, J. A. (2007). Trehalose, digeneaside and floridoside in the Florideophyceae (Rhodophyta) – a re-evaluation of its chemotaxonomic value. Phycologia 46:143–50. [, , ]CrossRefGoogle Scholar
Karyophyllis, D., Katsaros, C., and Galatis, B. (2000). F-actin involvement in apical cell morphogenesis of Sphacelaria rigidula (Phaeophyceae): mutual alignment between cortical actin filaments and cellulose microfibrils. Eur. J. Phycol. 35:195–203.CrossRefGoogle Scholar
Katoh, T., and Ehara, T. (1990). Supramolecular assembly of fucoxanthin-chlorophyll-protein complexes isolated from a brown alga, Petalonia fascia. Electron microscopic studies. Plant Cell Physiol. 31:439–47.Google Scholar
Katsaros, C., and Galatis, B. (1988). Thallus development in Dictyopteris membranacea (Phaeophyta, Dictyotales). Br. Phycol. J. 23:71–88.CrossRefGoogle Scholar
Katsaros, C., Karyophyllis, D., and Galatis, B. (2006). Cytoskeleton and morphogenesis in brown algae. Ann. Bot.-London 97:679–93.CrossRefGoogle ScholarPubMed
Katsaros, C., Motomura, T., Nagasato, C., and Galantis, B. (2009). Diaphragm development in cytokinetic vegetative cells of brown algae. Bot. Mar. 52:150–61.CrossRefGoogle Scholar
Kavanaugh, M. T., Nielsen, K. J., Chan, F. T., et al. (2009). Experimental assessment of the effects of shade on an intertidal kelp: do phytoplankton blooms inhibit growth of open-coast macroalgae? Limnol. Oceanogr. 54:276–88.CrossRefGoogle Scholar
Kawai, H., Müller, D. G., Fölster, E., and Häder, D. P (1990). Phototactic responses in the gametes of the brown alga, Ectocarpus siliculosus. Planta 182:292–7.Google ScholarPubMed
Kawai, H., Nakamura, S., Mimuro, M., Furuya, M., and Watanabe, M. (1996). Microspectrofluorometry of the autofluorescent flagellum in phototactic brown algal zoids. Protoplasma 191:172–7.CrossRefGoogle Scholar
Kawamata, S. (1998). Effect of wave-induced oscillatory flow on grazing by a subtidal sea urchin Strongylocentrotus nudus (A. Agassiz). J. Exp. Mar. Biol. Ecol. 224:31–48.CrossRefGoogle Scholar
Kawamata, S. (2001). Adaptive mechanical tolerance and dislodgement velocity of the kelp Laminaria japonica in wave-induced water motion. Mar. Ecol. Prog. Ser. 211:89–104.CrossRefGoogle Scholar
Kawamata, S. (2010). Inhibitory effects of wave action on destructive grazing by sea urchins: a review. Bull. Fish. Res. Agen. 32:95–102.Google Scholar
Kawamata, S., Yoshimitsu, S., Tokunaga, S., Kubo, S., and Tanaka, T. (2012). Sediment tolerance of Sargassum algae inhabiting sediment-covered rocky reefs. Mar. Biol. 159:723–33.CrossRefGoogle Scholar
Kawamitsu, Y., and Boyer, J. S. (1999). Photosynthesis and carbon storage between tides in a brown alga, Fucus vesiculosus. Mar. Biol. 133:361–9.CrossRefGoogle Scholar
Kawashima, S. (1984). Kombu cultivation in Japan for human foodstuff. Jpn. J. Phycol. 32:379–94.Google Scholar
Keats, D. W., Knight, M. A., and Pueschel, C. M. (1997). Antifouling effects of epithallial shedding in three crustose coralline algae (Rhodophyta, Coralinales) on a coral reef. J. Exp. Mar. Biol. Ecol. 213:281–93.CrossRefGoogle Scholar
Keeling, P. J. (2009). Chromalveolates and the evolution of plastids by secondary endosymbiosis. J. Eukaryot. Microbiol. 56:1–8.CrossRefGoogle ScholarPubMed
Keeling, P. J. (2010). The endosymbiotic origin, diversification and fate of plastids. Phil. Trans. R. Soc. B. 365:729–48.CrossRefGoogle ScholarPubMed
Keiji, I., and Kanji, H. (1989). Seaweeds: chemical composition and potential food uses. Food Rev. Internat. 5:101–44. [, , ]Google Scholar
Keiter, S., Braunbeck, T., Heise, S., et al. (2009). A fuzzy logic-classification of sediments based on data from in vitro biotests. J. Soils Sed. 9:168–79.CrossRefGoogle Scholar
Kelly, G. J. (1989). A comparison of marine photosynthesis with terrestrial photosynthesis: a biochemical perspective. Oceanogr. Mar. Biol. Annu. Rev. 27:11–44.Google Scholar
Kendrick, G. A. (1991). Recruitment of coralline crusts and filamentous turf algae in the Galapagos Archipelago: effect of simulated scour, erosion and accretion. J. Exp. Mar. Biol. Ecol. 147:47–63.CrossRefGoogle Scholar
Kennelly, S. J. (1987a). Inhibition of kelp recruitment by turfing algae and consequences for an Australian kelp community. J. Exp. Mar. Biol. Ecol. 112:49–60.CrossRefGoogle Scholar
Kennelly, S. J. (l987b). Physical disturbances in an Australian kelp community. I. Temporal effects. Mar. Ecol. Prog. Ser. 40: 145–53.CrossRefGoogle Scholar
Kenyon, K. W., and Rice, D. W. (1959). Life history of the Hawaiian monk seal. Pac. Sci. 13:215–52.Google Scholar
Keogh, S. M., Aldahan, A., and Possnert, G. (2007). Trends in the spatial and temporal distribution of 129I and 99Tc in coastal waters surrounding Ireland using Fucus vesiculosus as a bio-indicator. J. Environ. Radioact. 95:23–38.CrossRefGoogle ScholarPubMed
Kerby, N. W., and Raven, J. A. (1985). Transport and fixation of inorganic carbon by marine algae. Adv. Bot. Res. 11:71–123.CrossRefGoogle Scholar
Kerby, N. W., and Evans, L. V. (1983). Phosphoenolpyruvate carboxykinase activity in Ascophyllum nodosum (Phaeophyceae). J Phycol 19:1–3.CrossRefGoogle Scholar
Kerr, R. A. (2011). First detection of ozone hole recovery claimed. Science 332:160.CrossRefGoogle ScholarPubMed
Kershaw, P. J., McCubbin, D., and Leonard, K. S. (1999). Continuing contamination of north Atlantic and Arctic waters by Sellafield radionuclides. Sci. Total Envir. 237/238:119–32.CrossRefGoogle ScholarPubMed
Kerswell, A. P. (2006). Global biodiversity patterns of benthic marine algae. Ecology 87:2479–88.CrossRefGoogle ScholarPubMed
Keser, M., Swenarton, J. T., and Foertch, J. F. (2005). Effects of thermal input and climate change on growth of Ascophyllum nodosum (Fucales, Phaeophyceae) in eastern Long Island Sound. J. Sea Res. 54:211–20.CrossRefGoogle Scholar
Kessler, W. S. (2006). The circulation of the eastern tropical Pacific: a review. Prog. Oceanogr. 69:181–217.CrossRefGoogle Scholar
Khailov, K. M., and Burlakova, Z. P. (1969). Release of dissolved organic matter by marine seaweeds and distribution of their total organic production to inshore communities. Limnol. Oceanogr. 14:521–7.CrossRefGoogle Scholar
Khailov, K. M., Kholodov, V. I., Firsov, Y. K., and Prazukin, A. V. (1978). Thalli of Fucus vesiculosus in ontogenesis: changes in morpho-physiological parameters. Bot. Mar. 21:289–311.CrossRefGoogle Scholar
Kharlamenko, V. I., Kiyashko, S. I., Imbs, A. B., and Vyshkvartzev, D. I. (2001). Identification of food sources of invertebrates from the seagrass Zostera marina community using carbon and sulfur stable isotope ratio and fatty acid analyses. Mar. Ecol. Prog. Ser. 220:103–17.CrossRefGoogle Scholar
Khotimchemko, S.V. (2003). The fatty acid composition of glycolipids of marine macrophytes. Russian J. Mar. Biol. 29(2):126–8.CrossRefGoogle Scholar
Kilar, J. A., Littler, M. M., and Littler, D. S. (1989). Functionalmorphological relationships in Sargassum polyceratium (Phaeophyta): phenotypic and ontogenetic variability in apparent photosynthesis and dark respiration. J. Phycol. 25:713–20.CrossRefGoogle Scholar
Kim, G. H., Lee, I. K., and Fritz, L. (1996). Cell–cell recognition during fertilization in the red alga, Aglaothamnion sparsum (Ceramiaceae, Rhodophyta). Plant Cell Physiol. 37:621–8.CrossRefGoogle Scholar
Kim, G. H., Klotchkova, T. A., Lee, B.-C., and Kim, S. H. (2001a). FITC-phalloidin staining of F-actin in Aglaothamnion oosumiense and Griffithsia japonica (Rhodophyta). Bot. Mar. 44:501–8.CrossRefGoogle Scholar
Kim, G. H., Klotchkova, T. A., and Kang, Y. M. (2001b). Life without a cell membrane: regeneration of protoplasts from disintegrated cells of the marine green alga Bryopsis plumosa. J. Cell Sci. 114:2009–14.Google ScholarPubMed
Kim, G. H., Klotchkova, T. A., and West, J. A. (2002). From protoplasm to swarmer: regeneration of protoplasts from disintegrated cells of the multicellular marine green alga Microdictyon umbilicatum (Chlorophyta). J. Phycol. 38:174–83.CrossRefGoogle Scholar
Kim, G. H., Klochkova, T. A., Yoon, K.-S., Song, Y.-S., and Lee, K. P. (2005). Purification and characterization of a lectin, bryohealin, involved in the protoplast formation of a marine green alga Bryopsis plumosa (Chlorophyta). J. Phycol. 42:86–95.CrossRefGoogle Scholar
Kim, J. K., Kraemer, G. P., and Yarish, C. (2013). Integrated multi-tropic aquaculture in the United States. In Chopin, T., Neori, A., Robinson, S., and Troell, M. (eds), Integrated Multi-Trophic Aquaculture (IMTA). New York: Springer Science.Google Scholar
Kim, S.-H., and Kim, G. H. (1999). Cell–cell recognition during fertilization in the red alga, Aglaothamnion oosumiense (Ceramiaceae, Rhodophyta). Hydrobiologia 398/399:81–9.CrossRefGoogle Scholar
Kimura, K., Nagasto, C., Kogame, K., and Motomura, T. (2010). Disappearance of male mitochondrial DNA after the four-cell stage in sporophytes of the isogamous brown alga Scytosiphon lomentaria (Scytosiphonaceae, Phaeophyceae). J. Phycol. 46:143–52.CrossRefGoogle Scholar
Kindig, A. C., and Littler, M. M. (1980). Growth and primary productivity of marine macrophytes exposed to domestic sewage effluents. Mar. Environ. Res. 3:81–100.CrossRefGoogle Scholar
King, R. J. (1984). Oil pollution and marine plant systems. In Cheng, M. H. and Field, C. D. (eds). Pollution and Plants (pp. 127–42). Melbourne: Insearch Ltd.Google Scholar
King, R. J. (1990). Macroalgae associated with the mangrove vegetation of Papua New Guinea. Bot. Mar. 33:55–62.CrossRefGoogle Scholar
Kingham, D. L., and Evans, L. V. (1986). The Pelvetia–Mycosphaerella interrelationship. In Moss, S. T. (ed.), The Biology of Marine Fungi (pp. 177–87). Cambridge: Cambridge University Press.Google Scholar
Kingsford, M., and Battershill, C. (eds) (1998). Studying Temperate Marine Environments: A Handbook for Ecologists. Christchurch, NZ: Canterbury University Press, 335 pp.Google Scholar
Kinne, O. (1970). Temperature: general introduction. In Kinne, O. (ed), Marine Ecology (vol. I, pt. 1, pp. 321–46). New York: Wiley.Google Scholar
Kirk, J. T. O. (2010). Light and Photosynthesis in Aquatic Ecosystems, 3rd edn. Cambridge: Cambridge University Press. 649 pp. [, , , ]CrossRefGoogle Scholar
Kirkman, H. (1981). The first year in the life history and the survival of the juvenile marine macrophyte, Ecklonia radiata (Turn.) 1. Agardh. J. Exp. Mar. Biol. Ecol. 55:243–54.CrossRefGoogle Scholar
Kirst, G. O. (1988). Turgor pressure regulation in marine macroalgae. In Lobban, C. S., Chapman, D. J., and Kremer, B. P. (eds), Experimental Phycology: A Laboratory Manual (pp. 203–9). Cambridge: Cambridge University Press.Google Scholar
Kirst, G. O. (1990). Salinity tolerance of eukaryotic marine algae. Annu. Rev. Plant Physiol. Plant Mol. Biol. 41:21–53.CrossRefGoogle Scholar
Kirst, G. O., and Wiencke, C. (1995). Ecophysiology of polar algae. J. Phycol. 31:181–99.CrossRefGoogle Scholar
Kitade, Y., Nakamura, M., Uji, T., et al. (2008). Structural features and gene-expression profiles of actin homologs in Porphyra yezoensis (Rhodophyta). Gene 423:79–84.CrossRefGoogle Scholar
Kitagawa, D., Vakonakis, I., Olieric, N., et al. (2011). Structural basis of the 9-fold symmetry of centrioles. Cell 144:364–75.CrossRefGoogle ScholarPubMed
Klenell, M., Snoeijs, P., and Pedersen, M. (2002) The involvement of a plasma membrane H+-ATPase in the blue-light enhancement of photosynthesis in Laminaria digitata (Phaeophyta). J. Phycol. 38:1143–9.CrossRefGoogle Scholar
Klerks, P. L., and Weis, J. S. (1987). Genetic adaptation to heavy metals in aquatic organisms: a review. Environ. Pollut. 45:173–205.CrossRefGoogle ScholarPubMed
Kling, R., and Bodard, M. (1986). La construction du thalle de Gracilaria verrucosa (Rhodophyceae, Gigartinales): edification de la fronde; essai d’interpretation phylogenetique. Crypt. Algol. 7:231–46.Google Scholar
Klinger, T., and De Wreede, R. E. (1988). Stipe rings, age, and size in populations of Laminaria setchellii Silva (Laminariales, Phaeophyta) in British Columbia, Canada. Phycologia 27:234–40.CrossRefGoogle Scholar
Kloareg, B., and Quatrano, B. R. S. (1988). Structure of the cell walls of marine algae and ecophysiological functions of the matrix polysaccharides. Oceanogr. Mar. Biol. Annu. Rev. 26:259–315.Google Scholar
Kloareg, B., Demarty, M., and Mabeau, S. (1986). Polyanionic characteristics of purified sulphated homofucans from brown algae. Int. J. Biol. Macromol. 8:380–6.CrossRefGoogle Scholar
Kloareg, B., Demarty, M., and Mabeau, S. (1987). Ion-exchange properties of isolated cell walls of brown algae: the interstitial solution. J. Exp. Bot. 38:1652–62.CrossRefGoogle Scholar
Klotchkova, T. A., Chah, O.-K., West, J. A., and Kim, G. H. (2003). Cytochemical and ultrastructural studies on protoplast formation from disintegrated cells of the marine alga Chaetomorpha aerea (Chlorophyta). Eur. J. Phycol. 38:205–16.CrossRefGoogle Scholar
Klumpp, D. W., and McKinnon, A. D. (1989). Temporal and spatial patterns in primary productivity of a coral reef epilithic algal community. J. Exp. Mar. Biol. Ecol. 131:1–22.CrossRefGoogle Scholar
Klumpp, D. W., and Polunin, N. V. C. (1989). Partitioning among grazers of food resources within damselfish territories on a coral reef. J. Exp. Mar. Biol. Ecol. 125:145–69.CrossRefGoogle Scholar
Klumpp, D. W., McKinnnon, D., and Daniel, P. (1987). Damselfish territories: zones of high productivity on coral reefs. Mar. Ecol. Prog. Ser. 40:41–51.CrossRefGoogle Scholar
Knight, M., and Parke, M. (1950). A biological study of Fucus vesiculosus L. and F serratus L. J. Mar. Biol. Assoc. UK 29:439–514.CrossRefGoogle Scholar
Knoop, W. T., and Bate, G. C. (1990). A model for the description of photosynthesis-temperature responses by subtidal Rhodophyta. Bot. Mar. 33:165–71.CrossRefGoogle Scholar
Knoth, A., and Wiencke, C. (1984). Dynamic changes of protoplasmic volume and of fine structure during osmotic adaptation in the intertidal red alga Porphyra umbilicalis. Plant Cell Envir. 7:113–19.CrossRefGoogle Scholar
Koch, E. W. (1994). Hydrodynamics, diffusion-boundary layers and photosynthesis of the seagrasses Thalassia testudinum and Cymodocea nodosa. Mar. Biol. 118:767–76.CrossRefGoogle Scholar
Koehl, M. A. R. (1982). The interaction of moving water and sessile organisms. Sci. Am. 247(6): 124–35.CrossRefGoogle Scholar
Koehl, M. A. R. (1986). Seaweeds in moving water: form and mechanical function. In Givnish, T. J. (ed.), On the Economy of Plant Form and Function (pp. 603–34). Cambridge: Cambridge University Press.Google Scholar
Koehl, M. A. R., and Alberte, R. S. (1988). Flow, flapping, and photosynthesis of Nereocystis luetkeana: a functional comparison of undulate and flat blade morphologies. Mar. Biol. 99:435–44. [, , ]CrossRefGoogle Scholar
Koehl, M. A. R., and Wainwright, S. A. (1977). Mechanical adaptations of a giant kelp. Limnol. Oceanogr. 22:1067–71.CrossRefGoogle Scholar
Koehl, M. A. R., and Wainwright, S. A. (1985). Biomechanics. In Littler, M. M. and Littler, D. S. (eds) Handbook of Phycological Methods: Ecological Field Methods: Macroalgae (pp. 291–313). Cambridge: Cambridge University Press.Google Scholar
Koehl, M. A. R., Silk, W. K., Liang, H., and Mahadevan, L. (2008). How kelp produce blade shapes suited to different flow regimes: a new wrinkle. Integr. Comp. Biol. 48:834–51.CrossRefGoogle ScholarPubMed
Kohlmeyer, J., and Kohlmeyer, E. (1972). Is Ascophyllum lichenized? Bot. Mar. 15:109–12.CrossRefGoogle Scholar
Kohlmeyer, J., and Hawkes, M. W. (1983). A suspected case of mycophycobiosis between Mycosphaerella apophlaeae (Ascomycetes) and Apophlaea spp. (Phodophyta). J. Phycol. 19:257–60.CrossRefGoogle Scholar
Kongelschatz, J., Solórzano, L., Barber, R., and Mendoza, P. (1985). Oceanographic conditions in the Galapagos Islands during the 1982/1983 El Niño. In Robinson, G. and del Pino, E. M. (eds) El Niño en las Islas Galapagos. El evento de 1982/1983. Fund (pp. 91–123). Quito: Charles Darwin.Google Scholar
Konstantinou, I. K., and Albanis, T. A. (2004). Worldwide occurrence and effects of antifouling paint booster biocides in the aquatic environment: a review. Environ. Internat. 30:235–48.CrossRefGoogle ScholarPubMed
Kooistra, W. H. C. F., Joosten, A. M. T., and van den Hoek, C. (1989). Zonation patterns in intertidal pools and their possible causes: a multivariate approach. Bot. Mar. 32:9–26.CrossRefGoogle Scholar
Kopczak, C. D., Zimmerman, R. C., and Kremer, J. N. (1991). Variation in nitrogen physiology and growth among geographically isolated populations of the giant kelp, Macrocystis pyrifera (Phaeophyta). J. Phycol. 27:149–58.CrossRefGoogle Scholar
Kopp, D., Bouchon-Navaro, Y., Cordonnier, S., et al. (2010). Evaluation of algal regulation by herbivorous fishes on Caribbean coral reefs. Helgoland Mar. Res. 64:181–90.CrossRefGoogle Scholar
Korb, R. E., and Gerard, V. A. (2000). Nitrogen assimilation characteristics of polar seaweeds from differing nutrient environments. Mar. Ecol. Prog. Ser. 198:83–92.CrossRefGoogle Scholar
Kornmann, P. (1970). Advances in marine phycology on the basis of cultivation. Helgol. Meeresunters. 20:39–61.CrossRefGoogle Scholar
Kornmann, P., and Sahling, P.-H. (1977). Marine algae of Helgoland: benthic green brown and red algae. Helgol. Meeresunters. 29:1–289.CrossRefGoogle Scholar
Korpinen, S., Jormalainen, V., and Honkanen, T. (2007). Bottom-up and cascading top-down control of macroalgae along a depth gradient. J. Exp. Mar. Biol. Ecol. 343:52–63.CrossRefGoogle Scholar
Kottmeier, S. T., and Sullivan, C. W. (1988). Sea ice microbial communities (SIMCO). Polar Biol. 8:293–304.CrossRefGoogle Scholar
Kraan, S., and Guiry, M. D. (2000). Molecular and morphological character inheritance in hybrids of Alaria esculenta and A. praelonga (Alariaceae, Phaeophyceae). Phycologia 39:554–9.CrossRefGoogle Scholar
Kraberg, A. C., and Norton, T. A. (2007). Effect of epiphytism on reproductive and vegetative lateral formation in the brown, intertidal seaweed Ascophyllum nodosum (Phaeophyceae). Phycol. Res. 55:17–24.CrossRefGoogle Scholar
Kraemer, G. P., and Chapman, D. J. (1991a). Biomechanics and alginic acid composition during hydrodynamic adaptation by Egregia menziesii (Phaeophyceae) juveniles. J. Phycol. 27:47–53.CrossRefGoogle Scholar
Kraemer, G. P., and Chapman, D. J. (1991b). Effects of tensile force and nutrient availability on carbon uptake and cell wall synthesis in blades of juvenile Egregia menziesii (Turn.) Aresch. (Phaeophyceae). J. Exp. Mar. Biol. Ecol. 149:267–77.CrossRefGoogle Scholar
Kraufvelin, P., Moy, F. E., Christie, H., and Bokn, T. L. (2006). Nutrient addition to experimental rocky shore communities revisited: delayed responses, rapid recovery. Ecosystems 9:1076–93.CrossRefGoogle Scholar
Kraufvelin, P., Lindholm, A., Pedersen, M. F., Kirkerud, L. A., and Bonsdorff, E. (2010). Biomass, diversity and production of rocky shore macroalgae at two nutrient enrichment and wave action levels. Mar. Biol. 157:29–47.CrossRefGoogle Scholar
Krause, G. H., and Weis, E. (1991). Chlorophyll fluorescence and photosynthesis: the basics. Annu. Rev. Plant Physiol. Plant Mol. Biol. 42: 313–49.CrossRefGoogle Scholar
Kregting, L. T., Hurd, C. L., Pilditch, C. A., and Stevens, C. L. (2008a). The relative importance of water motion on nitrogen uptake by the subtidal macroalga Adamsiella chauvinii (Rhodophyta) in winter and summer. J. Phycol. 44:320–30.CrossRefGoogle Scholar
Kregting, L. T., Hepburn, C. D., Hurd, C. L., and Pilditch, C. A. (2008b). Seasonal patterns of growth and nutrient status of the macroalga Adamsiella chauvinii (Rhodophyta) in soft sediment environments. J. Exp. Mar. Biol. Ecol. 360:94–102.CrossRefGoogle Scholar
Kregting, L. T., Stevens, C. L., Cornelisen, C. D., Pilditch, C. A., and Hurd, C. L. (2011). Effects of a small-bladed macroalgal canopy on benthic boundary layer dynamics: implications for nutrient transport. Aquat. Biol. 14:41–56.CrossRefGoogle Scholar
Kreimer, G. (2001). Light perception and signal modulation during photoorientation of flagellate green algae. In Häder, D.-P. and Lebert, M. (eds), Photomovement; Comprehensive Series in Photosciences (Vol. 1, pp. 193–227). Amsterdam: Elsevier.Google Scholar
Kreimer, G., Kawai, H., Miiller, D. G., and Melkonian, M. (1991). Reflective properties of the stigma in male gametes of Ectocarpus siliculosus (Phaeophyceae) studied by confocal laser scanning microscopy. J. Phycol. 27:268–76.CrossRefGoogle Scholar
Kremer, B. P. (1977). Biosynthesis of polyols in Pelvetia canaliculata. Z. Pflanzenphysiol. 81:68–73.CrossRefGoogle Scholar
Kremer, B. P. (l981a). Carbon metabolism. In Lobban, C. S. and Wynne, M. J. (eds) The Biology of Seaweeds (pp. 493–533). Oxford: Blackwell Scientific. [, , ]Google Scholar
Kremer, B. P. (l981b). Metabolic implications of non-photosynthetic carbon fixation in brown macroalgae. Phycologia 20:242–50.CrossRefGoogle Scholar
Kremer, B. P. (1983). Carbon economy and nutrition of the alloparasitic red algaHarveyella mirabilis. Mar. Biol. 76:231–9.CrossRefGoogle Scholar
Kremer, B. P. (1985). Aspects of cellular compartmentation in brown marine macroalgae. J. Plant Physiol. 120:401–7.CrossRefGoogle Scholar
Kremer, B. P., and Markham, J. W. (1982). Primary metabolic effects of cadmium in the brown alga. Laminaria saccharina. Z. Pflanzenphysiol. 1008:125–30.CrossRefGoogle Scholar
Kroeker, K. J., Kordas, R. J., Crim, R. N., and Singh, G. G. (2010). Meta-analysis reveals negative yet variable effects of ocean acidification on marine organisms. Ecol. Lett. 13: 1419–34.CrossRefGoogle ScholarPubMed
Krom, M. D., Ellner, S., van Rijin, J., and Neori, A. (1995). Nitrogen and phosphorus cycling and transformations in a prototype “non-polluting” integrated mariculture system, Eilat, Israel. Mar. Ecol. Prog. Ser. 118:25–36.CrossRefGoogle Scholar
Krumhansl, K. A., Lee, J. M. and Scheibling, R. S. (2011). Grazing damage and encrustation by an invasive bryozoan reduce the ability of kelps to withstand breakage by waves. J. Exp. Mar. Biol. Ecol. 407:12–18.CrossRefGoogle Scholar
Kubanek, J. K., Lester, S. E., Fenical, W., and Hay, M. E. (2004). Ambiguous role of phlorotannins as chemical defenses in the brown alga Fucus vesiculosus. Mar. Ecol. Prog. Ser. 277:79–93.CrossRefGoogle Scholar
Kübler, J. E., and Dudgeon, S. R. (1996). Temperature dependent change in the complexity of form of Chondrus crispus fronds. J. Exp. Mar. Biol. Ecol. 207:15–24.CrossRefGoogle Scholar
Kübler, J. E., and Raven, J. A. (1994). Consequences of light limitation for carbon acquisition in three rhodophytes. Mar. Ecol. Prog. Ser. 110:203–9.CrossRefGoogle Scholar
Kübler, J. E., and Raven, J. A. (1996). Nonequilibrium rates of photosynthesis and respiration under dynamic light supply. J. Phycol. 32:963–9.CrossRefGoogle Scholar
Kübler, J. E., Johnston, A. M., and Raven, J. A. (1999). The effects of reduced and elevated CO2 and O2 on the seaweed Lomentaria articulata. Plant Cell Environ. 22:1303–10.CrossRefGoogle Scholar
Kucera, H., and Saunders, G. W. (2008). Assigning morphological variants of Fucus (Fucales, Phaeophyceae) in Canadian waters to recognized species using DNA barcoding. Botany 86:1065–79.CrossRefGoogle Scholar
Kuffner, I. B., Andersson, A. J., Jokel, P. L., and Rogers, K. S. (2008). Decreased abundance of crustose coralline algae due to ocean acidification. Nature Geosci. 1:114–17.CrossRefGoogle Scholar
Kumar, M., Kumari, P., Gupta, V., Reddy, C. R. K. and Jha, B. (2010a). Biochemical responses of red alga Gracilaria corticata (Gracilariales, Rhodophyta) to salinity induced oxidative stress. J. Exp. Mar. Biol. Ecol. 391:27–34.CrossRefGoogle Scholar
Kumar, M., Kumari, P., Gupta, V., Anisha, P. A., and Reddy, C. R. K. (2010b). Differential responses to cadmium induced oxidative stress in marine macroalga Ulva lactuca (Ulvales, Chlorophyta). Biometals 23:315–25.CrossRefGoogle Scholar
Kumar, M., Gupta, V., Trivedi, N., et al. (2011). Desiccation induced oxidative stress and its biochemical responses in intertidal red alga Gracilaria corticata (Gracilariales, Rhodophyta). Envir. Exp. Bot. 72:194–201.CrossRefGoogle Scholar
Kumar, M., Bijo, A. J., Baghel, R. S., Reddy, C. R. K., and Jha, B. (2012). Selenium and spermine alleviate cadmium induced toxicity in the red seaweed Gracilaria dura by regulating antioxidants and DNA methylation. Plant Physiol. Biochem. 51:129–38.CrossRefGoogle ScholarPubMed
Küpper, F. C., Schweigert, N., Ar Gall, E., et al. (1998). Iodine uptake in Laminariales involves extracellular, haloperoxidase-mediated oxidation of iodide. Planta 207:163–71.Google Scholar
Küpper, F. C., Müller, D. G., Peters, A. F., Kloareg, B., and Potin, P. (2002a). Oligoalginate recognition and oxidative burst play a key role in natural and induced resistance of sporophytes of Laminariales. J. Chem. Ecol. 28:2057–81.CrossRefGoogle ScholarPubMed
Küpper, H., Šetlik, I., Spiller, M., Küpper, F. C., and Prášil, O. (2002b). Heavy metal-induced inhibition of photosynthesis: targets of in vivo heavy metal chlorophyll formation. J. Phycol. 38:429–41.CrossRefGoogle Scholar
Küpper, F. C., Carpenter, L. J., McFiggans, G. B., et al. (2008). Iodine accumulation provides kelp with inorganic antioxidant impacting atmospheric chemistry. PNAS 105:6954–8.CrossRefGoogle ScholarPubMed
Küppers, U., and Kremer, B. P. (1978). Longitudinal profiles of carbon dioxide fixation capacities in marine macroalgae. Plant Physiol. 62:49–53.CrossRefGoogle ScholarPubMed
Küppers, U. and Weidner, M. (1980). Seasonal variation of enzyme activities inLaminaria hyperborea. Planta 148:222–30.Google ScholarPubMed
Kurihara, A., Abe, T., Tani, M., and Sherwood, A. R. (2010). Molecular phylogeny and evolution of red algal parasites: a case study of Benzaitenia, Janczewskia, and Ululania (Ceramiales). J. Phycol. 46:580–90.CrossRefGoogle Scholar
Kurle, C. M., Croll, D. A., and Tershy, B. R. (2008). Introduced rats indirectly change marine rocky intertidal communities from algae- to invertebrate-dominated. PNAS 105:3800–4.CrossRefGoogle ScholarPubMed
Kurogi, M., and Hirano, K. (1956). Influences of water temperature on the growth, formation of monosporangia and monospore-liberation in the Conchocelis-phase of Porphyra tenera Kjellm. Bull. Tohoku Reg. Fish. Res. Lab. 8:45–61.Google Scholar
Kutser, T., Vahtmäe, E., and Martin, G. (2006). Assessing suitability of multispectral satellites for mapping benthic macroalgal cover in turbid coastal waters by means of model simulations. Estuar. Coast. Shelf S. 67:521–9.CrossRefGoogle Scholar
Kuwano, K., Sakurai, R., Motozu, Y., Kitade, Y., and Saga, N. (2008). Diurnal cell division regulated by gating the G1/S transition in Enteromorpha compressa (Chlorophyta). J. Phycol. 44:364–73.CrossRefGoogle Scholar
Kylin, H. (1956). Die Gattungen der Rhodophyceen. Lund, Sweden: Gleerups Förlag.Google Scholar
La Claire, J. W. (l982a). Cytomorphological aspects of wound healing in selected Siphonocladales (Chlorophyceae). J. Phycol. 18:379–84.CrossRefGoogle Scholar
La Claire, J. W. (l982b). Wound-healing motility in the green alga Ernodesmis: calcium ions and metabolic energy are required. Planta 156:466–74.CrossRefGoogle Scholar
La Claire, J. W. (1989). Actin cytoskeleton in intact and wounded coenocytic green algae. Planta 177:47–57.CrossRefGoogle ScholarPubMed
La Claire, J. W., and Wang, J. (2000). Localization of plasmidlike DNA in giant-celled marine green algae. Protoplasma 213:157–64.CrossRefGoogle Scholar
Ladah, L. B., and Zertuche-González, J. A. (2007). Survival of microscopic stages of a perennial kelp (Macrocystis pyrifera) from the center and the southern extreme of its range in the northern hemisphere after exposure to simulated El Niño stress. Mar. Biol. 152:677–86.CrossRefGoogle Scholar
Ladah, L. B., Feddersen, F., Pearson, G. A., and Serrão, E. A. (2008). Egg release and settlement patterns of dioecious and hermaphroditic fucoid algae during the tidal cycle. Mar. Biol. 155:583–91.CrossRefGoogle Scholar
Lago-Lestón, A., Mota, C., Kautsky, L., and Pearson, P. A. (2010). Functional divergence in heat shock response following rapid speciation of Fucus spp. in the Baltic Sea. Mar. Biol. 157: 683–8.CrossRefGoogle Scholar
Lahaye, M., and Robic, A. (2007). Structure and functional properties of ulvan, a polysaccharide from green seaweeds. Biomacromol. 8:1765–74. [, , ]CrossRefGoogle ScholarPubMed
Lane, C. E., and Saunders, G. W. (2005). Molecular investigation reveals epi/endophytic extrageneric kelp (Laminariales, Phaeophyceae) gametophytes colonizing Lessoniopsis littoralis thalli. Bot. Mar. 48:426–36.CrossRefGoogle Scholar
Lane, C. E., Mayes, C., Druehl, L. D., and Saunders, G. W. (2006). A multi-gene molecular investigation of the kelp (Laminariales, Phaeophyceae) supports substantial taxonomic re-organization. J. Phycol. 42: 493–512.CrossRefGoogle Scholar
Lane, C.E., Lindstrom, S., and Saunders, G. W. (2007). A molecular assessment of northeast Pacific Alaria species (Laminariales, Phaeophyceae) with reference to the utility of DNA barcoding. Mol. Phylogenet. Evol. 44:634–48.CrossRefGoogle ScholarPubMed
Lang, J. C. (1974). Biological zonation at the base of a reef. Amer. Sci. 62: 271–81.Google Scholar
Langer, G., Nehrke, G., Probert, I., Ly, J., and Ziveri, P. (2009). Strain-specific responses of Emiliania huxleyi to changing seawater carbonate chemistry. Biogeosci. 6: 2637–46.CrossRefGoogle Scholar
Langford, T. E. L. (1990). Ecological Effects of Thermal Discharges. New York: Elsevier Applied Science, 468 pp.Google Scholar
Langston, W. J. (1990). Toxic effects of metals and the incidence of metal pollution in marine ecosystems. In Furness, R. W. and Rainbow, P.S. (eds), Heavy Metals in the Marine Environment (pp. 101–22). Boca Raton, FL: CRC Press.Google Scholar
Lapointe, B. E. (1985). Strategies for pulsed nutrient supply to Gracilaria cultures in the Florida Keys: interactions between concentration and frequency of nutrient pulses. J. Exp. Mar. Biol. Ecol. 93:211–22.CrossRefGoogle Scholar
Lapointe, B. E. (1986). Phosphorus-limited photosynthesis and growth of Sargassum natans and Sargassum fluitans (Phaeophyceae) in the western North Atlantic. Deep Sea Res. 33:391–9.CrossRefGoogle Scholar
Lapointe, B. E. (1987). Phosphorus- and nitrogen-limited photosynthesis and growth of Gracilaria tikvahiae (Rhodophyceae) in the Florida Keys: an experimental field study. Mar. Biol. 93:561–8.CrossRefGoogle Scholar
Lapointe, B. E. (1997). Nutrient thresholds for bottom-up control of macroalgal blooms on coral reefs in Jamaica and southeast Florida. Limnol. Oceanogr. 42:1119–31. [, , ]CrossRefGoogle Scholar
Lapointe, B. E. (1999). Simultaneous top-down and bottom-up forces control macroalgal blooms on coral reefs (Reply to comment by Hughes et al.). Limnol. Oceanogr. 44:1586–92.CrossRefGoogle Scholar
Lapointe, B. E., and Bedford, B. J. (2010). Ecology of nutrition of invasive Caulerpa brachypus f. parvifolia blooms on coral reefs off southeast Florida, USAHarmful Algae 9:1–12.CrossRefGoogle Scholar
Lapointe, B. E., and Duke, C. S. (1984). Biochemical strategies for growth of Gracilaria tikvahiae (Rhodophyta) in relation to light intensity and nitrogen availability. J. Phycol. 20:488–95.CrossRefGoogle Scholar
Lapointe, B. E., and O’Connell, J. (1989). Nutrient-enhanced growth of Cladophora prolifera in Harrington Sound, Bermuda: eutrophication of a confined, phosphorus-limited marine ecosystem. Estu. Cstl. Shelf Sci. 28:347–60.CrossRefGoogle Scholar
Lapointe, B. E., and Ryther, J. H. (1979). The effects of nitrogen and seawater flow rate on the growth and biochemical composition of Gracilaria foliifera v. angustissima in mass outdoor cultures. Bot. Mar. 22:529–37.CrossRefGoogle Scholar
Lapointe, B. E., Littler, M. M., and Littler, D. S. (1987). A comparison of nutrient-limited productivity in macroalgae from a Caribbean barrier reef and from a mangrove ecosystem. Aquat. Bot. 28:243–55.CrossRefGoogle Scholar
Lapointe, B. E., Littler, M. M., and Littler, D. S. (1992). Nutrient availability to marine macroalgae in siliciclastic versus carbonate-rich coastal waters. Estuaries 15:75–82. [, , , ]CrossRefGoogle Scholar
Larkum, A. W. D., and Barrett, J. (1983). Light-harvesting processes in algae. Adv. Bot. Res. 10:1–219. [, , , ]CrossRefGoogle Scholar
Larkum, A. W. D., and Kühl, M. (2005). Chlorophyll d: the puzzle resolved. Trends Plant Sci. 10:355–7.CrossRefGoogle ScholarPubMed
Larkum, A. W. D., and Vesk, M. (2003). Algal plastids: their fine structure and properties. In Larkum, A. W. D., Douglas, S. E., and Raven, J. A. (eds), Photosynthesis in Algae. Advances in Photosynthesis and Respiration (Vol. 14, pp. 11–28). Dordrecht, The Netherlands:Kluwer Academic Publishers. [, , ]CrossRefGoogle Scholar
Larkum, A. W. D., Drew, E. A., and Crossett, R. N. (1967). The vertical distribution of attached marine algae in Malta. J. Ecol. 55:361–71.CrossRefGoogle Scholar
Larkum, A. W. D., Orth, R. J., and Duarte, C. M. (eds) (2006). Seagrasses: Biology, Ecology and Conservation. Dordrecht, The Netherlands: Springer, 691 pp.Google Scholar
Larkum, A. W. D., Salih, A., and Kühl, M. (2011). Rapid mass movement of chloroplasts during segment formation of the calcifying siphonalean green alga Halimeda macroloba. PloS ONE 6:e20841.CrossRefGoogle ScholarPubMed
Larned, S. T., and Atkinson, M. J. (1997). Effects of water velocity on NH4 and PO4 uptake and nutrient-limited growth in the macroalgaDictyosphaeria cavernosa. Mar. Ecol. Prog. Ser. 157:295–302.CrossRefGoogle Scholar
Larsen, B., Haug, A., and Painter, T. J. (1970). Sulphated polysaccharides in brown algae. III. The native state of fucoidan in Ascophyllum nodosum and Fucus vesiculosus. Acta Chem. Scand. 24:3339–52.CrossRefGoogle Scholar
Lartigue, J., and Sherman, T. D. (2005). Response of Enteromorpha sp. (Chlorophyceae) to a nitrate pulse: nitrate uptake, inorganic nitrogen storage and nitrate reductase activity. Mar. Ecol. Prog. Ser. 292:147–57.CrossRefGoogle Scholar
Lary, D. J. (1997). Catalytic destruction of stratospheric ozone. J. Geophys. Res. 102(21):515–26.CrossRefGoogle Scholar
Lasley-Rasher, R. S., Rasher, D. B., Marion, A. H., Taylor, R. B., and Hay, M. E. (2011). Predation constrains host choice for a marine mesograzer. Mar. Ecol. Prog. Ser. 434:91–9.CrossRefGoogle Scholar
Lassuy, D. R. (1980). Effects of “farming” behavior by Eupomacentrus lividus and Hemiglyphidodon plagiometopon on algal community structure. Bull. Mar. Sci. 30:304–12.Google Scholar
Lauzon-Guay, J.-S., and Scheibling, R. E. (2007). Seasonal variation in movement, aggregation and destructive grazing of the green sea urchin (Strongylocentrotus droebachiensis) in relation to wave action and sea temperature. Mar. Biol. 151:2109–18.CrossRefGoogle Scholar
Lauzon-Guay, J.-S., and Scheibling, R. E. (2010). Spatial dynamics, ecological thresholds and phase shifts: modeling grazer aggregation and gap formation in kelp beds. Mar. Ecol. Prog. Ser. 403:29–41.CrossRefGoogle Scholar
Lavery, P. S., and McComb, A. J. (l991a). Macroalgal–sediment nutrient interactions and their importance to macroalgal nutrition in a eutrophic estuary. Estu. Cstl. Shelf Sci. 32:281–95.CrossRefGoogle Scholar
Lavery, P. S., and McComb, A. J. (l991b). The nutritional ecophysiology of Chaetomorpha linum and Ulva rigida in Peel Inlet, Western Australia. Bot. Mar. 34:251–60. [, , , ]CrossRefGoogle Scholar
Lavery, P. S., and Vanderklift, M. A. (2002). A comparison of spatial and temporal patterns in epiphytic macroalgal assemblages of the seagrasses Amphibolis griffithii and Posidonia coriacea. Mar. Ecol. Prog. Ser. 236:99–112.CrossRefGoogle Scholar
Lavery, P. S., Lukatelich, R. J., and McComb, A. J. (1991). Changes in the biomass and species composition of macroalgae in a eutrophic estuary. Estu. Cstl. Shelf Sci. 33:1–22.CrossRefGoogle Scholar
Laws, E. A. (2000). Aquatic Pollution: An Introductory Text. New York: J. Wiley, 655 pp. [, , , , ]Google Scholar
Laycock, M. V., and Craigie, J. S. (1977). The occurrence and seasonal variation of gigartinine and L-citrullinyl-L-arginine in Chondrus crispus Stackh. Can. J. Biochem. 55:27–30.CrossRefGoogle ScholarPubMed
Laycock, M. V., Morgan, K. C., and Craigie, J. S. (1981). Physiological factors affecting the accumulation of L-citrullinyl-L-arginine in Chondrus crispus. Can. J. Bot. 59:522–7.CrossRefGoogle Scholar
Le Bail, A., Billoud, B., Maisonneuve, C., et al. (2008). Early development pattern of the brown alga Ectocarpus siliculosus (Ectocarpales, Phaeophyceae) sporophyte. J. Phycol. 44:1269–81.CrossRefGoogle ScholarPubMed
Le Bail, A., Billoud, B., Kowalczyk, N., et al. (2010). Auxin metabolism and function in the multicellular brown alga Ectocarpus siliculosus. Plant Physiol. 153:128–44.CrossRefGoogle ScholarPubMed
Le Bail, A., Billoud, B., Le Panse, S., Chenivesse, S., and Charrier, B. (2011). ETOILE regulates developmental patterning in the filamentous brown alga Ectocarpus siliculosus. The Plant Cell 23:1666–78.CrossRefGoogle ScholarPubMed
Le Corguillé, G., Pearson, G., Valente, M., Viegas, C., et al. (2009). Plastid genomes of two brown algae, Ectocarpus siliculosus and Fucus vesiculosus: further insights on the evolution of red algal derived plastids. BMC Evol. Biol. 9:240–53.CrossRefGoogle ScholarPubMed
Le Gall, L., and Saunders, G. W. (2007). A nuclear phylogeny of the Florideophyceae (Rhodophyta) inferred from combined EF2, small subunit and large subunit ribosomal DNA: establishing the new red algal subclass Corallinophycidae. Mol. Phylogenet. Evol. 43:1118–30.CrossRefGoogle ScholarPubMed
Le Gall, L., and Saunders, G. W. (2010). DNA barcoding is a powerful tool to uncover algal diversity: a case study of the Phyllophoraceae (Gigartinales, Rhodophyta) in the Canadian flora. J. Phycol. 46:374–89.CrossRefGoogle Scholar
Le Gall, L., Rusig, A.-M., and Cosson, J. (2004). Organisation of the microtubular cytoskeleton in protoplasts from Palmaria palmate (Palmariales, Rhodophyta). Bot. Mar. 47:231–7.Google Scholar
Leal, M. F. C., Vasconcelos, M. T., Sousa-Pinto, I., and Cabral, J. P. S. (1997). Biomonitoring with benthic macroalgae and direct assay of heavy metals in seawater of the Oporto coast (NW Portugal). Mar. Poll. Bull. 34:1006–15.CrossRefGoogle Scholar
Lechat, H., Amat, M., Mazoyer, J., et al. (2000). Structure and distribution of glucomannan and sulfated glucan in the cell walls of the red alga Kappaphycus alvarezii (Gigartinales, Rhodophyta). J. Phycol. 36:891–902.CrossRefGoogle Scholar
Ledlie, M., Graham, N., Bythell, J., et al. (2007). Phase shifts and the role of herbivory in the resilience of coral reefs. Coral Reefs 26:641–53.CrossRefGoogle Scholar
Lee, S.-H., Motomura, T., and Ichimura, T. (2002). Light and electron microscopic observations of preferential destruction of chloroplast and mitochondrial DNA at early male gametogenesis of the anisogamous green alga Derbesia tenuissima (Chlorophyta). J. Phycol. 38:534–42.CrossRefGoogle Scholar
Lee, T. M. (1998). Investigations of some intertidal green macroalgae to hyposaline stress: detrimental role of putrescine under extreme hyposaline conditions. Plant Sci. 138:1–8.CrossRefGoogle Scholar
Lee, T. M., Tsac, P.-F., Shyu, Y.T., and Sheu, F. (2005). The effects of phosphate on phosphate starvation responses of Ulva lactuca (Ulvales, Chlorophyta). J. Phycol. 41:975–82.CrossRefGoogle Scholar
Lehninger, A. L. (1975). Biochemistry, 2nd edn. New York: Worth, 1104 pp.Google Scholar
Leichter, J. J., Stokes, M. D., and Genovese, S. J. (2008). Deep water macroalgal communities adjacent to the Florida Keys reef tract. Mar. Ecol. Prog. Ser. 356:123–38.CrossRefGoogle Scholar
Leigh, E. G., Paine, 1 R. T.. Quinn, F., and Suchanek, T. H. (1987). Wave energy and intertidal productivity. Proc. Natl. Acad. Sci. USA 84:1314–18. [8.0, ]CrossRefGoogle ScholarPubMed
Leighton, D. L. (1971). Grazing activities of benthic invertebrates in southern California kelp beds. Nova Hedwegia Beiheft 32:421–53.Google Scholar
Leighton, D. L., Jones, L. G., and North, W. J. (1966). Ecological relationships between giant kelp and sea urchins in southern California. Proc. Intl. Seaweed Symp. 5:141–53.Google Scholar
Lenanton, R. C. J., Robertson, A. I., and Hansen, J. A. (1982). Nearshore accumulations of detached macrophytes as nursery areas for fish. Mar. Ecol. Prog. Ser. 9:51–7.CrossRefGoogle Scholar
Leonardi, P. I., and Vasquez, J. A. (1999). Effects of copper on the ultrastructure of Lessonia spp. Hydrobiol. 398/399:375–83.CrossRefGoogle Scholar
Leonardi, P. I., Miravalles, A. B., Faugeron, S. et al. (2006). Diversity, phenomenology and epidemiology of epiphytism in farmed Gracilaria chilensis (Rhodophyta) in northern Chile. Eur. J. Phycol. 41:247–57.CrossRefGoogle Scholar
Lepoint, G., Nyssen, F., Gobert, S., Dauby, P., and Bouquegneau, J. M. (2000). Relative impact of a seagrass bed and its adjacent epilithic algal community in consumer diets. Mar. Biol. 136(3):513–18.CrossRefGoogle Scholar
Leroux, S. J., and Loreau, M. (2008). Subsidy hypothesis and strength of trophic cascades across ecosystems. Ecology Letters 11:1147–56.CrossRefGoogle ScholarPubMed
Lesser, M. P. (2006). Oxidative stress in marine environments: biochemistry and physiological ecology. Annu. Rev. Physiol. 68:253–78.CrossRefGoogle ScholarPubMed
Lessios, H. A. (1988). Mass mortality of Diadema antillarum in the Caribbean: what have we learned? Annu. Rev. Ecol. Syst. 19:371–93.CrossRefGoogle Scholar
Lessios, H. A. (2005). Diadema antillarum populations in Panama twenty years following mass mortality. Coral Reefs 24:125–7.CrossRefGoogle Scholar
Leustek, T., Martin, M. N., Bick, J-A. and Davies, J. P. (2000). Pathways and regulation of sulfur metabolism revealed through molecular and genetic studies. Annu. Rev. Plant Phys. 51:141–65.CrossRefGoogle ScholarPubMed
Levenbach, S. (2008). Behavioral mechanism for an associational refuge for macroalgae on temperate reefs. Mar. Ecol. Prog. Ser.370:45–52.CrossRefGoogle Scholar
Levin, S. A. (1992). The problem of pattern and scale in ecology. Ecology 73:1943–67.CrossRefGoogle Scholar
Lewis, J. R. (1964). The Ecology of Rocky Shores. London: English Universities Press, 323 pp. [, , , ]Google Scholar
Lewis, J. R. (1980). Objectives in littoral ecology: a personal viewpoint. In Price, J. H., Irvine, D. E. G., and Farnham, W. F. (eds), The Shore Environment. Vol. 1: Methods (pp. 118). New York: Academic Press.Google Scholar
Lewis, S. M., Norris, J. N., and Searles, R. B. (1987). The regulation of morphological plasticity in tropical reef algae by herbivory. Ecology 68:636–41.CrossRefGoogle Scholar
Lewis, S. E., Brodie, J. E., Bainbridge, Z. T., et al. (2009). Herbicides: a new threat to the Great Barrier Reef. Environ. Poll. 157:2470–84.CrossRefGoogle ScholarPubMed
Li, N., and Cattolico, R. A. (1987). Chloroplast genome characterization in the red alga Griffithsia pacifica. Mol. Gen. Genet. 209:343–51.CrossRefGoogle ScholarPubMed
Li, T., Wang, C., and Miao, J. (2007). Identification and quantification of indole-3-acetic acid in the kelp Laminaria japonica Areschoug and its effect on growth of marine microalgae. J. Appl. Phycol. 19:479–84.CrossRefGoogle Scholar
Libes, S. M. (1992). Introduction to Marine Biogeochemistry (2nd edn). Amsterdam: Elsevier, 909 pp.Google Scholar
Lignell, A., and Pedersen, M. (1987). Nitrogen metabolism in Gracilaria secundata. Hydrobiologia 151/152:431–41.CrossRefGoogle Scholar
Lilley, S. A., and Schiel, D. R. (2006). Community effects following the deletion of a habitat-forming alga from rocky marine shores. Oecologia 148:672–81.CrossRefGoogle ScholarPubMed
Lin, D. T., and Fong, P. (2008). Macroalgal bioindicators (growth, tissue N, δ15N) detect nutrient enrichment from shrimp farm effluent entering Opunohu Bay, Moorea, French Polynesia. Mar. Poll. Bull. 56:245–9.CrossRefGoogle Scholar
Lin, R., and Stekoll, M. S. (2007). Effects of plant growth substances on the conchocelis phase of Alaskan Porphyra (Bangiales, Rhodophyta) species in conjunction with environmental variables. J. Phycol. 43:1094–103.CrossRefGoogle Scholar
Lin, T. Y., and Hassid, W. Z. (1966). Pathway of alginic acid synthesis in the marine brown alga, Fucus gardneri Silva. J. Biol Chem 241: 5284–97.Google Scholar
Lindstrom, S. C. (2008). Cryptic diversity and phylogenetic relationships within the Mastocarpus papillatus species complex (Rhodophyta, Phyllophoraceae). J. Phycol. 44:1300–8.CrossRefGoogle Scholar
Ling, S. D., and Johnson, C. R. (2009). Population dynamics of an ecologically important range-extender: kelp beds versus sea urchin barrens. Mar. Ecol. Prog. Ser. 374:113–25.CrossRefGoogle Scholar
Littler, M. M. (1979). The effects of bottle volume, thallus weight, oxygen saturation levels, and water movement on apparent photosynthetic rates in marine algae. Aquat. Bot. 7:21–34.CrossRefGoogle Scholar
Littler, M. M., and Arnold, K. E. (1980). Sources of variability in macroalgal primary productivity: sampling and interpretative problems. Aquat Bot 8: 141–56.CrossRefGoogle Scholar
Littler, M. M., and Arnold, K. E. (1982). Primary productivity of marine macroalgal functional-form groups from southwestern North America. J Phycol 18: 307–11.CrossRefGoogle Scholar
Littler, M. M., and Kauker, B. J. (1984). Heterotrichy and survival strategies in the red alga Corallina officinalis L. Bot. Mar. 27:37–44.CrossRefGoogle Scholar
Littler, M. M., and Littler, D. S. (1980). The evolution of thallus form and survival strategies in benthic marine macroalgae: field and laboratory tests of a functional form model. Am. Nat. 116:25–44. [, , , ]CrossRefGoogle Scholar
Littler, M. M., and Littler, D. S. (1984). Models of tropical reef biogenesis: the contribution of algae. Prog. Phycol. Res. 3:323–64.Google Scholar
Littler, M. M., and Littler, D. S. (1987). Effects of stochastic processes on rocky intertidal biotas: an unusual flash flood near Corona del Mar, California. Bull. So. Cal. Acad. Sci. 86:95–106.Google Scholar
Littler, M. M., and Littler, D. S. (1988). Structure and role of algae in tropical reef communities. In Lembi, C. A. and Waaland, J. R. (eds), Algae and Human Affairs (pp. 29–56). Cambridge: Cambridge University Press.Google Scholar
Littler, M. M., and Littler, D. S. (1990). Productivity and nutrient relationships in psammophytic versus epilithic forms of Bryopsidales (Chlorophyta): comparisons based on a short-term physiological assay. Hydrobiologia 204/205:49–55.CrossRefGoogle Scholar
Littler, M. M., and Littler, D. S. (1999). Blade abandonment/proliferation: a novel mechanism for rapid epiphyte control in marine macrophytes. Ecology 80:1736–46.CrossRefGoogle Scholar
Littler, M. M., and Littler, D. S. (2007). Assessment of coral reefs using herbivory/nutrient assays and indicator groups of benthic primary producers: a critical synthesis, proposed protocols, and critique of management strategies. Aq. Cons.: Mar. Fresh. Ecosyst. 17:195–215.CrossRefGoogle Scholar
Littler, M. M., and Littler, D. S. (2011a). Algae: macro. In Hopley, D. (ed.) Encyclopedia of Modern Coral Reefs: Structure, Form and Process (pp. 30–8). Berlin: Springer-Verlag.CrossRefGoogle Scholar
Littler, M. M., and Littler, D. S. (2011b). Algae, turf. In Hopley, D. (ed.) Encyclopedia of Modern Coral Reefs: Structure, Form and Process (pp. 38–9). Berlin: Springer-Verlag.CrossRefGoogle Scholar
Littler, M. M., and Littler, D. S. (2011c). Algae, coralline. In Hopley, D. (ed.) Encyclopedia of Modern Coral Reefs: Structure, Form and Process (pp. 20–30). Berlin: Springer-Verlag.CrossRefGoogle Scholar
Littler, M. M., and Littler, D. S. (2011d). Algae, blue-green boring. In Hopley, D. (ed.) Encyclopedia of Modern Coral Reefs: Structure, Form and Process (pp. 18–20). Berlin: Springer-Verlag. [, , ]CrossRefGoogle Scholar
Littler, M. M., and Murray, S. N. (1975). Impact of sewage on the distribution, abundance and community structure of rocky intertidal macro-organisms. Mar. Biol. 30:277–91.CrossRefGoogle Scholar
Littler, M. M., Littler, D. S., and Taylor, P R. (1983a). Evolutionary strategies in a tropical barrier reef system: functional-form groups of marine macroalgae. J. Phycol. 19:229–37. [, , ]CrossRefGoogle Scholar
Littler, M. M., Martz, D. R., and Littler, D. S. (1983b) Effects of recurrent sand deposition on rocky intertidal organisms: importance of substrate heterogeneity in a fluctuating environment. Mar. Ecol. Prog. Ser. 11: 129–39.CrossRefGoogle Scholar
Littler, M. M., Littler, D. S., Blair, S. M., and Norris, J. N. (1985). Deepest known plant life discovered on an uncharted seamount. Science 227:57–9. [, , , ]CrossRefGoogle ScholarPubMed
Littler, M. M., Littler, D. S., Blair, S. M., and Norris, J. N. (1986a). Deep-water plant communities from an uncharted seamount off San Salvador Island, Bahamas: distribution, abundance, and primary productivity. Deep Sea Res. 33:881–92.CrossRefGoogle Scholar
Littler, M. M., Taylor, P. R., and Littler, D. S. (1986b). Plant defense associations in the marine environment. Coral Reefs 5:63–71.CrossRefGoogle Scholar
Littler, M. M., Littler, D. S., and Taylor, P. R. (1987). Animal–plant defense associations: effects on the distribution and abundance of tropical reef macrophytes. J. Exp. Mar. Biol. Ecol. 105:107–21.CrossRefGoogle Scholar
Littler, M. M., Littler, D. S., and Lapointe, B. E. (1988). A comparison of nutrient- and light-limited photosynthesis in psammophytic versus epilithic forms of Halimeda (Caulerpales, Halimediaceae) from the Bahamas. Coral Reefs 6:219–25.CrossRefGoogle Scholar
Littler, M. M., Littler, D. S., and Brooks, B. L. (2006). Harmful algae on tropical coral reefs: bottom-up eutrophication and top-down herbivory. Harmful Algae 5:565–85.CrossRefGoogle Scholar
Liu, D., Keesing, J. K., Xing, Q., and Shi, P. (2009). World’s largest macroalgal bloom caused by expansion of seaweed aquaculture in China. Mar. Poll. Bull. 58:888–95.CrossRefGoogle Scholar
Liu, J., Dong, S., Liu, X., and Ma, S. (2000). Responses of the macroalgae Gracilaria tenuistipitata var. liui (Rhodophyta) to iron stress. J. Appl. Phycol. 12:605–12.CrossRefGoogle Scholar
Lobban, C. S. (l978a). The growth and death of the Macrocystis sporophyte (Phaeophyceae, Laminariales). Phycologia 17: 196–212.CrossRefGoogle Scholar
Lobban, C. S. (l978b). Translocation of 14C in Macrocystis pyrifera (giant kelp). Plant Physiol. 61: 585–9.CrossRefGoogle Scholar
Lobban, C. S. (1978c). Translocation of 14C in Macrocystis integrifolia (Phaeophyceae). 1. Phycol. 14:178–82.CrossRefGoogle Scholar
Lobban, C. S. (1989). Environmental factors, plant responses, and colony growth in relation to tube-dwelling diatom blooms in the Bay of Fundy, Canada, with a review of the biology of tube-dwelling diatoms. Diatom Res. 4:89–109.CrossRefGoogle Scholar
Lobban, C. S., and Baxter, D. M. (1983). Distribution of the red algal epiphyte Polysiphonia lanosa on its brown algal host Ascophyllum nodosum in the Bay of Fundy, Canada. Bot. Mar. 26:533–8.CrossRefGoogle Scholar
Lobban, C. S., and Harrison, P. J. (1994). Seaweed Ecology and Physiology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Lobban, C. S., and Jordan, R. W. (2010). Diatoms on coral reefs and in tropical marine lakes. In Smol, J. P. and Stoermer, E. F. (eds), The Diatoms: Applications for the Environmental and Earth Sciences (2nd edn) (pp. 346–56). Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Lodish, H., Berk, A., Kaiser, C. A., et al. (2008). Molecular Cell Biology. New York: W. H. Freeman, 973 pp.Google Scholar
Longstaff, B. J., Kildea, T., Runcie, J. W., et al. (2002). An in situ study of photosynthetic oxygen exchange and electron transport rate in the marine macroalga Ulva lactuca (Chlorophyta). Photosynth. Res. 74:281–93.CrossRefGoogle Scholar
Longtin, C. M., and Scrosati, R. A. (2009). Role of surface wounds and brown algal epiphytes in the colonization of Ascophyllum nodosum (Phaeophyceae) fronds by Vertebrata lanosa (Rhodophyta). J. Phycol. 45:535–9.CrossRefGoogle Scholar
Lopez-Figueroa, F., and Niell, X. (1990). Effects of light quality on chlorophyll and biliprotein accumulation in seaweeds. Mar. Biol. 104:321–7.CrossRefGoogle Scholar
López-Figueroa, F., Lindemann, P., Braslavsky, K., et al. (1989). Detection of a phytochrome-like protein in macroalgae. Bot. Acta 102:178–80.CrossRefGoogle Scholar
Lorb, R. E., and Gerard, V. A. (2000). Nitrogen assimilation characteristics of polar seaweeds from differing nutrient environments. Mar. Ecol Prog. Ser. 198:83–92.Google Scholar
Loreau, M., Mouquet, N., and Holt, R. D. (2003). Meta-ecosystems: a theoretical framework for a spatial ecosystem ecology. Ecol. Lett. 6:673–9.CrossRefGoogle Scholar
Lotze, H., and Schramm, W. (2000). Ecophysiological traits explain species dominance in macroalgal blooms. J. Phycol. 36:287–95.CrossRefGoogle Scholar
Lotze, H. K., Worm, B., and Sommer, U. (2000). Propagule banks, herbivory and nutrient supply control population development and dominance patterns in macroalgal blooms. Oikos 89:46–58.CrossRefGoogle Scholar
Lotze, H. K., Worm, B. and Sommer, U. (2001). Strong bottom-up and top-down control of early life stages of macroalgae. Limnol. Oceanogr. 46:749–57.CrossRefGoogle Scholar
Løvstad Holdt, S., and Kraan, S. (2011). Bioactive compounds in seaweed: Functional food applications and legislation. J. Appl. Phycol. 23:543–97.CrossRefGoogle Scholar
Lowe, R. J., Koseff, J. R., and Monismith, S. G. (2005). Oscillatory flow through submerged canopies: 1. Velocity structure. J. Geophys. Res. 110, C10016, 17 pp.Google Scholar
Loya, Y., and Rinkevich, B. (1980). Effects of oil pollution on coral reef communities. Mar. Ecol. Prog. Ser. 3:176–80.CrossRefGoogle Scholar
, F., , W., Tian, C., et al. (2011). The Bryopsis hypnoides plastid genome: multimeric forms and complete nucleotide sequence. PloS ONE. 6:e14663.CrossRefGoogle ScholarPubMed
Lubchenco, J. (1978). Plant species diversity in a marine intertidal community: importance of herbivore food preference and algal competitive abilities. Am. Nat. 112:23–39.CrossRefGoogle Scholar
Lubchenco, J. (1983). Littorina and Fucus: effects of herbivores, substratum heterogeneity, and plant escapes during succession. Ecology 64:1116–23.CrossRefGoogle Scholar
Lubchenco, J. (1986). Relative importance of competition and predation: early colonization by seaweeds in New England. In Diamond, J. M. and Case, T. (eds). Community Ecology (pp. 537–55). New York: Harper and Row.Google Scholar
Lubchenco, J., and Cubit, J. (1980). Heteromorphic life histories of certain marine algae as adaptations to variations in herbivory. Ecology 61:676–87.CrossRefGoogle Scholar
Lubchenco, J., and Gaines, S. D. (1981). A unified approach to marine plant-herbivore interactions. I. Populations and communities. Annu. Rev. Ecol. Syst. 12:405–37.CrossRefGoogle Scholar
Lubchenco, J., and Menge, B. A. (1978). Community development and persistence in a low rocky intertidal zone. Ecol. Monogr. 48:67–94.CrossRefGoogle Scholar
Lubimenko, V., and Tichovskaya, Q. (1928). Recherches sur la Photosynthese et [‘Adaptation Chromatique chez les Algues Marines. Moscow: Academy of Sciences.Google Scholar
Lück, J., Lück, H. B., L’Hardy-Halos, M.-Th., and Lambert, C. (1999). Simulation of the thallus development of Antithamnion plumula (Ellis) Le Jolis, (Rhodophyceae, Ceramiales). Acta Biotheor. 47:329–51.CrossRefGoogle Scholar
Lüder, U. H., and Clayton, M. N. (2004). Induction of phlorotannins in the brown macroalga Ecklonia radiata (Laminariales, Phaeophyta) in response to simulated herbivory – the first microscopic study. Planta 218:928–37.CrossRefGoogle ScholarPubMed
Lundberg, P., Weich, R. G., Jensen, P., and Vogel, H. J. (1989).Phosphorus-31 and nitrogen-14 studies of the uptake of phosphorus and nitrogen compounds in the marine macroalgaUlva lactuca. Plant Physiol. 89:1380–7.CrossRefGoogle Scholar
Lundheim, R. (1997). Ice nucleation in seaweeds in relation to vertical zonation. J. Phycol. 33: 739–42.CrossRefGoogle Scholar
Lüning, K. (1969). Growth of amputated and dark-exposed individuals of the brown alga Laminaria hyperborea. Mar. Biol. 2: 218–23.CrossRefGoogle Scholar
Lüning, K. (1980). Control of algal life-history by daylength and temperature. In Price, J. H., Irvine, D. E. G., and Farnham, W. F. (eds), The Shore Environments. Vol. 2: Ecosystems (pp. 915–45). New York: Academic Press.Google Scholar
Lüning, K. (l981a). Light. In: Lobban, C. S., and Wynne, M. J. (eds), The Biology of Seaweeds (pp. 326–55). Oxford: Blackwell Scientific.Google Scholar
Lüning, K. (1981b). Photomorphogenesis of reproduction in marine macroalgae. Ber Deutsch. Bot. Ges. 94:401–17.Google Scholar
Lüning, K. (1984). Temperature tolerance and biogeography of seaweeds: the marine algal flora of Helgoland (North Sea) as an example. Helgol. Meeresunters 38: 305–17. [, , ]CrossRefGoogle Scholar
Lüning, K. (1986). New frond formation in Laminaria hyperborea (Phaeophyta): a photoperiodic response. Br. Phycol. J. 21:269–73.[]CrossRefGoogle Scholar
Lüning, K. (1988). Photoperiodic control of sorus formation in the brown alga Laminaria saccharina. Mar. Ecol. Prog. Ser. 45:137–44.CrossRefGoogle Scholar
Lüning, K. (1990). Seaweeds. Their Environment, Biogeography, and Ecophysiology (trans. and ed. Yarish, C. and Kirkman, H.). New York: Wiley-Interscience. [, , , , ]Google Scholar
Lüning, K. (1991). Circannual growth rhythm in a brown alga, Pterygophora californica. Bot. Acta 104:157–62.CrossRefGoogle Scholar
Lüning, K. (1994). When do algae grow? The third Founders’ Lecture. Eur. J. Phycol. 29:61–7. [, , ]CrossRefGoogle Scholar
Lüning, K., and tom Dieck, I. (1989). Environmental triggers in algal seasonality. Bot. Mar. 32:389–97.CrossRefGoogle Scholar
Lüning, K., and tom Dieck, I. (1990). The distribution and evolution of the Laminariales: North Pacific-Atlantic relationships. In Garbary, D. J., and South, G. R. (eds), Evolutionary Biogeography of the Marine Algae of the North Atlantic (pp. 187–204). Berlin: Springer-Verlag.CrossRefGoogle Scholar
Lüning, K., and Dring, M. J. (1972). Reproduction induced by blue light in female gametophytes of Laminaria saccharina. Planta 104:252–6.CrossRefGoogle ScholarPubMed
Lüning, K., and Dring, M. J. (1979). Continuous underwater light measurements near Helgoland (North Sea) and its significance for characteristic light limits in the sublittoral region. Helgol. Meeresunters 32:403–24. [3.0, ]CrossRefGoogle Scholar
Lüning, K., and Freshwater, W. (1988). Temperature tolerance of northeast Pacific marine algae. J. Phycol. 24:310–15.CrossRefGoogle Scholar
Lüning, K., and Neushul, M. (1978). Light and temperature demands for growth and reproduction of laminarian gametophytes in southern and central California. Mar. Biol. 45:297–309.CrossRefGoogle Scholar
Lüning, K., Schmitz, K., and Willenbrink, J. (1973). CO2 fixation and translocation in benthic marine algae. III. Rates and ecological significance of translocation in Laminaria hyperborea and L. saccharina. Mar. Biol. 23:275–81.CrossRefGoogle Scholar
Lüning, K., Titlyanov, E. A., and Titlyanov, T. (1997). Diurnal and circadian periodicity of mitosis and growth in marine macroalgae. III. The red alga Porphyra umbilicalis. Eur. J. Phycol. 32:167–73.CrossRefGoogle Scholar
Lüning, K., Kadel, P., and Pang, S. (2008). Control of reproduction rhythmicity by environmental and endogenous signals in Ulva pseudocurvata (Chlorophyta). J. Phycol. 44:866–73.CrossRefGoogle Scholar
Lyngby, J. E. (1990). Monitoring of nutrient availability and limitation using the marine macroalga Ceramium rubrum (Huds.) C. Ag. Aquat. Bot. 38:153–61.CrossRefGoogle Scholar
Lyons, D. A., and Scheibling, R. E. (2009). Range expansion by invasive marine algae: rates and patterns of spread at a regional scale. Diversity Distrib. 15:762–75.CrossRefGoogle Scholar
Maberly, S. C. (1990). Exogenous sources of inorganic carbon for photosynthesis by marine macroalgae. J. Phycol. 26:439–49.CrossRefGoogle Scholar
Maberly, S. C., and Madsen, T. V. (1990). Contribution of air and water to the carbon balance of Fucus spiralis. Mar. Ecol. Prog. Ser.62:175–83.CrossRefGoogle Scholar
Maberly, S. C., Raven, J. A., and Johnston, A. M. (1992). Discrimination of 12C and 13C by marine plants. Oecologia 91:481–92.CrossRefGoogle Scholar
MacArtain, P., Gill, C. I. R., Brooks, M., Campbell, R., and Rowland, I. R. (2007). Nutritional value of edible seaweeds. Nutr. Rev. 65:535–43.CrossRefGoogle ScholarPubMed
MacArthur, R. H. (1955). Fluctuations of animal populations and a measure of community stability. Ecology 36:533–6.CrossRefGoogle Scholar
MacArthur, R. H. (1965). Patterns of species diversity. Biol. Rev. 40:510–33.CrossRefGoogle Scholar
MacArthur, R. H. (1972). Strong, or weak, interactions? Transactions of the Connecticut Academy of Arts and Sciences 44:177–88.Google Scholar
MacArthur, R. H., and Levins, R. (1967). The limiting similarity, convergence, and divergence of coexisting species. Am. Nat. 101:377–85.CrossRefGoogle Scholar
MacArthur, R. H., and Wilson, E. O. (1963). An equilibrium theory of insular zoogeography. Evolution 17:373–87.CrossRefGoogle Scholar
Macaya, E. C., and Zuccarello, G. C. (2010). Genetic structure of the giant kelp Macrocystis pyrifera along the southeastern Pacific. Mar. Ecol. Prog. Ser. 420:103–12.CrossRefGoogle Scholar
Mach, K. J. (2009). Mechanical and biological consequences of repetitive loading: crack initiation and fatigue failure in the red macroalga Mazzaella. J. Exp. Biol. 212:961–76.CrossRefGoogle ScholarPubMed
Mach, K. J., Nelson, D. V., and Denny, M. W. (2007). Techniques for predicting the lifetimes of wave-swept macroalgae: a primer on fracture mechanics and crack growth. J. Exp. Biol. 210:2213–30.CrossRefGoogle ScholarPubMed
Machalek, K. M., Davison, I. R., and Falkowski, P. G. (1996). Thermal acclimation and photoacclimation of photosynthesis in the brown alga Laminaria saccarina. Plant Cell. Envir. 19: 1005–16.CrossRefGoogle Scholar
Mackerness, S. A.-H., and Jordan, B. R. (1999). Changes in gene expression in response to UV-B induced stress. In Pessarakli, M. (ed.), Handbook of Plant and Crop Stress (pp. 749–68). New York: Marcel Dekker Inc.Google Scholar
Mackie, W., and Preston, R. D. (1974). Cell wall and intercellular region polysaccharides. In Stewart, W. D. P (ed.), Algal Physiology and Biochemistry (pp. 40–85). Oxford: Blackwell Scientific.Google Scholar
Madronich, S., McKenzie, R. L., Björn, L. O., and Caldwell, M. M. (1998). Changes in biologically active ultraviolet radiation reaching the earth´s surface. J. Photochem. Photobiol. B: Biol. 46:5–19.CrossRefGoogle ScholarPubMed
Madsen, T. V., and Maberly, S. C. (1990). A comparison of air and water as environments for photosynthesis by the intertidal alga Fucus spiralis L. J. Phycol. 26:24–30.CrossRefGoogle Scholar
Maggs, C. A. (1988). Intraspecific life history variability in the Florideophycidae (Rhodophyta). Bot. Mar. 31:465–90.CrossRefGoogle Scholar
Maggs, C. A. (1989). Erythrodermis allenii Batters in the life history of Phyllophora traillii Holmes ex Batters (Phyllophoraceae, Rhodophyta). Phycologia 28:305–17.CrossRefGoogle Scholar
Maggs, C. A. (1998). Life history variation in Dasya ocellata (Dasyaceae, Rhodophyta). Phycologia 37:100–5.CrossRefGoogle Scholar
Maggs, C. A., and Cheney, D. P. (1990). Competition studies of marine macroalgae in laboratory culture. J. Phycol. 26:17–24. [4.0]CrossRefGoogle Scholar
Maggs, C. A., and Guiry, M. D. (1987). Environmental control of macroalgal phenology. In Crawford, R. M. M. (ed.), Plant Life in Aquatic and Amphibious Habitats (pp. 359–73). Oxford: Blackwell Scientific.Google Scholar
Maggs, C. A., and Pueschel, C. M. (1989). Morphology and development of Ahnfeltia plicata (Rhodophyta): proposal of Ahnfeltiales ord. novoJ. Phycol. 25:333–51.CrossRefGoogle Scholar
Maggs, C. A., Fletcher, H. L., and Fewer, D. (2011). Speciation in red algae: members of the Ceramiales as model organisms. Integr. Comp. Biol. 51:492–504.CrossRefGoogle ScholarPubMed
Magruder, W. H. (1984). Specialized appendages on spermatia from the red alga Aglaothamnion neglectum (Ceramiales, Ceramiaceae) specifically bind with trichogynes. J. Phycol. 20:436–40.CrossRefGoogle Scholar
Maier, C. M., and Pregnall, A. M. (1990). Increased macrophyte nitrate reductase activity as a consequence of groundwater input of nitrate through sandy beaches. Mar. Biol. 107:263–71.CrossRefGoogle Scholar
Maier, I. (1997). Fertilization, early embryogenesis and parthenogenesis in Durvillaea potatorum (Durvillaeales, Phaeophyceae). Nova Hedwigia 64:41–50.Google Scholar
Maier, I., and Müller, D. G. (1986). Sexual pheromones in algae. Biol. Bull. 170:145–75.CrossRefGoogle Scholar
Maier, I., Hertweck, C., and Boland, W. (2001). Stereochemical specificity of lamoxirene, the sperm-releasing pheromone in kelp (Laminariales, Phaeophyceae). Biol. Bull. 201:121–5.CrossRefGoogle Scholar
Makarov, V. N., Schoschina, E. V., and Lüning, K. (1995). Diurnal and circadian periodicity of mitosis and growth in marine macroalgae. I. Juvenile sporophytes of Laminariales (Phaeophyta). Eur. J. Phycol. 30:261–6.CrossRefGoogle Scholar
Malta, E. J., Ferreira, D. G., Vergara, J. J., and Perez-Llorens, J. L. (2005). Nitrogen load and irradiance affect morphology, photosynthesis and growth of Caulerpa prolifera (Bryopsidales: Chlorophyta). Mar. Ecol. Prog. Ser. 298:101–14.CrossRefGoogle Scholar
Mance, G. (1987). Pollution Threat of Heavy Metals in Aquatic Environments. Amsterdam: Elsevier.CrossRefGoogle Scholar
Mandal, P., Mateu, C. G., Chattopadhay, K., et al. (2007). Structural features and antiviral activity of sulphated fucans from the brown seaweed Cystoseira indica. Antivir. Chem. Chemother. 18:153–62.CrossRefGoogle ScholarPubMed
Mandoli, D. F. (1998a). What ever happened to Acetabularia? Bringing a once-classic model system into the age of molecular genetics. Int. Rev. Cytol. 182:1–67. [, , ]CrossRefGoogle Scholar
Mandoli, D. F. (1998b). Elaboration of body plan and phase change during development of Acetabularia: how is the complex architecture of a giant unicell built? Annu. Rev. Plant Physiol. Plant Mol. Biol. 49:173–98.CrossRefGoogle Scholar
Maneveldt, G. W., and Keats, D. W. (2008). Effects of herbivore grazing on the physiognomy of the coralline alga Spongites yendoi and on associated competitive interactions. Afr. J. Mar. Sci. 30:581–93.CrossRefGoogle Scholar
Maneveldt, G. W., Eager, R. C., and Bassier, A. (2009). Effects of long-term exclusion of the limpet Cymbula oculus (Born) on the distribution of intertidal organisms on a rocky shore. Afr. J. Mar. Sci. 31:171–9.CrossRefGoogle Scholar
Manley, S. L. (1983). Composition of sieve tube sap from Macrocystis pyrifera (Phaeophyta) with emphasis on the inorganic constituents. J. Phycol. 19:118–21.CrossRefGoogle Scholar
Manley, S. L., and North, W.J. (1984). Phosphorus and the growth of juvenile Macrocystis pyrifera (Phaeophyta) sporophytes. J. Phycol. 20:389–93.CrossRefGoogle Scholar
Mann, K. H. (1972). Ecological energetics of the seaweed zone in a marine bay on the Atlantic coast of Canada. II. Productivity of the seaweeds. Mar. Biol. 14:199–209.Google Scholar
Mann, K. H., Jarman, N., and Dieckmann, G. (1979). Development of a method for measuring the productivity of the kelp Ecklonia maxima (Osbeck) Papenf. Trans. R. Soc. S. Afr. 44:27–41.CrossRefGoogle Scholar
Manney, G. L. , Santee, M. L., Rex, M., et al. (2011). Unprecedented Arctic ozone loss in 2011. Nature 478:469–75.CrossRefGoogle ScholarPubMed
Manning, W.M., and Strain, H. H. (1943). Chlorophyll d, a green pigment of red algae. J. Biol. Chem. 151:1–19.Google Scholar
Mantyka, C. S., and Belllwood, D. R. (2007). Macroalgal grazing selectivity among herbivorous coral reef fishes. Mar. Ecol. Prog. Ser. 352:177–85.CrossRefGoogle Scholar
Marande, W., and Kohl, L. (2011). Flagellar kinesins in protests. Future Microbiol. 6:231–46.CrossRefGoogle Scholar
Marian, F. D., Garcia-Jiménez, P., and Robaina, R. R. (2000). Polyamines in marine macroalgae: levels of putrescine, spermidine and spermine in the thalli and changes in their concentration during glycerol-induced cell growth in vitro. Physiol. Plantarum 110:530–4.CrossRefGoogle Scholar
Markager, S., and Sand-Jensen, K. (1990). Heterotrophic growth of Ulva lactuca (Chlorophyceae). J. Phycol. 26:670–3.CrossRefGoogle Scholar
Markham, J. W. (1973). Observations on the ecology of Laminaria sinclairii on three northern Oregon beaches. J. Phycol. 9:336–41.Google Scholar
Markham, J. W., and Newroth, P. R. (1972). Observations on the ecology of Gymnogongrus linearis and related species. Proc. Intl. Seaweed Symp. 7:127–30.Google Scholar
Markham, J. W., Kremer, B. P., and Sperling, K. R. (1980). Effect of cadmium on Laminaria saccharina in culture. Mar. Ecol. Prog. Ser. 3:31–9.CrossRefGoogle Scholar
Marquardt, R., Schubert, H., Varela, D. A. et al. (2010). Light acclimation strategies of three commercially important red algal species. Aquaculture 299:140–8.CrossRefGoogle Scholar
Marsden, A. D., and DeWreede, R. E. (2000). Marine macroalgal community structure, metal content and reproductive function near an acid mine drainage outflow. Environ. Poll. 110:431–40.CrossRefGoogle ScholarPubMed
Marshall, K., Joint, I., Callow, M. E., and Callow, J. A. (2006). Effect of marine bacterial isolates on the growth and morphology of axenic plantlets of the green algaUlva linza. Microbial Ecol. 52:302–10.CrossRefGoogle ScholarPubMed
Martinez, B., and Rico, J. M. (2004). Inorganic nitrogen and phosphorus uptake kinetics in Palmaria palmata (Rhodophyta). J. Phycol. 40:642–50.CrossRefGoogle Scholar
Martinez, E. A., Destombe, C., Quillet, M. C., and Valero, M. (1999). Identification of random amplified polymorphic DNA (RAPD) markers highly linked to sex determination in the red alga Gracilaria gracilis. Mol. Ecol. 8:1533–8.CrossRefGoogle ScholarPubMed
Martins, G. M., Hawkins, S. J., Thompson, R. C., and Jenkins, S. R. (2007). Community structure and functioning in intertidal rock pools: effects of pool size and shore height at different successional stages. Mar. Ecol. Prog. Ser. 329:43–55.CrossRefGoogle Scholar
Martone, P. T. (2006). Size, strength and allometry of joints in the articulated coralline Calliarthron. J. Exp. Biol. 209:1678–89.CrossRefGoogle ScholarPubMed
Martone, P. T. (2007). Kelp versus coralline: cellular basis for mechanical strength in the wave-swept seaweed Calliarthron (Corallinaceae, Rhodophyta). J. Phycol. 43:882–91.CrossRefGoogle Scholar
Martone, P. T., and Denny, M. W. (2008a). To break a coralline: mechanical constraints on the size and survival of a wave-swept seaweed. J. Exp. Biol. 211: 3433–41.CrossRefGoogle ScholarPubMed
Martone, P. T., and Denny, M. W. (2008b). To bend a coralline: effect of joint morphology on flexibility and stress amplification in an articulated calcified seaweed. J. Exp. Biol. 211:3421–32.CrossRefGoogle Scholar
Martone, P. T., Estevez, J. M., Lu, F., et al. (2009). Discovery of lignin in seaweed reveals convergent evolution of cell wall architecture. Curr. Biol. 19:169–75.CrossRefGoogle ScholarPubMed
Martone, P. T., Boller, M., Burgert, I., et al. (2010a). Mechanics without muscle: biomechanical inspiration from the plant world. Integr. Comp. Biol. 50:888–907.CrossRefGoogle ScholarPubMed
Martone, P. T., Navarro, D. A., Stortz, C. A., and Estevez, J. M. (2010b). Differences in polysaccharide structure between calcified and uncalcified segments in the coralline Calliarthron cheilosporiodes. J. Phycol. 46:507–15.CrossRefGoogle Scholar
Martone, P. T., Kost, L., and Boller, M. (2012). Drag reduction in wave-swept macroalgae: alternative strategies and new predictions. Am. J. Bot. 99:806–15.CrossRefGoogle ScholarPubMed
Maschek, J. A., and Baker, B. J. (2008). The chemistry of algal secondary metabolites. In Amsler, C. D. (ed.), Algal Chemical Ecology (pp. 1–24). London: Springer (Limited).Google Scholar
Mass, T., Genin, A., Shavit, U., Grinstein, M., and Tchernov, D. (2010). Flow enhances photosynthesis in marine benthic autotrophs by increasing the efflux of oxygen from the organism to the water. P. Natl. Acad. Sci. USA 107:2527–31.CrossRefGoogle Scholar
Masterson, P., Arenas, F. A., Thompson, R. C., and Jenkins, S. R. (2008). Interaction of top down and bottom up factors in intertidal rockpools: effects on early successional macroalgal community composition, abundance and productivity. J. Exp. Mar. Biol. Ecol. 363:12–20.CrossRefGoogle Scholar
Mathews, S. (2006). Phytochrome-mediated development in land plants: red light sensing evolves to meet the challenges of changing light environments. Mol. Ecol. 15:3483–503.CrossRefGoogle ScholarPubMed
Mathieson, A. C., and Norall, T. L. (l975). Physiological studies of subtidal red algae. J. Exp. Mar. Biol. Ecol. 20:237–47.CrossRefGoogle Scholar
Mathieson, A. C., and Penniman, C. A. (1991). Floristic patterns and numerical classification of New England estuarine and open coastal seaweed populations. Nova Hedwigia 52:453–85.Google Scholar
Mathieson, A. C., Dawes, C. J., Anderson, M. L., and Hehre, E. J. (2001). Seaweeds of the Brave Boat Harbor salt marsh and adjacent open coast of southern Maine. Rhodora 103:1–46.Google Scholar
Mathieson, A. C., Dawes, C. J., Wallace, A. L., and Klein, A. S. (2006). Distribution, morphology, and genetic affinities of dwarf embedded Fucus populations from the Northwest Atlantic Ocean. Bot. Mar. 49:283–303.CrossRefGoogle Scholar
Mathieson, A. C., Hehre, E. J., Dawes, C. J., and Neefus, C. D. (2008). An historical comparison of seaweed populations from Casco Bay, Maine. Rhodora 110:1–102.CrossRefGoogle Scholar
Matilsky, M. B., and Jacobs, W. (1983). Accumulation of amyloplasts on the bottom of normal and inverted rhizome tips of Caulerpa prolifera (Forsskal) Lamouroux. Planta 159:189–92.CrossRefGoogle ScholarPubMed
Matson, P. G., and Edwards, M. S. (2007). Effects of ocean temperature on the southern range limits of two understory kelps, Pterygophora californica and Eisenia arborea, at multiple life-stages. Mar. Biol. 151:1941–9.CrossRefGoogle Scholar
Matsunaga, S., Uchida, H., Iseki, M., Watanabe, M., and Murakami, A. (2010). Flagellar motions in phototactic steering in a brown algal swarmer. Photochem. Photobiol. 86:374–81.CrossRefGoogle Scholar
Matsuo, Y., Imagawa, H., Nishizawa, M., and Shizuri, Y. (2005). Isolation of an algal morphogenesis inducer from a marine bacterium. Science 307:1598.CrossRefGoogle ScholarPubMed
Matz, C. (2011). Competition, communication, cooperation: molecular crosstalk in multi-species biofilms. In Flemming, H.-C., Wingender, J., and Szewzyk, U. (eds), Biofilm Highlights (pp. 29–40). Springer Series on Biofilms 5. Berlin and Heidelberg: Springer-Verlag.CrossRefGoogle Scholar
Maumus, F., Rabinowicz, P., Bowler, C., and Rivarola, M. (2011). Stemming epigenetics in marine stramenopiles. Curr. Genomics 12:357–70.CrossRefGoogle ScholarPubMed
Maxell, B. A., and Miller, K. A. (1996). Demographic studies of the annual kelps Nereocystis luetkeana and Costaria costata (Laminariales, Phaeophyta) in Puget Sound, Washington. Bot. Mar. 39:479–89.CrossRefGoogle Scholar
Maximilien, R., de Nys, R., Holmström, C., et al. (1998). Chemical mediation of bacterial surface colonization by secondary metabolites from the red alga Delisea pulchra. Aquat. Microb. Ecol. 15:233–46.CrossRefGoogle Scholar
Maximova, O. V., and Sazhin, A. F. (2010). The role of gametes of the macroalgae Ascophyllum nodosum (L.) Le Jolis and Fucus vesiculosus L. (Fucales, Phaeophyceae) in summer nanoplankton of the White Sea coastal waters. Mar. Biol. 50:218–29.Google Scholar
Mayhoub, H., Gayral, P., and Jacques, R. (1976). Action de la composition spectrale de la lumière sur la croissance et la reproduction de Calosiphonia vermicularis (J. Agardh) Schmitz (Rhodophycées Gigartinales). C. R. Acad. Sci. Paris 283(D):1041–4.Google Scholar
Mayr, E. (1982). The Growth of Biological Thought: Diversity, Evolution, and Inheritance. Cambridge, MA: Belknap/Harvard Press, 992 pp.Google Scholar
McArthur, D. M., and Moss, B. L. (1977). The ultrastructure of cell walls in Enteromorpha intestinalis (L.) Link. Br. Phycol. J. 12:359–68.CrossRefGoogle Scholar
McCandless, E. L. (1981). Polysaccharides of the seaweeds. In Lobban, C. S. and Wynne, M. J. (eds) The Biology of Seaweeds (pp. 559–88). Oxford: Blackwell Scientific.Google Scholar
McCandless, E. L. J., and Craigie, J. S. (1979). Sulfated polysaccharides in red and brown algae. Annu. Rev. Plant Physiol. 30:41–53.CrossRefGoogle Scholar
McClanahan, T. R., Sala, E., Stickels, P., et al. (2003). Interaction between nutrients and herbivory in controlling algal communities and coral condition on Glover’s Reef, Belize. Mar. Ecol Prog. Ser. 261:135–47.CrossRefGoogle Scholar
McClanahan, T. R., Sala, E., Mumby, P. J., and Jones, S. (2004). Phosphorus and nitrogen enrichment do not enhance brown frondose “macroalgae”. Mar. Poll. Bull. 48:196–9.CrossRefGoogle Scholar
McClanahan, T. R., Steneck, R. S., Pietri, D., Cokos, B., and Jones, S. (2005). Interaction between inorganic nutrients and organic matter in controlling coral reef communities in Glovers Reef Belize. Mar. Poll. Bull. 50:566–75.CrossRefGoogle ScholarPubMed
McClintock, J. B., and Baker, B. J. (2010). Marine Chemical Ecology. Taylor and Francis, 624 pp.
McClintock, M., Higinbotham, N., Uribe, E. G., and Cleland, R. E. (1982). Active, irreversible accumulation of extreme levels of H2SO4 in the brown alga, Desmarestia. Plant Physiol. 70:771–4.CrossRefGoogle Scholar
McClintock, J. B., Amsler, C. D., and Baker, B. J. (2010). Overview of the chemical ecology of benthic marine invertebrates along the western Antarctic Peninsula. Integr. Compar. Biol. (Advanced Access) 50:967–80.CrossRefGoogle ScholarPubMed
McConnaughey, T. (1991). Calcification in Chara corallina: CO2 hydroxylation generates protons for bicarbonate assimilation. Limnol. Oceanogr. 36:619–28.CrossRefGoogle Scholar
McConnaughey, T. (1998). Acid secretion, calcification, and photosynthetic carbon concentrating mechanisms. Can. J. Bot. 76:1119–26.Google Scholar
McConnaughey, T., and Whelan, J. F. (1997). Calcification generates protons for nutrient and bicarbonate uptake. Earth Sci. Rev. 42:95–117.CrossRefGoogle Scholar
McConnico, L. A., and Foster, M. S. (2005). Population biology of the intertidal kelp, Alaria marginata Postels and Ruprecht: a non-fugitive annual. J. Exp. Mar. Biol. Ecol. 324:61–75.CrossRefGoogle Scholar
McCook, L. J., and Chapman, A. R. O. (1992). Vegetative regeneration of Fucus rockweed canopy as a mechanism of secondary succession on an exposed rocky shore. Bot. Mar. 35:35–46.CrossRefGoogle Scholar
McCook, L. J., Jompa, J., and Diaz-Pulido, G. (2001). Competition between corals and algae on coral reefs: a review of evidence and mechanisms. Coral Reefs 19:400–17.CrossRefGoogle Scholar
McCord, J. M., and Fridovich, I. (1969). Superoxide dismutase—an enzymatic function for erythrocuprein (hemocuprein). J. Biol. Chem. 22:6049–55.Google Scholar
McCracken, D. A., and Cain, J. R. (1981). Amylose in floridean starch. New Phytol. 88:67–71.CrossRefGoogle Scholar
McDevit, D.C., and Saunders, G. W. (2009). On the utility of DNA barcoding for species differentiation among brown macroalgae (Phaeophyceae) including a novel extraction protocol. Phycol. Res. 57:131–41.CrossRefGoogle Scholar
McDevit, D.C., and Saunders, G. W. (2010). A DNA barcode examination of the Laminariaceae (Phaeophyceae) in Canada reveals novel biogeographical and evolutionary insights. Phycologia 49: 235–48.CrossRefGoogle Scholar
McGlathery, K. J., Marino, R., and Howarth, R. W. (1994). Variable rates of phosphorus uptake by shallow marine carbonate sediments: Mechanisms and ecological significance. Biogeochem. 25:127–46.CrossRefGoogle Scholar
McGlathery, K. J., Pedersen, M. F., and Borum, J. (1996). Changes in intracellular nitrogen pools and feedback controls on nitrogen uptake in Chaetomorpha linum (Chlorophyta). J. Phycol. 32:393–401.CrossRefGoogle Scholar
McGlathery, K. J., Sundback, K., and Anderson, I. C. (2007). Eutrophication in shallow coastal bays and lagoons: the role of plants in the coastal filter. Mar. Ecol. Prog. Ser. 348:1–18.CrossRefGoogle Scholar
McHugh, D. J. (2004). A Guide to the Sewaweed Industry. FAO Fosheries Technical Paper. No. 441, Rome: FAO, 105 pp. [, , , , , , , , , , ]Google Scholar
McKay, R. M. L., and Gibbs, S. P. (1990). Phycoerythrin is absent from the pyrenoid of Porphyridium cruentum: photosynthetic implications. Planta 180:249–56.CrossRefGoogle ScholarPubMed
McKenzie, G. H., Ch’ng, A. L., and Gayler, K. R. (1979). Glutamine synthetase/glutamine:α-ketoglutarate aminotransferase in chloroplasts from the marine alga Caulerpa simpliciuscula. Plant Physiol. 63:578–82.CrossRefGoogle ScholarPubMed
McKenzie, P. F., and Bellgrove, A. (2009). Dislodgement and attachment strength of the intertidal macroalga Hormosira banksii (Fucales, Phaeophyceae). Phycologia 48:335–43.CrossRefGoogle Scholar
McLachlan, P. J., and Bidwell, R. G. S. (1978). Photosynthesis of eggs, sperm, zygotes, and embryos ofFucus serratus. Can. J. Bot. 56:371–3.CrossRefGoogle Scholar
Meinesz, A. (1980). Connaissances actuelles et contribution a l'etude de la reproduction et du cycle des Udoteacees (Caulerpales,Chlorophytes). Phycologia 19:110–38.CrossRefGoogle Scholar
Meinesz, A. (2007). Methods for identifying and tracking seaweed invasions. Bot. Mar. 50:373–84.CrossRefGoogle Scholar
Mejia, A. Y., Puncher, G. N., and Engelen, A. H. (2012). Macroalgae in tropical marine coastal systems. In Wiencke, C. and Bischof, K. (eds), Seaweed Biology: Novel Insights Into Ecophysiology, Ecology and Utilization (Series: Ecological Studies, Volume 219, pp. 329–57). Berlin and Heidelberg: Springer.CrossRefGoogle Scholar
Melkonian, M., and Robenek, H. (1984). The eyespot apparatus of flagellated green algae. Prog. Phycol. Res. 3:193–268.Google Scholar
Meneses, I., and Santelices, B. (1999). Strain selection and genetic variation in Gracilaria chiliensis (Gracilariales, Rhodophyta). J. Appl. Phycol. 11:241–6.CrossRefGoogle Scholar
Menge, B. A. (1972a). Competition for food between two intertidal starfish species and its effect on body size and feeding. Ecology 53:635–44.CrossRefGoogle Scholar
Menge, B. A. (1972b). Foraging strategy of a starfish in relation to actual prey availability and environmental predictability. Ecol. Monogr. 42:25–50.CrossRefGoogle Scholar
Menge, B. A. (1976). Organization of the New England rocky intertidal community: role of predation, competition and environmental heterogeneity. Ecol. Monogr. 46:355–93.CrossRefGoogle Scholar
Menge, B. A. (1978a). Predation intensity in a rocky intertidal community. Effect of an algal canopy, wave action and desiccation on predator feeding rates. OecologiaBerlin 34:17–35.Google Scholar
Menge, B. A. (1978b). Predation intensity in a rocky intertidal community. Relation between predator foraging activity and environmental harshness. Oecologia Berlin 34:1–16.CrossRefGoogle Scholar
Menge, B. A. (1983). Components of predation intensity in the low zone of the New England rocky intertidal region. Oecologia 58:141–55.CrossRefGoogle ScholarPubMed
Menge, B. A. (1991). Generalizing from experiments: is predation strong or weak in the New England rocky intertidal? Oecologia 88:1–8.CrossRefGoogle ScholarPubMed
Menge, B. A. (1992). Community regulation: under what conditions are bottom-up factors important on rocky shores? Ecology 73:755–65.CrossRefGoogle Scholar
Menge, B. A., and Branch, G. M. (2001). Rocky intertidal communities. In Bertness, M. D., Gaines, S. D., and Hay, M. E. (eds), Marine Community Ecology (pp. 221–51). Sunderland, MA: Sinauer Associates.Google Scholar
Menge, B. A., and Sutherland, J. P. (1976). Species diversity gradients: synthesis of the roles of predation, competition, and temporal heterogeneity. Am. Nat. 110:351–69.CrossRefGoogle Scholar
Menge, B. A., and Sutherland, J. P. (1987). Community regulation: variation in disturbance, competition, and predation in relation to environmental stress and recruitment. Am. Nat. 130:730–57.CrossRefGoogle Scholar
Menge, B. A., Lubchenco, J., Gaines, S. D., and Ashkenas, L. R. (1986). A test of the Menge-Sutherland model of community organization in a tropical rocky intertidal food web. Oecologia (Berlin) 71:75–89.CrossRefGoogle Scholar
Menge, B. A., Berlow, E. L., Blanchette, C. A., Navarrete, S. A., and Yamada, S. B. (1994). The keystone species concept: variation in interaction strength in a rocky intertidal habitat. Ecol. Monogr. 64:249–86.CrossRefGoogle Scholar
Menge, B. A., Lubchenco, J., Bracken, M. E. S, et al. (2003). Coastal oceanography sets the pace of rocky intertidal community dynamics. Proc. Nat. Acad. Science USA 100: 12,229–34.CrossRefGoogle ScholarPubMed
Menge, J. L., and Menge, B. A. (1974). Role of resource allocation, aggression and spatial heterogeneity in coexistence of two competing intertidal starfish. Ecol. Monogr. 44:189–209.CrossRefGoogle Scholar
Menzel, D. (1988). How do giant plant cells cope with injury? The wound response in siphonous green algae. Protoplasma 144:73–91.CrossRefGoogle Scholar
Menzel, D. (1994). Cell differentiation and the cytoskeleton in Acetabularia. New Phytol. 128:369–93.CrossRefGoogle Scholar
Menzel, D., and Elsner-Menzel, C. (l989). Actin-based chloroplast rearrangements in the cortex of the giant coenocytic green alga Caulerpa. Protoplasma 150: 1–8.CrossRefGoogle Scholar
Mercado, J., Jimenez, C., Niell, F. X., and Figueroa, F. L. (1996). Comparison of methods for measuring light absorption by algae and their application to the estimation of the package effect. Sci. Mar. 60:39–45.Google Scholar
Mercado, J. M., Carmona, R., and Niell, F. X. (1998). Bryozoans increase available CO2 for photosynthesis in Gelidium sesquipedale (Rhodophyceae). J. Phycol. 34:925–7.CrossRefGoogle Scholar
Mercado, J. M., de los Santos, C. B., Perez-Llorens, J. L., and Vergara, J. J. (2009). Carbon isotopic fractionation in macroalgae from Cadiz Bay (Southern Spain): comparision with other bio-geographic regions. Estuar. Coast Shelf Sci. 85:449–58.CrossRefGoogle Scholar
Metaxas, A., and Scheibling, R. E. (1994). Spatial and temporal variability of tidepool hyperbenthos on a rocky shore in Nova Scotia, Canada. Mar. Ecol. Prog. Ser. 108:175–84.CrossRefGoogle Scholar
Michael, T. S. (2009). Glycoconjugate organization of Enteromorpha (=Ulva) flexuosa and Ulva fasciata (Chlorophyta) zoospores. J. Phycol. 45:660–77.CrossRefGoogle ScholarPubMed
Michel, G., Tonon, T., Scorner, D., Cock, J. M., and Kloareg, G. (2010a). The cell wall polysaccharide metabolism of the brown alga Ectocarpus siliculosus. Insights into the evolution of extracellular matrix polysaccharides in eukaryotes. New Phytol. 188:82–97. [, , ]CrossRefGoogle ScholarPubMed
Michel, G., Tonon, T., Scornet, D., Cock, J. M., and Kloareg, B. (2010b). Central and storage carbon metabolism of the brown alga Ectocarpus siliculosus: insights into the origin and evolution of storage carbohydrates in eukaryotes. New Phytol. 188:67–81.CrossRefGoogle ScholarPubMed
Michener, R. H., and Schell, D. M. (1994). Stable isotope ratios as tracers in marine aquatic food webs. In Lajtha, K., and Michener, R. H. (eds), Stable Isotopes in Ecology and Environmental Science (pp. 138–57). Oxford: Blackwell Scientific Publications.Google Scholar
MillerIII, H. L., and Dunton, K. H. (2007). Stable isotope (13C) and O2 micro-optode alternatives for measuring photosynthesis in seaweeds. Mar. Ecol. Prog. Ser. 329:85–97.CrossRefGoogle Scholar
MillerIII, H. L., Neale, P. J., and Dunton, K. H. (2009). Biological weighting functions for UV inhibition of photosynthesis in the kelp Laminaria hyperborea (Phaeophyceae). J. Phycol. 45:571–84.CrossRefGoogle Scholar
Miller, K. A., Olsen, J. L., and Stam, W. T. (2000). Genetic divergence correlates with morphological and ecological subdivision in the deep-water elk kelp, Pelagophycus porra (Phaeophyceae). J. Phycol. 36:862–70.CrossRefGoogle Scholar
Miller, R. J., Lenihan, H. S., Muller, E. B., et al. (2010). Impacts of metal oxide nanoparticles on marine phytoplankton. Environ. Sci. Technol. 44:7329–34.CrossRefGoogle ScholarPubMed
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): .Google ScholarPubMed
Miller, S. M., Wing, S. R., and Hurd, C. L. (2006). Photoacclimation of Ecklonia radiata (Laminariales, Heterokontophyta) in Doubtful Sound, Fjordland, Southern New Zealand. Phycologia 45:44–52.CrossRefGoogle Scholar
Milligan, A. J., and Harrison, P. J. (2000). Effects of non-steady-state iron limitation on nitrogen assimilatory enzymes in the marine diatom Thalassiosira weissfloggi (Bacillariophyta). J. Phycol. 36:78–86.CrossRefGoogle Scholar
Milligan, K. L. D., and De Wreede, R. E. (2000). Variations in holdfast attachment mechanics with developmental stage, substratum-type, season, and wave-exposure for the intertidal kelp species Hedophyllum sessile (C. Agardh) Setchell. J. Exp. Mar. Biol. Ecol. 254:189–209.CrossRefGoogle ScholarPubMed
Millner, A., and Evans, L. V. (1980). The effects of triphenyltin chloride on respiration and phytosynthesis in the green algae Enteromorpha intestinalis and Ulothrix pseudoflacca. Plant Cell Environ. 3:339–48.CrossRefGoogle Scholar
Millner, A., and Evans, L. V. (1981). Uptake of triphenyltin chloride by Enteromorpha intestinalis and Ulothrix pseudoflacca. Plant Cell Environ. 4:383–9.CrossRefGoogle Scholar
Mimura, T., Reid, R. J., and Smith, F. A. (1998). Control of phosphate transport across the plasmalemma of Chara corallina. J. Exp. Bot. 49:13–19.CrossRefGoogle Scholar
Mimuro, M., and Akimoto, S. (2003). Carotenoids of light harvesting systems: energy transfer processes from fucoxanthin and peridinin to chlorophyll. In Larkum, A. W. D., Douglas, S. E., and Raven, J. A. (eds) Photosynthesis in Algae. Advances in Photosynthesis and Respiration, (Vol. 14 pp. 335–49). Dordrecht, The Netherlands: Kluwer Academic Publishers.CrossRefGoogle Scholar
Minagawa, M., and Wada, E. (1984). Stepwise enrichment of 15N along food chains: further evidence and the relationship between σ 15N and animal age. Geochim. Cosmochim. Acta 48:1135–40.[]CrossRefGoogle Scholar
Mine, I., Anota, Y., Menzel, D., and Okuda, K. (2005). Poly (A)+RNA and cytoskeleton during cyst formation in the cap ray of Acetabularia peniculus. Protoplasma 226:199–206.CrossRefGoogle ScholarPubMed
Mine, I., Menzel, D., and Okuda, K. (2008). Morphogenesis in giant-celled algae. Int. Rev. Cell Mol. Biol. 266:37–83.CrossRefGoogle ScholarPubMed
Miner, B. G., Sultan, S. E., Morgan, S. G., Padilla, D. K., and Relyea, R. A. (2005). Ecological consequences of phenotypic plasticity. Trends Ecol. Evol. 20:685–92.CrossRefGoogle ScholarPubMed
Mishkind, M., Mauzerall, D., and Beale, S. I. (1979). Diurnal variation in situ of photosynthetic capacity in Ulva caused by a dark reaction. Plant Physiol. 64: 896–9.CrossRefGoogle ScholarPubMed
Mitman, G. G., and van der Meer, J. P. (1994). Meiosis, blade development, and sex determination in Porphyra purpurea (Rhodophyta). J. Phycol. 30:147–59.CrossRefGoogle Scholar
Miura, A. (1975). Porphyra cultivation in Japan. In Tokida, J. and Hirose, H. (eds), Advance of Phycology in Japan (pp. 273–304). The Hague: Dr. W. Junk.Google Scholar
Miyagishima, S.-Y., and Nakanishi, H. (2010). The chloroplast division machinery: origin and evolution. In Seckbach, J. and Chapman, D. J. (eds), Red Algae in the Genomic Age (Cellular Origin, Life in Extreme Habitats and Astrobiology 13) (pp. 3–23). Dordrecht, The Netherlands: Springer.Google Scholar
Miyamura, S. (2010). Cytoplasmic inheritance in green algae: patterns, mechanisms and relation to sex type. J. Plant Res. 123:171–84.CrossRefGoogle ScholarPubMed
Miyamura, S., Sakaushi, S., Hori, T., and Nagumo, T. (2010). Behavior of flagella and flagellar root systems in the planozygotes and settled zygotes of the green alga Bryopsis maxima Okamura (Ulvophyceae, Chlorophyta) with reference to spatial arrangement of eyespot and cell fusion site. Phycol. Res. 58:258–69.CrossRefGoogle Scholar
Miyashita, H., Ikemoto, H., Kurano, N., and Miyachi, S. (2003). Acaryochloris marina ge. Et. Sp. nov (Cyabobacteria), an oxygenic photosynthetic prokaryote containing chlorophyll d as a major pigment. J. Phycol. 39:1247–53.CrossRefGoogle Scholar
Mizuno, M. (1984). Environment at the front shore of the Institute of Algological Research of Hokkaido University. Sci. Pap. Inst. Algol. Res., Fac. Sci. Hokkaido U 7:263–92.Google Scholar
Mizuta, H., Kai, T., Tabuchi, K., and Yasui, H. (2007). Effects of light quality on the reproduction and morphology of sporophytes of Laminaria japonica (Phaeophyceae). Aquac. Res. 38:1323–9.CrossRefGoogle Scholar
Mobley, C. D. (1989). A numerical model for the computation of radiance distributions in natural waters with windroughened surfaces. Limnol. Oceanogr. 34: 1473–83.CrossRefGoogle Scholar
Moe, R. L., and Silva, P. C. (1981). Morphology and taxonomy of Himanthothallus (including Phaeoglossum and Phyllogigas), an Antarctic member of the Desmarestiales (Phaeophyceae). J. Phycol. 17: 15–29.CrossRefGoogle Scholar
Molis, M., Körner, J., Ko, Y. W., Kim, J. H., and Wahl, M. (2006). Inducible responses in the brown seaweed Ecklonia cava: the role of grazer identity and season. J. Ecol. 94:243–9.CrossRefGoogle Scholar
Molis, M., Wessels, H., Hagen, W., et al. (2009). Do sulphuric acid and the brown alga Desmarestia viridis support community structure in Arctic kelp patches by altering grazing impact, distribution patterns, and behaviour of sea urchins? Polar Biol. 32:71–82.CrossRefGoogle Scholar
Moncreiff, C. A., and Sullivan, M. J. (2001). Trophic importance of epiphytic algae in subtropical seagrass beds: evidence from multiple stable isotope analyses. Mar. Ecol. Prog. Ser. 215:93–106.CrossRefGoogle Scholar
Monro, K., and Poore, A. G. B. (2009a). The potential for evolutionary responses to cell-lineage selection on growth form and its plasticity in a red seaweed. Am. Nat. 173:151–63.CrossRefGoogle Scholar
Monro, K., and Poore, A. G. B. (2009b). Performance benefits of growth-form plasticity in a clonal red seaweed. Biol. J. Linn. Soc. 97:80–9.CrossRefGoogle Scholar
Monro, K., Poore, A. G. B., and Brooks, R. (2007). Multivariate selection shapes environment-dependent variation in the clonal morphology of a red seaweed. Evol. Ecol. 21:765–82.CrossRefGoogle Scholar
Monteiro, C. A., Engelen, A. H., and Santos, R. (2009). Macro- and mesoherbivores prefer native seaweeds over the invasive brown seaweed Sargassum muticum: a potential regulating role on invasions. Mar. Biol. 156:2505–15.CrossRefGoogle Scholar
Moon, D. A., and Goff, L. J. (1997). Molecular characterization of two large DNA plasmids in the red alga Porphyra pulchra. Curr. Genet. 32:132–8.CrossRefGoogle ScholarPubMed
Morel, F. M. M., Rueter, J. G., Anderson, D. M., and Guillard, R. R. L. (1979). Aquil: a chemically defined phytoplankton culture medium for trace metal studies. J. Phycol. 15:135–41.CrossRefGoogle Scholar
Moreno, C. A., and Jaramillo, E. (1983). The role of grazers in the zonation of intertidal macroalgae on the Chilean coast. Oikos 41:73–6.CrossRefGoogle Scholar
Morris, C.A., Nicolaus, B., Sampson, V., Harwood, J. L., and Kille, P. (1999). Identification and characterization of a recombinant metallothionein protein from a marine alga, Fucus vesiculosus. Biochem. J. 338:553–60.CrossRefGoogle ScholarPubMed
Moss, B. (1964). Wound healing and regeneration in Fucus vesiculosus L. Proc. Intl. Seaweed Symp. 4: 117–22.Google Scholar
Moss, B. (l974). Attachment and germination of the zygotes of Pelvetia canaliculata (L.) Dcne. et Thur. (Phaeophyceae, Fucales). Phycologia 13:317–22.CrossRefGoogle Scholar
Moss, B. L. (1982). The control of epiphytes by Halidrys siliquosa (L.) Lyngb. (Phaeophyta, Cystoseiraceae). Phycologia 21:185–91.CrossRefGoogle Scholar
Moss, B. L. (1983). Sieve elements in the Fucales. New Phytol 93:433–7.CrossRefGoogle Scholar
Motomura, T. (1990). Ultrastructure of fertilization in Laminaria angustata (Phaeophyta, Laminariales) with emphasis on the behavior of centrioles, mitochondria and chloroplasts of the sperm. J. Phycol. 26:80–9.CrossRefGoogle Scholar
Motomura, T., and Nagasato, C. (2004). The first spindle formation in brown algal zygotes. Hydrobiologia 512:171–6.CrossRefGoogle Scholar
Motomura, T., and Sakai, Y. (1988). The occurrence of flagellated eggs in Laminaria angustata (Phaeophyta, Laminariales). J. Phycol. 24:282–5.Google Scholar
Motomura, T., Nagasato, C., and Kimura, K. (2010). Cytoplasmic inheritance of organelles in brown algae. J. Plant Res. 123:185–92.CrossRefGoogle ScholarPubMed
Muhlin, J. F., Engel, C. R., Stessel, R., Weatherbee, R. A., and Brawley, S. H. (2008). The influence of coastal topography, circulation patterns, and rafting in structuring populations of an intertidal alga. Mol. Ecol. 17:1198–210.CrossRefGoogle ScholarPubMed
Mulholland, M. R., and Lomas, M. W. (2008). Nitrogen uptake and assimilation. In Capone, D.G., Bronk, D.A., Mulholland, M.R., and Carpenter, E.J. (eds), Nitrogen in the Marine Environment (pp. 303–84). New York: Academic Press.CrossRefGoogle Scholar
Müller, D. G. (1963). Die Temperaturabhangigkeit der Sporangienbildung bei Ectocarpus siliculosus von verschiedenen Standorten. Publ. Staz. Zool. Napoli 33:310–14.Google Scholar
Müller, D. G. (1981). Sexuality and sex attraction. In Lobban, C. S. and Wynne, M. J. (eds), The Biology of Seaweeds (pp. 661–74). Oxford: Blackwell Scientific.Google Scholar
Müller, D. G. (1989). The role of pheromones in sexual reproduction of brown algae. In Coleman, A. W., Goff, L., and 1. Stein-Taylor, R. (eds), Algae as Experimental Systems (pp. 201–13). New York: Alan R. Liss.Google Scholar
Müller, D. G., Maier, I., and Gassmann, G. (1985). Survey on sexual pheromone specificity in Laminariales (Phaeophyceae). Phycologia 24:475–7.CrossRefGoogle Scholar
Müller, R., Wiencke, C., and Bischof, K. (2008). Interactive effects of UV radiation and temperature on microstages of Laminariales (Phaeophyceae) from the Arctic and North Sea. Clim. Res. 37: 203–13. [, , ]CrossRefGoogle Scholar
Müller, R., Wiencke, C., Bischof, K., and Krock, B. (2009a). Zoospores from three Arctic Laminariales under different UV radiation and temperature conditions: exceptional spectral absorbance properties and lack of phlorotannin induction. Photochem. Photobiol. 85: 970–7.CrossRefGoogle ScholarPubMed
Müller, R., Laepple, T., Bartsch, I., and Wiencke, C. (2009b). Impact of oceanic warming on the distribution of seaweeds in polar and cold-temperate watersBot. Mar. 52 617–38. [, , , ]CrossRefGoogle Scholar
Mumby, P. J., and Steneck, R. S. (2011). The resilience of coral reefs and its implications for reef management. In Dubinsky, Z., and Stambler, N. (eds), Coral Reefs: An Ecosystem in Transition (pp. 509–19). Dordrecht, The Netherlands: Springer.CrossRefGoogle Scholar
Mumford, T. E. (1990). Nori cultivation in North America: growth of the industry. Hydrobiol. 204/205:89–98.CrossRefGoogle Scholar
Mumford, T. E., and Miura, A. (1988). Porphyra as food: cultivation and economics. In Lembi, C. A. and Waaland, J. R. (eds), Algae and Human Affairs (pp. 87–117). Cambridge: Cambridge University Press.Google Scholar
Munda, I. M. (1984). Salinity dependent accumulation of Zn, Co and Mn in Scytosiphon lomentaria (Lyngb.) Link and Enteromorpha intestinalis (L.) from the Adriatic Sea. Bot. Mar. 27:371–6.CrossRefGoogle Scholar
Munda, I. M., and Hudnik, V. (1986). Growth response of Fucus vesiculosus to heavy metals, singly and in dual combinations, as related to accumulation. Bot. Mar. 29:401–12.CrossRefGoogle Scholar
Munda, I. M., and Hudnik, V. (1988). The effects of Zn, Mn, and Co accumulation on growth and chemical composition of Fucus vesiculosus L. under different temperature and salinity conditions. Mar. Ecol. 9:213–25.CrossRefGoogle Scholar
Munns, R., Greenway, H., and Kirst, G. O. (1983). Halotolerant eukaryotes. In Pirson, A., and Zimmerman, M. H. (eds) Encyclopedia of Plant Physiology. Vol. 12: Physiological Plant Ecology III (pp. 59–135). Berlin: Springer-Verlag.CrossRefGoogle Scholar
Muñoz, J., Cancino, J. M., and Molina, M. X. (1991). Effect of encrusting bryozoans on the physiology of their algal substratum. J. Mar. Biol. Assoc. UK. 71:877–82.CrossRefGoogle Scholar
Murray, S. N., and Dixon, P. S. (1992). The Rhodophyta: some aspects of their biology. III. Oceanogr. Mar. Biol. 30:1–148. [, , ]Google Scholar
Murray, S. N., and Littler, M. M. (1978). Patterns of algal succession in a perturbated marine intertidal community. J. Phycol. 14:506–12.CrossRefGoogle Scholar
Murray, S. N., Ambrose, R. F., and Dethier, M. N. (2006). Monitoring Rocky Shores. Princeton, NJ: University of California Press, 240 pp.CrossRefGoogle Scholar
Murthy, M. S., Ramakrishna, T., Sarat Babu, G. V., and Rao, Y. N. (1986). Estimation of net primary productivity of intertidal seaweeds: limitations and latent problems. Aquat. Bot. 23:383–7.CrossRefGoogle Scholar
Mutchler, T., Sullivan, M. J., and Fry, B. (2005). Potential of N-14 isotope enrichment to resolve ambiguities in coastal trophic relationships. Mar. Ecol. Prog. Ser. 266:27–33.CrossRefGoogle Scholar
Myers, J. H., Gunthorpe, L., Allinson, G., and Duda, S. (2006). Effects of antifouling biocides to the germination and growth of the marine macroalga Hormosira banksii (Turner) Desicaine. Mar. Poll. Bull. 52:1048–55.CrossRefGoogle ScholarPubMed
Myers, J. H., Duda, S., Gunthorpe, L., and Allinson, G. (2007). Evaluation of the Hormosira banksii (Turner) Desicaine germination and growth inhibition bioassay for use as a regulatory assay. Chemosphere 69:955–60.CrossRefGoogle ScholarPubMed
Nagasato, C. (2005). Behavior and function of paternally inherited centrioles in brown algal zygotes. J. Plant Res. 118:361–9.CrossRefGoogle ScholarPubMed
Nagasato, C., Inoue, A., Mizuno, M., et al. (2010). Membrane fusion process and assembly of cell wall during cytokinesis in the brown alga, Silvetia babingtonii (Fucales, Phaeophyceae). Planta. 232:287–98.CrossRefGoogle Scholar
Nagashima, H., Nakamura, S., Nisizawa, K., and Hori, T. (1971). Enzymic synthesis of floridean starch in a red alga, Serraticardia maxima. Plant Cell Physiol. 12:243–53.Google Scholar
Nakahara, H., and Nakamura, Y. (1973). Parthenogenesis, apogamy and apospory in Alaria crassifolia (Laminariales). Mar. Biol. 18:327–32.Google Scholar
Nakajima, Y., Endo, Y., Ionue, Y., et al. (2006). Ingestion of Hijiki seaweed and risk of arsenic poisoning. Appl. Organomet. Chem. 20:557–64.CrossRefGoogle Scholar
Naldi, M., and Wheeler, P. A. (1999). Changes in nitrogen pools in Ulva fenestrata (Chlorophyta) and Gracilaria pacifica (Rhodophyta) under nitrate and ammonium enrichment. J. Phycol. 35:70–7.CrossRefGoogle Scholar
Naldi, M., and Wheeler, P. A. (2002). 15N measurements of ammonium and nitrate uptake by Ulva fenestrata (Chlorophyta) and Gracilaria pacifica (Rhodophyta): comparison of net nutrient disappearance, release of ammonium and nitrate and 15N accumulation in algal tissue. J. Phycol. 38:135–44.CrossRefGoogle Scholar
Nanba, N., Kado, R., and Ogawa, H. (2005). Effects of irradiance and water flow on formation and growth of spongy and filamentous thalli of Codium fragile. Aquat. Bot. 81:315–25.CrossRefGoogle Scholar
Navarrete, S. A., Broitman, B. R., Wieters, E. A., and Castilla, J. C. (2005). Scales of benthic-pelagic coupling and the intensity of species interactions: from recruitment limitation to top down control. Proc. Nat. Acad. Sci. USA 102:18,046–51.CrossRefGoogle ScholarPubMed
Neefus, C., Mathieson, A. C., Bray, T. L., and Yarish, C. (2008). The occurrence of three introduced Asiatic species of Porphyra (Bangiales, Rhodophyta) in the northwestern Atlantic. J. Phycol. 44:1399–414.CrossRefGoogle ScholarPubMed
Neill, K., Heesch, S., and Nelson, W. (2008). Diseases, Pathogens and Parasites of Undaria pinnatifida. Ministry of Agriculture and Forestry, Biosecurity New Zealand, Tech. Paper No. 2009/44. Available at . [, , ]Google Scholar
Neish, A. C., Shacklock, P. F., Fox, C. H., and Simpson, F. J. (1977). The cultivation of Chondrus crispus. Factors affecting growth under greenhouse conditions. Can. J. Bot. 55:2263–71.CrossRefGoogle Scholar
Neiva, J., Pearson, G. A., Valero, M., and Serrão, E. A. (2010). Surfing the wave on a borrowed board: range expansion and spread of introgressed organellar genomes in the seaweed Fucus ceranoides L. Mol. Ecol. 19:4812–22.CrossRefGoogle ScholarPubMed
Nelson, S. G., and Siegrist, A. W. (1987). Comparison of mathematical expressions describing light-saturation curves for photosynthesis by tropical marine macroalgae. Bull. Mar. Sci. 41:617–22.Google Scholar
Nelson, T. A., Lee, D. J., and Smith, B. C. (2003). Are “green tides” harmful algal blooms? Toxic properties of water-soluble extracts from two bloom-forming macroalgae, Ulva fenestrata and Ulvaria obscura (Ulvophyceae). J. Phycol. 39:874–9.CrossRefGoogle Scholar
Nelson, W. A. (2005). Life history and growth in culture of the endemic New Zealand kelp Lessonia variegata J. Agardh in response to differing regimes of temperature, photoperiod and light. J. Appl. Phycol. 17:23–8.CrossRefGoogle Scholar
Nelson, W.A. (2009). Calcified macroalgae – critical to coastal ecosystems and vulnerable to climate change: a review. Mar. Freshwat. Res. 60:787–801. [, , ]CrossRefGoogle Scholar
Nelson, W. A., Brodie, J., and Guiry, M. D. (1999). Terminology used to describe reproduction and life history stages in the genus Porphyra (Bangiales, Rhodophyta). J. Appl. Phycol. 11:407–10.CrossRefGoogle Scholar
Nelson, W. G. (1982). Experimental studies of oil pollution on the rocky intertidal community of a Norwegian fjord. J. Exp. Mar. Biol. Ecol. 65:121–38.CrossRefGoogle Scholar
Nelson-Smith, A. (1972). Oil Pollution and Marine Ecology. London: Elek Science Press, 260 pp.Google Scholar
Neori, A., Chopin, T., Troell, M., et al. (2004). Integrated aquaculture: rationale, evolution and state of the art emphasizing seaweed biofiltration in modern mariculture. Aquaculture 321:361–91.CrossRefGoogle Scholar
Neori, A., Troell, M., Chopin, T., et al. (2007) The need for a balanced ecosystem approach to blue revolution aquaculture. Environment 49:36–43.Google Scholar
Neushul, M. (1972). Functional interpretation of benthic marine algal morphology. In Abbott, I. A. and Kurogi, M. (eds), Contributions to the Systematics of Benthic Marine Algae of the North Pacific (pp. 47–73). Tokyo: Japan Society for Phycology.Google Scholar
Neushul, M. (1981). The ocean as a culture dish: experimental studies of marine algal ecology. Proc. Intl. Seaweed Symp. 8:19–35.Google Scholar
Neville, A. C. (1988). The helicoidal arrangement of microfibrils in some algal cell walls. Prog. Phycol. Res. 6:1–21.Google Scholar
Newcombe, E. M., and Taylor, R. B. (2010) Trophic cascade in a seaweed-epifauna-fish food chain. Mar. Ecol. Prog. Ser. 408:161–7.CrossRefGoogle Scholar
Nezlin, N. P., Kamer, K., and Stein, E. D. (2007). Application of color infrared aerial photography to assess macroalgal distribution in an eutrophic estuary, Upper Newport Bay, California. Estuar. Coasts 30:855–68.CrossRefGoogle Scholar
Nicotri, M. E. (1980). Factors involved in herbivore food preference. J. Exp. Mar. Biol. Ecol. 42:13–26.CrossRefGoogle Scholar
Niell, F. X. (1976). C:N ratio in some marine macrophytes and its possible ecological significance. Bot. Mar. 14:347–50.Google Scholar
Nielsen, H. D., and Nielsen, S. L. (2008). Evaluation of imaging and conventional PAM as a measure of photosynthesis in thin- and thick-leaved marine macroalgae. Aquat. Biol. 3:121–31.CrossRefGoogle Scholar
Nielsen, H. D., and Nielsen, S. L. (2010). Adaptation to high light irradiances enhances the photosynthetic Cu2+ resistance in Cu2+ tolerant and non-tolerant populations of the brown macroalga Fucus serratus. Mar. Poll. Bull. 60:710–17.CrossRefGoogle Scholar
Nielsen, H. D., Brownlee, C., Coelho, S. M., and Brown, M. T. (2003a). Inter-population differences in inherited copper tolerance involve photosynthetic adaptation and exclusion mechanisms in Fucus serratus. New Phytol. 160:157–65.CrossRefGoogle Scholar
Nielsen, H. D., Brown, M. T., and Brownlee, C. (2003b). Cellular responses of developing Fucus serratus embryos exposed to elevated concentrations of Cu. Pl. Cell Environ. 26:1737–44.CrossRefGoogle Scholar
Nielsen, H. D., Burridge, T. R., Brownlee, C., and Brown, M. T. (2005). Prior exposure to Cu contamination influences the outcome of toxicological testing of Fucus serratus embryos. Mar. Poll. Bull. 50:1675–80.CrossRefGoogle ScholarPubMed
Nielsen, K. J. (2001). Bottom-up and top-down forces in tide pools: test of a food chain model in an intertidal community. Ecol. Monogr. 71:187–217.CrossRefGoogle Scholar
Nielsen, K. J., and Navarrete, S. A. (2004). Mesoscale regulation comes from the bottom-up: intertidal interactions between consumers and upwelling. Ecol. Lett. 7:31–41.CrossRefGoogle Scholar
Nienhuis, P. H. (1987). Ecology of salt-marsh algae in the Netherlands. In Huiskes, A. H. L., Blom, C. W. P. M., and Rozema, J. (eds), Vegetation Between Land and Sea (pp. 66–83). Dordrecht, The Netherlands: Dr. W. Junk.CrossRefGoogle Scholar
Niklas, K. J. (2009). Functional adaptation and phenotypic plasticity at the cellular and whole plant level. J. Biosci. 34:613–20. [, , ]CrossRefGoogle ScholarPubMed
Nilsen, G., and Nordby, Ø. (1975). A sporulation-inhibiting substance from vegetative thalli of the green alga, Ulva mutabilis Føyn. Planta 125:127–39.Google ScholarPubMed
Nishihara, G. N., and Ackerman, J. D. (2006). The effect of hydrodynamics on the mass transfer of dissolved inorganic carbon to the freshwater macrophyte Vallisneria americana. Limnol. Oceanogr. 51:2734–45.CrossRefGoogle Scholar
Nishihara, G. N., and Ackerman, J. D. (2007). On the determination of mass transfer in a concentration boundary layer. Limnol. Oceanogr. Methods 5:88–96.CrossRefGoogle Scholar
Nishihara, G. N., and Terada, R. (2010). Species richness of marine macrophytes is correlated to a wave exposure gradient. Phycol. Res. 58:280–92.CrossRefGoogle Scholar
Nishikawa, T., and Yamaguchi, M. (2008). Effect of temperature on light-limited growth of the harmful diatom Coscinodiscus wailesii, a causative organism in the bleaching of aquacultured Porphyra thalli. Harmful Algae 7:561–6.CrossRefGoogle Scholar
Nishimura, N. J., and Mandoli, D. F. (1992). Population analysis of reproductive cell structures of Acetabularia acetabulum (Chlorophyta). Phycologia 31(3/4):351–8.CrossRefGoogle Scholar
Nisizawa, K., Noda, H., Kikuchi, R., and Watanabe, T. (1987). The main seaweed foods in Japan. Hydrobiologia 151/152:5–29. [, , ]CrossRefGoogle Scholar
Niwa, K., and Sakamoto, T. (2010). Allopolyploidy in natural and cultivated populations of Porphyra (Bangiales, Rhodophyta). J. Phycol. 46:1097–105.CrossRefGoogle Scholar
Niwa, K., Furuita, H., and Yamamoto, T. (2008). Changes of growth characteristics and free amino acid content of cultivated Porphyra yezoensis Ueda (Bangiales Rhodophyta) blades with the progression of the number of harvests in a nori farm. J. Appl. Phycol. 20:687–93.CrossRefGoogle Scholar
Niwa, K., Iida, S., Kato, A., et al. (2009a). Genetic diversity and introgression in two cultivated species (Porphyra yezoensis and Porphyra tenera) and closely related wild species of Porphyra (Bangiales, Rhodophyta). J. Phycol. 45:493–502.CrossRefGoogle Scholar
Niwa, K., Hayashi, Y., Abe, T., and Aruga, Y. (2009b). Induction and isolation of pigmentation mutants of Porphyra yezoensis (Bangiales, Rhodophyta) by heavy-ion beam irradiation. Phycol. Res. 57:194–202.CrossRefGoogle Scholar
Niwa, K., Yamamoto, T., Furuita, H., and Abe, T. (2011). Mutation breeding in the marine crop Porphyra yezoensis (Bangiales, Rhodophyta): cultivation experiment of the artificial red mutant isolated by heavy ion beam mutagenesis. Aquaculture 314:182–7.CrossRefGoogle Scholar
Nöel, L. M.-L. J., Hawkins, S. J., Stuart, S. R., et al. (2009). Grazing dynamics in intertidal rockpools: connectivity of microhabitats. J. Exp. Mar. Biol. Ecol. 370:9–17.CrossRefGoogle Scholar
North, W. J. (1987) Biology of the Macrocystis resource in North America. In M. S. Doty, 1. F. Caddy, and B. Santelices (eds.), Case Studies of Seven Commerical Seaweed Resources (pp. 265–311). FAO Fish. Tech. Pap. 281.
Norton, T. A. (1977). Ecological experiments with Sargassum muticum. J. Mar. Biol. Assoc. UK 57:33–43.CrossRefGoogle Scholar
Norton, T. A. (1991). Conflicting constraints on the form of intertidal algae. Brit. Phycol. J. 26:203–18.CrossRefGoogle Scholar
Norton, T. A. (1992). Dispersal by algae. Br. Phycol. J. 27:293–301.CrossRefGoogle Scholar
Norton, T. A., and Fetter, R. (1981) The settlement of Sargassum muticum propagules in stationary and flowing water. J. Mar. Biol. Assoc. UK 61:929–40.CrossRefGoogle Scholar
Norton, T. A., and Mathieson, A. C. (1983). The biology of unattached seaweeds. Prog. Phycol. Res. 2:333–86.Google Scholar
Norton, T. A., Mathieson, A. C., and Neushul, M. (1981). Morphology and environment. In Lobban, C. S. and Wynne, M. J. (eds), The Biology of Seaweeds (pp. 421–51). Oxford: Blackwell Scientific.Google Scholar
Nott, A., Jung, H.-S., Koussevitzky, S., and Chory, J. (2006). Plastid-to-nucleus retrograde signaling. Annu. Rev. Plant Biol. 57:739–59.CrossRefGoogle ScholarPubMed
Nugues, M. M., and Bak, R. P. M. (2006). Differential competitive abilities between Caribbean coral species and a brown alga: a year of experiments and a long-term perspective. Mar. Ecol. Prog. Ser. 315:75–86.CrossRefGoogle Scholar
Nultsch, W., Pfau, J., and Rüffer, U. (1981). Do correlations exist between chromatophore arrangement and photosynthetic activity in seaweeds? Mar Biol 62: 111–17.CrossRefGoogle Scholar
Nyberg, C. D., and Wallentinus, I. (2005). Can species traits be used to predict marine macroalgal introductions? Biol. Invasions 7:265–79.CrossRefGoogle Scholar
Nylund, G. M., and Pavia, H. (2005). Chemical versus mechanical inhibition of fouling in the red alga Dilsea carnosa. Mar. Ecol. Prog. Ser. 299:111–21.CrossRefGoogle Scholar
Nylund, G. M., Cervin, G., Hermansson, M., and Pavia, H. (2005). Chemical inhibition of bacterial colonization by the red alga Bonnemaisonia hamifera. Mar. Ecol. Prog. Ser. 302:27–36.CrossRefGoogle Scholar
Oates, B. R. (1988). Water relations of the intertidal saccate alga Colpomenia peregrina (Phaeophyta, Scytosiphonales). Bot. Mar. 31:57–63.CrossRefGoogle Scholar
Oates, B. R. (1989). Articulated coralline algae as a refuge for the intertidal saccate species, Colpomenia peregrina and Leathesia difformis in southern California. Bot. Mar. 32:475–8.CrossRefGoogle Scholar
Oates, B. R., and Cole, K. M. (1994). Comparative studies on hair cells of two agarophyte algae: Gelidium vagum (Gelidiales, Rhodophyta) and Gracilaria pacifica (Gracilariales, Rhodophyta). Phycologia 33:420–33.CrossRefGoogle Scholar
O’Brien, M. C., and Wheeler, P. A. (1987). Short-term uptake of nutrients by Enteromorpha prolifera (Chlorophyceae). J. Phycol. 23:547–56.CrossRefGoogle Scholar
O’Brien, P. Y., and Dixon, P. S. (1976). The effects of oils and oil components on algae: a review. Br. Phycol. J. 11: 115–42.CrossRefGoogle Scholar
Ogata, E. (1971). Growth of conchocelis in artificial medium in relation to carbon dioxide and calcium metabolism. J. Shimonoseki U. Fish. 19:123–9.Google Scholar
Ogawa, H. (1984). Effects of treated municipal wastewater on the early development of sargassaceous plants. Hydrobiologia 116/117:389–92.CrossRefGoogle Scholar
Ogden, J. C., Brown, R. A., and Salesky, N. (1973). Grazing by the echinoid Diadema antillarum Philippi: formation of halos around West Indian patch reefs. Science 182:715–17.CrossRefGoogle ScholarPubMed
Okabe, Y., and Okada, M. (1990). Nitrate reductase activity and nitrate in native pyrenoids purified from the green algaBryopsis maxima. Plant Cell Physiol. 31:429–32.Google Scholar
O’Kelly, C. J., and Baca, B. J. (1984). Time course of carpogonial branch and carposporophyte development in Callithamnion cordatum (Rhodophyta) Ceramiales. Phycologia 23:407–17.CrossRefGoogle Scholar
Okuda, T., Noda, T., Yamamoto, T., Hori, M., and Nakaoka, M. (2010). Contribution of environmental and spatial processes to rocky intertidal metacommunity structure. Acta Oecol. 36:413–22.CrossRefGoogle Scholar
Olson, A. M., and Lubchenco, J. (1990). Competition in seaweeds: linking plant traits to competitive outcomes. J. Phycol. 26:1–6. [4.0]CrossRefGoogle Scholar
Oltmanns, F. (1892). Ueber die Cultur-und Lebensbedingungen der Meeresalgen. Jahr. Wissensch Bot. 23:349–440.Google Scholar
Oohusa, T. (1980). Diurnal rhythm in the rates of cell division, growth and photosynthesis of Porphyra yezoensis (Rhodophyceae) cultured in the laboratory. Bot. Mar. 23:1–5.CrossRefGoogle Scholar
Oppliger, L. V., Correa, J. A., and Peters, A. (2007). Parthenogenesis in the brown alga Lessonia nigrescens (Laminariales, Phaeophyceae) from Central Chile. J. Phycol. 43:1295–301.CrossRefGoogle Scholar
Orduña-Rojas, J., and Robledo, D. (1999). Effects of irradiance and temperature on the release and growth of carpospores from Gracilaria cornea J. Agardh (Gracilariales, Rhodophyta). Bot. Mar. 42:315–19.CrossRefGoogle Scholar
Osborne, B. A., and Raven, J. A. (1986). Light absorption by plants and its implications for photosynthesis. Biol. Rev. 61:1–61. [, , ]CrossRefGoogle Scholar
Osmond, C. B. (1994). What is photoinhibition? Some insights from comparisons of shade and sun plants. In Baker, N. R. and Bowyer, N. R. (eds), Photoinhibition of Photosynthesis, From the Molecular Mechanisms to the Field (pp. 1–24). Oxford: BIOS Scientific Publ.Google Scholar
Ouriques, L. C., and Bouzon, Z. L. (2003). Ultrastructure of germinating tetraspores of Hypnea musciformis (Gigartinales, Rhodophyta). Plant Biosys. 137:193–202.CrossRefGoogle Scholar
Ouriques, L. C., Schmidt, É. C., and Bouzon, Z. L. (2012). The mechanism of adhesion and germination in the carospores of Porphyra spiralis var. amplifolia (Rhodophyta, Bangiales). Micron 43:269–77.CrossRefGoogle Scholar
Owens, N. J. P. (1987). Natural variations in 15N in the marine environment. Adv. Mar. Biol. 24:389–451.CrossRefGoogle Scholar
Padilla, D. K., and Allen, B. J. (2000). Paradigm lost: reconsidering functional form and group hypotheses in marine ecology. J. Exp. Mar. Biol. Ecol. 250:207–21.CrossRefGoogle ScholarPubMed
Paerl, H. W., Rudek, J., and Mallin, M. A. (1990). Stimulation of phytoplankton in coastal waters by natural rainfall inputs: nutritional and trophic implications. Mar. Biol. 107:247–54.CrossRefGoogle Scholar
Paine, R. T. (1966). Food web complexity and species diversity. Am. Nat. 100:65–75.CrossRefGoogle Scholar
Paine, R.T. (1979). Disaster, catastrophe, and local persistence of the sea palmPostelsia palmaeformis. Science 205:685–7.Google ScholarPubMed
Paine, R. T. (1986). Benthic community-water column coupling during the 1982–1983 El Nino. Are community changes at high latitudes attributable to cause or coincidence? Limnol. Oceanogr. 31:351–60.CrossRefGoogle Scholar
Paine, R. T. (1988). Habitat suitability and local population persistence of the sea palmPostelsia palmaeformis. Ecology 69:1787–94.CrossRefGoogle Scholar
Paine, R. T. (1990). Benthic macroalgal competition: complications and consequences. J. Phycol. 26:12–17. [4.0, ]CrossRefGoogle Scholar
Paine, R. T. (1994). Marine Rocky Shores and Community Ecology: An Experimentalist’s Perspective, Excellence in Ecology 4. Oldendorf/Luhe, Germany: Ecology Institute, 152 pp. [, , , ]Google Scholar
Paine, R. T. (2010). Macroecology: does it ignore or can it encourage further ecological synthesis based on spatially local experimental manipulations? Am. Nat. 176:385–93. [,, ]CrossRefGoogle ScholarPubMed
Paine, R. T., and Levin, S. A. (1981). Intertidal landscapes: disturbance and the dynamics of pattern. Ecol. Monogr. 51:145–78. [,, , ]CrossRefGoogle Scholar
Paine, R. T., Ruesink, J. L., Sun, A., et al. (1996). Trouble in oiled waters: lessons from the Exxon Valdez oil spill. Ann. Rev. Ecol. Syst. 27:197–235.CrossRefGoogle Scholar
Pak, J. Y., Solorzano, C., Arai, M., and Nitta, T. (1991). Two distinct steps for spontaneous generation of subprotoplasts from a disintegrated Bryopsis cell. Plant Physiol. 96:819–25.CrossRefGoogle ScholarPubMed
Palmieri, M., and Kiss, J. Z. (2007). The role of plastids in gravitropism. In Wise, R. R. and Hoober, J. K. (eds), The Structure and Function of Plastids (pp. 507–25). Dordrecht, The Netherlands: Springer.CrossRefGoogle Scholar
Palumbi, S. R. (1984). Measuring intertidal wave forces. J. Exp. Mar. Biol. Ecol. 81:171–9.CrossRefGoogle Scholar
Palumbi, S. R. (2001). The ecology of marine protected areas. In Bertness, M. D., Gaines, S. D., and Hay, M. E. (eds), Marine Community Ecology (pp. 509–30). Sunderland, MA: Sinauer Associates, Inc.Google Scholar
Pandolfi, J. M., Bradbury, R. H., Sala, E., et al. (2003). Global trajectories of the long-term decline of coral reef ecosystems. Science 301:955–8.CrossRefGoogle ScholarPubMed
Pang, S.-J., and Lüning, K. (2004). Photoperiodic long-day control of sporophyll and hair formation in the brown alga Undaria pinnatifida. J. Appl. Phycol. 16:83–92.CrossRefGoogle Scholar
Papenfuss, G. F. (1958). Die Gattungen der Rhodophyceen. By Harold Kylin. Bull. Torrey Botan. Club 85: 142–3.CrossRefGoogle Scholar
Pareek, M., Mishra, A., and Jha, B. (2010). Molecular phylogeny of Gracilaria species inferred from molecular markers belonging to three different genomes. J. Phycol. 46:1322–8.CrossRefGoogle Scholar
Park, C. S., Park, K. Y., Back, J. M., and Hwang, E. K. (2008). The occurrence of pinhole disease in relation to developmental stage in cultivated Undaria pinnatifida (Harvey) Suringar (Phaeophyta) in Korea. J. Appl. Phycol. 20:485–90.CrossRefGoogle Scholar
Park, H. S., Jeong, W. J., Kim, E. C., et al. (2011). Heat shock protein gene family of the Porphyra seriata and enhancement of heat stress tolerance by PsHSP70 in Chlamydomonas. Mar. Biotechnol. 14:332–42.CrossRefGoogle ScholarPubMed
Parke, M. W. (1948). Studies of the British Laminariaceae. I. Growth in Laminaria saccharina (L.) Lamour. J. Mar. Biol. Assoc. UK 27:651–709.CrossRefGoogle Scholar
Parsons, T. R., Takahashi, M., and Hargrave, B. (1977). Biological Oceanographic Processes (2nd edn). New York: Pergamon Press, 332 pp.Google Scholar
Parsons, T. R., Maita, Y., and Lalli, C. M. (1984). A Manual of Chemical and Biological Methods of Seawater Analysis. New York: Pergamon Press, 173 pp.Google Scholar
Pastorok, R. A., and Bilyard, G. R. (1985). Effects of sewage pollution on coral reef communities. Mar. Ecol. Prog. Ser. 21:175–89.CrossRefGoogle Scholar
Patterson, D. J. (1989a). Stramenopiles: chromophytes from a protistan perspective. In Green, J. C., Leadbeater, B. S. C., and Diver, W. L. (eds), The Chromophyte Algae: Problems and Perspectives (Systematics Association Special Volume 38, pp. 357–79). Oxford: Clarendon Press.Google Scholar
Patterson, M. R. (1989b). Nearshore biomechanics [review of Denny’s book]. Science 243:1374. [8.0]CrossRefGoogle Scholar
Patwary, M. U., and van der Meer, J. P. (1994). Application of RAPD markers in an examination of heterosis in Geldium vagum (Rhodophyta). J. Phycol. 30:91–7.CrossRefGoogle Scholar
Paul, N. A., Cole, L., de Nys, R., and Steinberg, P. D. (2006a). Ultrastructure of the gland cells of the red alga Asparagopsis armata (Bonnemaisoniaceae). J. Phycol. 42:637–45.CrossRefGoogle Scholar
Paul, N.A., de Nys, R., and Steinberg, P. D. (2006b). Chemical defense against bacteria in the red alga Asparagopsis armata: linking structure with function. Mar. Ecol. Prog. Ser. 306:87–101.CrossRefGoogle Scholar
Paul, V. J., and Hay, M. E. (1986). Seaweed susceptibility to herbivory: chemical and morphological correlates. Mar. Ecol. Prog. Ser. 33:255–64.CrossRefGoogle Scholar
Paul, V. J., and Van Alstyne, K. L. (1988a). Chemical defense and chemical variation in some tropical Pacific species of Halimeda (Halimedaceae; Chlorophyta). Coral Reefs 6:263–9.CrossRefGoogle Scholar
Paul, V. J., and Van Alstyne, K. L. (1988b). Use of ingested algal diterpenoids by Elysia halimedae Macnae (Opisthobranchia: Ascoglossa) as antipredator defenses. J. Exp. Mar. Biol. Ecol. 119:15–29.CrossRefGoogle Scholar
Paul, V. J., and Van Alstyne, K. L. (1992). Activation of chemical defenses in the tropical green algae Halimeda spp. J. Exp. Mar. Biol. Ecol. 160:191–203.CrossRefGoogle Scholar
Paul, V. J., Puglisi, M. P., and Ritson-Williams, R. (2006c). Marine chemical ecology. Nat. Prod. Rep. 23:153–80.CrossRefGoogle ScholarPubMed
Paul, V. J., Ritson-Williams, R., and Sharp, K. (2011). Marine chemical ecology in benthic environments. Nat. Prod. Rep. 28:345–87.CrossRefGoogle ScholarPubMed
Pavia, H., and Toth, G. B. (2000). Inducible chemical resistance to herbivory in the brown seaweed Ascophyllum nodosum. Ecology 81:3212–25.CrossRefGoogle Scholar
Pavia, H., and Toth, G. B. (2008). Macroalgal models in testing and extending defense theories. In Amsler, C. D. (ed.), Algal Chemical Ecology (pp. 147–72). London: Springer (Limited).CrossRefGoogle Scholar
Pavia, H., Cervin, G., Lindgren, A., and Åberg, P. (1997). Effects of UV-B radiation and simulated herbivory on phlorotannins in the brown alga Ascophyllum nodosum. Mar. Ecol. Prog. Ser. 157:139–46.CrossRefGoogle Scholar
Pavia, H., Toth, G., and Åberg, P. (1999). Trade-offs between phlorotannin production and annual growth in natural populations of the brown seaweed Ascophyllum nodosum. J. Ecol. 87:761–71.CrossRefGoogle Scholar
Pawlik-Skowrońska, B., Pirszel, J., and Brown, M. T. (2007). Concentrations of phytochelatins and glutathione found in natural assemblages of seaweeds depend on species and metal concentrations of the habitat. Aquat. Toxicol. 83:190–9.CrossRefGoogle ScholarPubMed
Pearson, G. A., and Davison, I. R. (1993). Freezing rate and duration determine the physiological response of intertidal fucoids to freezing. Mar Biol 115: 353–62.CrossRefGoogle Scholar
Pearson, G. A., and Davison, I. R. (1994). Freezing stress and osmotic dehydration in Fucus distichus (Phaeophyta): evidence for physiological similarity. J. Phycol. 30:257–67.CrossRefGoogle Scholar
Pearson, G. A., and Evans, L. V. (1990). Settlement and survival of Polysiphonia lanosa (Ceramiales) spores on Ascophyllum nodosum and Fucus vesiculosus (Fucales). J. Phycol. 26:597–603.CrossRefGoogle Scholar
Pearson, G. A., and Serrão, E. A. (2006). Revisiting synchronous gamete release by fucoid algae in the intertidal zone: fertilization success and beyond? Integr. Comp. Biol. 46:587–97. [, , , ]CrossRefGoogle ScholarPubMed
Pearson, G. A., Serrão, E. A., and Brawley, S. H. (1998). Control of gamete release in fucoid algae: sensing hydrodynamic conditions via carbon acquisition. Ecology 79:1725–39.CrossRefGoogle Scholar
Pearson, G., Kautsky, L., and Serrão, E. (2000). Recent evolution in Baltic Fucus vesiculosus: reduced tolerance to emersion stresses compared to intertidal (North Sea) populations. Mar. Ecol. Prog. Ser. 202:67–79.CrossRefGoogle Scholar
Pearson, G. A., Serrão, E. A., and Dring, M. (2004). Blue- and green-light signals for gamete release in the brown alga, Silvetia compressa. Oecologia 133:193–201.CrossRefGoogle Scholar
Pearson, G. A., Lago-Leston, A., and Mota, C. (2009). Frayed at the edges: selective pressure and adaptive response to abiotic stressors are mismatched in low diversity edge populations. J. Ecol. 97:450–62.CrossRefGoogle Scholar
Pearson, G. A., Hoarau, G., Lago-Leston, A., et al. (2010). An expressed sequence tag analysis of the intertidal brown seaweeds Fucus serratus (L.) and F. vesiculosus (L.) (Heterokontophyta, Phaeophyceae) in response to abiotic stressors. Mar. Biotechnol. 12:195–213.CrossRefGoogle Scholar
Peckol, P., and Ramus, J. (1988). Abundances and physiological properties of deep-water seaweeds from Carolina outer continental shelf. J. Exp. Mar. Biol. Ecol. 115:25–39.CrossRefGoogle Scholar
Peckol, P., Krane, J. M., and Yates, J. L. (1996). Interactive effects of inducible defense and resource availability on phlorotannins in the North Atlantic brown alga Fucus vesiculosus. Mar. Ecol. Prog. Ser. 138:209–17.CrossRefGoogle Scholar
Peddigari, S., Zhang, W., Takechi, K., Takano, H., and Takio, S. (2008). Two different clades of copia-like retrotransposons in the red alga, Porphyra yezoensis. Gene 424:153–8.CrossRefGoogle ScholarPubMed
Pedersen, M. F. (1994). Transient ammonium uptake in the macroalga Ulva lactuca (Chlorophyta): nature, regulation, and the consequences for the choice of measuring technique. J. Phycol. 30:980–6. [, , ]CrossRefGoogle Scholar
Pedersen, M. F., and Borum, J. (1996). Nutrient control of algal growth in estuarine waters. Nutrient limitation and the importance of nitrogen requirements and nitrogen storage among phytoplankton and species of macroalgae. Mar. Ecol. Prog. Ser. 142:261–72.CrossRefGoogle Scholar
Pedersen, M. F., and Borum, J. (1997). Nutrient control of estuarine macroalgae: growth strategy and the balance between nitrogen requirements and uptake. Mar. Ecol. Prog. Ser. 161:155–63.CrossRefGoogle Scholar
Pedersen, M. F., and Snoeijs, P. (2001). Patterns of macroalgal diversity, community composition and long-term changes along the Swedish west coast. Hydrobiologia 459:83–102.CrossRefGoogle Scholar
Pedersen, L. B., Geimer, S., and Rosenbaum, J. L. (2006). Dissecting the molecular mechanisms of intraflagellar transport in Chlamydomonas. Curr. Biol. 16:450–9.CrossRefGoogle ScholarPubMed
Pedersen, M. F., Borum, J., and Fotel, F. L. (2010). Phosphorus dynamics and limitation of fast- and slow-growing temperate seaweeds in Oslofjord, Norway. Mar. Ecol. Prog. Ser. 399:103–15. [, , ]CrossRefGoogle Scholar
Pedersen, P. M. (1981). Phaeophyta: life histories. In Lobban, C. S. and Wynne, M. J. (eds), The Biology of Seaweeds (pp. 194–217). Oxford: Blackwell Scientific.Google Scholar
Pelevin, V. N., and Rutkovskaya, V. A. (1977). On the optical classification of ocean waters from the spectral attenuation of solar radiation. Oceanology 17:28–32.Google Scholar
Pellegrini, L. (1980). Cytological studies on physodes in the vegetative cells of Cystoseira stricta Sauvageau (Phaeophyta, Fucales). J. Cell Sci. 41:209–31.[]
Pelletreau, K. N., and Muller-Parker, G. (2002). Sulfuric acid in the phaeophyte alga Desmarestia munda deters feeding by the sea urchin Strongylocentrotus droebachiensis. Mar. Biol. 141:1–9.Google Scholar
Pelletreau, K. N., and Targett, N. M. (2008). New perspectives for addressing patterns of secondary metabolites in marine algae. In Amsler, C. D. (ed.), Algal Chemical Ecology (pp. 121–46). London: Springer (Limited).CrossRefGoogle Scholar
Penela-Arenaz, M., Bellas, J., and Vázquez, E. (2009). Effects of the Prestige oil spill on the biota in NW Spain: 5 years of learning. Adv. Mar. Biol. 56:365–94.CrossRefGoogle Scholar
Pennings, S. C. (l990). Size-related shifts in herbivory: specialization in the sea hare Aplysia californica Cooper. J. Exp. Mar. Biol. Ecol. 142:43–61.CrossRefGoogle Scholar
Pennings, S. C., and Bertness, M. D. (2001). Salt-marsh communities. In Bertness, M. D., Gaines, S. D., and Hay, M. E. (eds), Marine Community Ecology (pp. 289–337). Sunderland, MA: Sinauer Associates, Inc.Google Scholar
Pennings, S. C., Carefoot, T. H., Zimmer, M., Danko, J. P., and Ziegler, A. (2000). Feeding preferences of supralittoral isopods and amphipods. Canadian Journal of Zoology-Revue Canadienne De Zoologie 78(11): 1918–29.CrossRefGoogle Scholar
Pentecost, A. (1985). Photosynthetic plants as intermediary agents between environmental HCO3- and carbonate deposition. In Lucas, W. J. and Berry, J. A. (eds), Inorganic Carbon Uptake by Aquatic Photosynthetic Organisms (pp. 459–80). Bethesda, MD: American Society of Plant Physiologists.Google Scholar
Percival, E. (1979). The polysaccharides of green, red and brown seaweeds: their basic structure, biosynthesis and function. Br. Phycol. J. 14:103–17. [, , ]CrossRefGoogle Scholar
Percival, E., and McDowell, R. H. (1967). Chemistry and Enzymology of Marine Algal Polysaccharides. New York: Academic Press, 219 pp.Google Scholar
Pereira, L., Critchley, A. T., Amado, A. M., and Ribeiro-Claro, P. J. A. (2009). A comparative analysis of phycolloids produced by underutilized versus industrially utilized carrageenophytes (Gigartinales, Rhodophyta). J. Appl. Phycol. 21:599–605.CrossRefGoogle Scholar
Pereira, R. C., and da Gama, B. A. P. (2008). Macroalgal chemical defenses and their roles in structuring tropical marine communities. In Amsler, C. D. (ed.), Algal Chemical Ecology (pp. 25–56). London: Springer (Limited).CrossRefGoogle Scholar
Pereira, R. C., Bianco, E. M., Bueno, L. B., et al. (2010). Associational defense against herbivory between brown seaweeds. Phycologia 49:424–8.CrossRefGoogle Scholar
Pereira, R., and Yarish, C. (2010). The role of Porphyra in sustainable culture systems: physiology and applications. In Israel, A. and Einav, R. (eds), Role of Seaweeds in a Globally Changing Environment (pp. 339–54). New York: Springer. [, , ]CrossRefGoogle Scholar
Pereira, R., Kramer, G., Yarish, C., and Sousa-Pinto, I. (2008). Nitrogen uptake by gametophytes of Porphyra dioica (Bangiales, Rhodophyta) under controlled-culture conditions. Eur. J. Phycol. 43:107–18.CrossRefGoogle Scholar
Pereira, R., Yarish, C., and Critchley, A. (2012). Seaweed aquaculture for human foods, in land-based and IMTA systems. In Meyers, R. A. (ed.). Encyclopedia of Sustainability Science and Technology (pp. 9109–28). New York: Springer Science. [, , ]CrossRefGoogle Scholar
Pereira, T. R., Engelen, A. H., Pearson, G. A., et al. (2011). Temperature effects on the microscopic haploid stage development of Laminaria ochroleuca and Saccorhiza polyschides, kelps with contrasting life histories. Cah. Biol. Mar. 52:395–403.Google Scholar
Perez, R., Durand, P., Kaas, R., et al. (1988). Undaria pinnatifida on the French coasts: cultivation method, biochemical composition of the sporophyte and the gametophyte. In Stadler, T. (ed.), Algal Biotechnology (pp. 315–27). Amsterdam: Elsevier.Google Scholar
Pérez-Rodríguez, E., Aguilera, J., Gómez, I., and Figueroa, F. L. (2001). Excretion of coumarins by the Mediterranean green alga Dasycladus vermicularis in response to environmental stress. Mar. Biol. 139:633–9.Google Scholar
Perrin, C., Daguin, C., van de Vliet, M., et al. (2007). Implications of mating system for genetic diversity of sister algal species: Fucus spiralis and Fucus vesiculosus (Heterokontophyta, Phaeophyceae). Eur. J. Phycol. 42:219–30.CrossRefGoogle Scholar
Perrone, C., and Felicini, G. P. (1976). Les bourgeons adventifs de Gigartina acicularis (Wulf.) Lamour. (Rhodophyta, Gigartinales) en culture. Phycologia 15:45–50.CrossRefGoogle Scholar
Perrot-Rechenmann, C. (2010). Cellular responses to auxin: division versus expansion. Cold Spring Harb. Perspect. Biol. 2:a001446.CrossRefGoogle ScholarPubMed
Peters, A. F., and Clayton, M. N. (1998). Molecular and morphological investigations of three brown algal genera with stellate plastids: evidence for Scytothamnales ord. Nov. (Phaeophyceae). Phycologia 37:106–13.CrossRefGoogle Scholar
Peters, A. F., and Schaffelke, B. (1996). Streblonema (Ectocarpales, Phaeophyceae) infection in the kelp Laminaria saccharina (Laminariales, Phaeophyceae) in the western Baltic. Hydrobiologia 326/327:111–16.CrossRefGoogle Scholar
Peters, A. F., van Oppen, M. J. H., Wiencke, C., et al. (1997). Phylogeny and historical ecology of the Desmarestiaceae (Phaeophyceae) support a southern hemisphere origin. J. Phycol. 33:294–309.CrossRefGoogle Scholar
Peters, A. F., Scornet, D., Müller, D. G., Kloareg, B., and Cock, J. M. (2004a). Inheritance of organelles in artificial hybrids of the isogamous multicellular chromist alga Ectocarpus siliculosus (Phaeophyceae). Eur. J. Phycol. 39:235–42.CrossRefGoogle Scholar
Peters, A. F., Marie, D., Scornet, D., Kloareg, B., and Cock, J. M. (2004b). Proposal of Ectocarpus siliculosus (Ectocarpales, Phaeophyceae) as a model organism for brown algal genetics and genomics. J. Phycol. 40:1079–88.CrossRefGoogle Scholar
Peters, A. F., Scornet, D., Ratin, M., et al. (2008). Life-cycle-generation-specific developmental processes are modified in the immediate upright mutant of the brown alga Ectocarpus siliculosus. Development 135:1503–12.CrossRefGoogle ScholarPubMed
Peterson, B. J., and Fry, B. (1987). Stable isotopes in ecosystem studies. Ann. Rev. Ecol. Syst. 18:293–320.CrossRefGoogle Scholar
Peterson, D. H., Perry, M. J., Bencala, K. E., and Talbot, M. C. (1987). Phytoplankton productivity in relation to light intensity: a simple equation. Estuar. Coast Shelf Sci. 24:813–32.CrossRefGoogle Scholar
Peterson, R. D. (1972). Effects of light intensity on the morphology and productivity of Caulerpa racemosa (Forsskal) 1. Agardh. Micronesica 8:63–86.Google Scholar
Petraitis, P. S., Methratta, E. T., Rhile, E. C., Vidargas, N. A., and Dudgeon, S. R. (2009). Experimental confirmation of multiple community states in a marine ecosystem. Oecologia 161:139–48.CrossRefGoogle Scholar
Petrell, R. J., Mazhari Tabrizi, K., Harrison, P. J., and Druehl, L. D. (1993). Mathematical model of Laminaria production near a British Columbian salmon sea cage farm. J. Appl. Phycol. 5:1–4.CrossRefGoogle Scholar
Pfister, C. A. (2007). Intertidal invertebrates locally enhance primary production. Ecology 88:1647–53.CrossRefGoogle ScholarPubMed
Pfister, C. A., and Hay, M. E. (1988). Associational plant refuges: convergent patterns in marine and terrestrial communities result from differing mechanisms. Oecologia 77:118–29.CrossRefGoogle ScholarPubMed
Philips, J. C., and Hurd, C. L. (2003). Nitrogen ecophysiology of intertidal seaweeds from New Zealand: N uptake, storage and utilization in relation to shore position and season. Mar. Ecol. Prog. Ser. 264:31–40.CrossRefGoogle Scholar
Philips, J. C., and Hurd, C. L. (2004). Kinetics of nitrate, ammonium and urea uptake by four intertidal seaweeds from New Zealand. J. Phycol. 40:534–45.CrossRefGoogle Scholar
Phillips, D. J. H. (1990). Use of macroalgae and invertebrates as monitors of metal levels in estuaries and coastal waters. In Furness, R. W. and Rainbow, P. S. (eds), Heavy Metals in the Marine Environment (pp. 82–99). Boca Raton, FL: CRC Press.Google Scholar
Phillips, D. J. H. (1991). Selected trace elements and the use of biomonitors in subtropical and tropical marine ecosystems. Rev. Environ. Contamin. Toxicol. 120:105–29.CrossRefGoogle Scholar
Phillips, J. A., Clayton, M. N., Maier, I., Boland, W., and Müller, D. G. (1990). Sexual reproduction in Dictyota diemensis (Dictyotales, Phaeophyta). Phycologia 29:367–79.CrossRefGoogle Scholar
Phillips, J. C., Kendrick, G. A., and Lavery, P. S. (1997). A test of a functional group approach to detecting shifts in macroalgal communities along a disturbance gradient. Mar. Ecol. Prog. Ser. 153:125–38.CrossRefGoogle Scholar
Phlips, E. J., and Zeman, C. (1990). Photosynthesis, growth and nitrogen fixation by epiphytic forms of filamentous cyanobacteria from pelagic Sargassum. Bull. Mar. Sci. 47:613–21.Google Scholar
Pianka, E. R. (1966). Latitudinal gradients in species diversity: a review of concepts. Am. Nat. 100:33–46.CrossRefGoogle Scholar
Pickett-Heaps, J. D., and West, J. (1998). Time-lapse video observations on sexual plasmogamy in the red alga Bostrychia. Eur. J. Phycol. 33:43–56.CrossRefGoogle Scholar
Pickett-Heaps, J. D., West, J. A., Wilson, S. M., and McBride, D. L. (2001). Time-lapse videomicroscopy of cell (spore) movement in red algae. Eur. J. Phycol. 36:9–22.CrossRefGoogle Scholar
Pils, B., and Heyl, A. (2009). Unraveling the evolution of cytokinin signaling. Plant Physiol. 151:782–91.CrossRefGoogle ScholarPubMed
Pinto, E., Sigaud-Kutner, T. C. S., Leitão, M. A. S., O. K., et al. (2003). Heavy metal-induced oxidative stress in algae. J. Phycol. 39:1008–18.CrossRefGoogle Scholar
Plastino, E., Ursi, S., and Fujii, M. T. (2003). Color inheritance, pigment characterization, and growth of a rare light green strain of Gracilaria birdiae (Gracilariales, Rhodophyta). Phycol. Res. 51:45–52.Google Scholar
Plettner, I., Steinke, M., and Malin, G. (2005). Ethene (ethylene) production in the marine macroalga Ulva (Enteromorpha) intestinalis L. (Chlorophyta, Ulvophyceae): effect of light-stress and co-production with dimethyl sulphide. Plant Cell Environ. 28:1136–45.CrossRefGoogle Scholar
Pohnert, G., and Boland, W. (2002). The oxylipin chemistry of attraction and defense in brown algae and diatoms. Nat. Prod. Rep. 19:108–22.Google ScholarPubMed
Polis, G. A., Anderson, W. B., and Holt, R. D. (1997). Toward an integration of landscape and food web ecology: the dynamics of spatially subsidized food webs. Ann. Rev. Ecol. Syst. 28:289–316.CrossRefGoogle Scholar
Polle, A. (1996). Mehler reaction: friend or foe in photosynthesis. Bot. Acta 109:84–9.CrossRefGoogle Scholar
Polne-Fuller, M., and Gibor, A. (1984). Development studies in Porphyra. I. Blade differentiation in Porphyra perforata as expressed by morphology, enzymatic digestion, and protoplast regeneration. J. Phycol. 20:609–16.CrossRefGoogle Scholar
Polne-Fuller, M., and Gibor, A. (1987). Tissue culture of seaweeds. In Bird, K. T. and Benson, P. H. (eds), Seaweed Cultivation for Renewable Resources (pp. 219–40). Amsterdam: Elsevier.Google Scholar
Polne-Fuller, M., and Gibor, A. (1990). Development studies in Porphyra (Rhodophyceae). III. Effect of culture conditions on wall regeneration and differentiation of protoplasts. J. Phycol. 26:674–82.CrossRefGoogle Scholar
Polunin, N. V. C. (1988). Efficient uptake of algal production by a single resident herbivorous fish on the reef. J. Exp. Mar. Biol. Ecol. 123:61–76.CrossRefGoogle Scholar
Pomin, V. H. (2010). Structural and functional insights into sulfated galactans: a systematic review. Glycoconjugate Journal 27:1–12.CrossRefGoogle ScholarPubMed
Pomin, V. H., and Mourao, P. A. S. (2008). Structure, biology, evolution, and medical importance of sulfated fucans and galactans. Glycobiology 18:1016–27.CrossRefGoogle ScholarPubMed
Poore, A. G. B., Campbell, A. H., and Steinberg, P. D. (2009). Natural densities of mesograzers fail to limit growth of macroalgae or their epiphytes in a temperate algal bed. J. Ecol. 97:164–75.CrossRefGoogle Scholar
Poore, A. G., Campbell, A. H., Coleman, R. A., et al. (2012). Global patterns in the impact of marine herbivores on benthic primary producers. Ecol. Lett. 15:912–22. [, , ]CrossRefGoogle ScholarPubMed
Porter, E. T., Lawrence, P. S., and Suttles, S. E. (2000). Gypsum dissolution is not a universal integrator of “water motion”. Limnol. Oceanogr. 45:145–58.CrossRefGoogle Scholar
Potin, P. (2008). Oxidative burst and related responses in biotic interactions of algae. In Amsler, C. D. (ed.) Algal Chemical Ecology (pp. 245–72). Berlin and Heidelberg: Springer-Verlag. [, . ]CrossRefGoogle Scholar
Potin, P. (2012). Intimate associations between epiphytes, endophytes, and parasites of seaweeds. In Wiencke, C., and Bischof, K. (eds), Seaweed Biology: Novel Insights into Ecophysiology, Ecology and Utilization, Ecological Studies 219 (pp. 203–34). Berlin and Heidelberg: Springer-Verlag.CrossRefGoogle Scholar
Preston, C. H. (2001). The Exxon Valdez oil spill in Alaska: acute, indirect and chronic effects on the ecosystem. Adv. Mar. Biol. 39:1–101.Google Scholar
Preston, M. R. (1988). Marine pollution. In Riley, J. P. (ed.), Chemical Oceanography (pp. 53–196). Orlando, FL: Academic Press.Google Scholar
Price, I. R. (1989). Seaweed phenology in a tropical Australian locality (Townsville, North Queensland). Bot. Mar. 32:399–406.CrossRefGoogle Scholar
Price, I. R., Fricker, R. L., and Wilkinson, C. R. (1984) Ceratodictyon spongiosum (Rhodophyta), the macroalgal partner in an alga-sponge symbiosis, grown in unialgal culture. J. Phycol. 20:156–8.CrossRefGoogle Scholar
Price, N. M., and Harrison, P. J. (1988). Specific selenium-containing macromolecules in the marine diatom Thalassiosira pseudonana. Plant Physiol. 86:192–9.CrossRefGoogle ScholarPubMed
Prince, E. K., and Pohnert, G. (2010). Searching for signals in the noise: metabolomics in chemical ecology. Anal. Bioanal. Chem. 396:193–7.CrossRefGoogle ScholarPubMed
Prince, J. S., and Trowbridge, C. D. (2004). Reproduction in the green macroalga Codium (Chlorophyta): characterization of gametes. Bot. Mar. 47:461–70.CrossRefGoogle Scholar
Probyn, T. A. (1984). Nitrate uptake by Chordaria flagelliformis (Phaeophyta). Bot. Mar. 17:271–5.Google Scholar
Probyn, T. A., and Chapman, A. R. O. (1982). Nitrogen uptake characteristics of Chordaria flagelliformis (Phaeophyta) in batch mode and continuous mode experiments. Mar. Biol. 71:129–33. [, , ]CrossRefGoogle Scholar
Probyn, T. A., and Chapman, A. R. O. (1983). Summer growth of Chordaria flagelliformis (0. F. Muell.) C. Ag.: physiological strategies in a nutrient stressed environment. J. Exp. Mar. Biol. Ecol. 73:243–71.CrossRefGoogle Scholar
Probyn, T. A., and McQuaid, C. D. (1985). In situ measurements of nitrogenous nutrient uptake by kelp (Ecklonia maxima) and phytoplankton in a nitrate-rich upwelling environment. Mar. Biol. 88:149–54.CrossRefGoogle Scholar
Provasoli, L., and Pintner, I. J. (1980). Bacteria induced polymorphism in an axenic laboratory strain of Ulva lactuca (Chlorophyceae). J. Phycol. 16:196–201.CrossRefGoogle Scholar
Pu, R., and Robinson, K. R. (2003). The involvement of Ca2+ gradients, Ca2+ fluxes, and CaM kinase II in polarization and germination of Silvetia compressa zygotes. Planta 217:407–16.CrossRefGoogle Scholar
Pueschel, C. M. (1989). An expanded survey of the ultrastructure of red algal pit plugs. J. Phycol. 25:625–36.CrossRefGoogle Scholar
Pueschel, C. M. (1990). Cell structure. In Cole, K. M. and Sheath, R. G. (eds), Biology of the Red Algae (pp. 7–41). Cambridge: Cambridge University Press.Google Scholar
Pueschel, C. M., and Cole, K. M. (1982). Rhodophycean pit plugs: an ultrastructural survey with taxonomic implications. Am. J. Bot. 69:703–20.CrossRefGoogle Scholar
Pueschel, C. M., and Korb, R. E. (2001). Storage of nitrogen in the form of protein bodies in the kelp Laminaria solidungula. Mar. Ecol. Prog Ser. 218:107–14.CrossRefGoogle Scholar
Purton, S. (2002). Algal chloroplasts. In: eLS (pp. 1–9). Chichester: John Wiley and Sons Ltd.Google Scholar
Quillet, M., and de Lestang-Bremond, G. (1981). The MeCDPS, a carrying sulphate’s nucleotide of the red seaweed Catenella opuntia (Grev.)Proc. Intl Seaweed Symp. 10:503–7.Google Scholar
Raffaelli, D. G., and Hawkins, S. J. (1996). Intertidal Ecology. Dordrecht, The Netherlands: Kluwer Academic Pubs, 356 pp. [3, , , , , , ]CrossRefGoogle Scholar
Ragan, M. A., and Glombitza, K.-W. (1986). Phlorotannins, brown algal polyphenols. Prog. Phycol. Res. 4:129–241.Google Scholar
Raghothama, K. G. (1999). Phosphate acquisition. Ann. Rev. Plant Physiol. Plant Mol. Biol. 50:655–93.CrossRefGoogle ScholarPubMed
Raikar, V., and Wafar, M. (2006). Surge ammonium uptake in macroalgae from a coral atoll. J. Exp. Mar. Biol. Ecol. 339:236–40.CrossRefGoogle Scholar
Raimondi, P. T., Reed, D. C., Gaylord, B., and Washburn, L. (2004). Effects of self-fertilization in the giant kelp Macrocystis pyrifera. Ecology 85:3267–76.CrossRefGoogle Scholar
Raimonet, M., Guillon, G., Mornet, F., and Richard, P. (2013). Macroalgal δ15N values in well-mixed estuaries: indicator of anthropogenic nitrogen input or macroalgal metabolism? Estu. Coastal Shelf Sci. 119:126–38.CrossRefGoogle Scholar
Rainbow, P. S. (1995). Biomonitoring of heavy metal availability in the marine environment. Mar. Poll. Bull. 31:183–92.CrossRefGoogle Scholar
Rainbow, P. S., and Phillips, D. J. H. (1993). Cosmopolitan biomonitors of trace metals. Mar. Poll. Bull. 26:593–601.CrossRefGoogle Scholar
Ramirez, M. E., Müller, D. G., and Peters, A. F. (1986). Life history and taxonomy of two populations of ligulate Desmarestia (Phaeophyceae) from Chile. Can. J. Bot. 64:2948–54.Google Scholar
Ramon, E. (1973). Germination and attachment of zygotes of Himanthalia elongata (L.) S. F. Gray. J. Phycol. 9:445–9.Google Scholar
Ramus, J. (1978). Seaweed anatomy and photosynthetic performance: the ecological significance of light guides, heterogeneous absorption and multiple scatter. J. Phycol. 14: 352–62.CrossRefGoogle Scholar
Ramus, J. (1981). The capture and transduction of light energy. In Lobban, C. S. and Wynne, M. J. (eds) The Biology of Seaweeds (pp. 458–92). Oxford: Blackwell Scientific. [, , , ]Google Scholar
Ramus, J. (1982). Engelmann’s theory: the compelling logic. In Srivastava, L. M. (ed.), Synthetic and Degradative Processes in Marine Macrophytes (pp. 29–46). Berlin: Walter de Gruyter.Google Scholar
Ramus, J. S. (1990). A form-function analysis of photon capture for seaweeds. Proc. Intl Seaweed Symp. 13:65–71.Google Scholar
Ramus, J., and Rosenberg, G. (1980). Diurnal photosynthetic performance of seaweeds measured under natural conditions. Mar. Biol. 56:21–8.CrossRefGoogle Scholar
Ramus, J., and Venable, M. (1987). Temporal ammonium patchiness and growth rate in Codium and Ulva (Ulvophyceae). J. Phycol. 23:518–23.CrossRefGoogle Scholar
Rasher, D. B., and Hay, M. E. (2010). Chemically rich seaweeds poison corals when not controlled by herbivores. PNAS 107:9683–8.CrossRefGoogle Scholar
Rasher, D., Engel, S., Bonito, V., et al. (2012). Effects of herbivory, nutrients, and reef protection on algal proliferation and coral growth on a tropical reef. Oecologia 169:187–98.CrossRefGoogle ScholarPubMed
Rausch, C., and Bucher, M. (2002). Molecular mechanisms of phosphate transport in plants. Planta 216:23–37.CrossRefGoogle ScholarPubMed
Rautenberger, R., and Bischof, K. (2006). Impact of temperature on UV susceptibility of two species of Ulva (Chlorophyta) from Antarctic and Subantarctic regions. Polar Biol. 29:988–96.CrossRefGoogle Scholar
Raven, J. A. (1984). Energetics and Transport in Aquatic Plants (MBL Lectures in Biology, Vol. 4). New York: Alan R. Liss.Google Scholar
Raven, J. A. (1991). Implications of inorganic carbon utilization: ecology, evolution, and geochemistry. Can. J. Bot. 69:908–24.CrossRefGoogle Scholar
Raven, J. A. (1996). Into the voids: the distribution, function, development and maintenance of gas spaces in plants. Ann. Bot.-London. 78:137–42.CrossRefGoogle Scholar
Raven, J. A. (1997a). Miniview: multiple origins of plasmodesmata. Eur. J. Phycol. 32:95–101.CrossRefGoogle Scholar
Raven, J. A. (1997b). Putting the C in phycology. Eur. J. Phycol. 32:319–33.CrossRefGoogle Scholar
Raven, J. A. (2003). Long-distance transport in non-vascular plants. Plant Cell Envir. 26:73–85. [, , ]CrossRefGoogle Scholar
Raven, J. A. (2010). Inorganic carbon acquisition by eukaryotic algae: four current questions. Photosynth. Res. 106:123–34. [, , ]CrossRefGoogle ScholarPubMed
Raven, J. A. (2011). The cost of photoinhibition. Physiol. Plant 142:87–104.CrossRefGoogle ScholarPubMed
Raven, J. A., and Beardall, J. (1981). Respiration and photorespiration. Can. Bull. Fish Aquat. Sci. 210:55–82.Google Scholar
Raven, J. A., and Geider, R. J. (1988). Temperature and algal growth. New Phytol. 110:441–61. [, , ]CrossRefGoogle Scholar
Raven, J. A., and Geider, R. J. (2003). Adaptation, acclimation and regulation in algal photosynthesis. In Larkum, A. W. D., Douglas, S. E., and Raven, J. A. (eds), Photosynthesis in Algae. Advances in Photosynthesis and Respiration (Volume 14, pp. 385–412). Dordrecht, The Netherlands: Kluwer Academic Publishers.CrossRefGoogle Scholar
Raven, J. A., and Hurd, C. L. (2012). Ecophysiology of photosynthesis in macroalgae. Photosynth. Res. 113:105–25.CrossRefGoogle ScholarPubMed
Raven, J. A., and Lucas, W. J. (1985). The energetics of carbon acquisition. In Lucas, W. J. and Berry, J. A. (eds) Inorganic Carbon Uptake by Aquatic Photosynthetic Organisms (pp 305–24). Rockville MD: The American Society of Plant Physiologists.Google Scholar
Raven, J. A., and Taylor, R. (2003). Macroalgal growth in nutrient-enriched estuaries: a biochemical and evolutionary perspective. Water, Air Soil Poll. 3:7–26.CrossRefGoogle Scholar
Raven, J. A., Johnston, A. M., and MacFarlane, J. J. (1990). Carbon metabolism. In Cole, K. M. and Sheath, R. G. (eds) Biology of the Red Algae (pp. 171–202). Cambridge: Cambridge University Press.Google Scholar
Raven, J. A., Kübler, J. E., and Beardall, J. (2000). Put out the light, and then put out the light. J. Mar. Ass. UK 80:1–25.CrossRefGoogle Scholar
Raven, J. A., Johnston, A. M., Kübler, J. E., et al. (2002a). Mechanistic interpretation of carbon isotope discrimination by marine macroalgae and seagrass. Funct. Plant Biol. 29:355–78.CrossRefGoogle Scholar
Raven, J. A., Johnston, A. M., Kübler, J. E., et al. (2002b). Seaweeds in cold seas: evolution and carbon acquisition. Ann. Bot. 90:525–36.CrossRefGoogle ScholarPubMed
Raven, J. A., Giordano, M., and Beardall, J. (2008). Insights into the evolution of CCMs from comparisons with other resource acquisition and assimilation processes. Physiol. Plant. 133:4–14.CrossRefGoogle ScholarPubMed
Raven, J.A., Beardall, J., Giordano, M., and Maberly, S. C. (2012). Algal evolution in relation to atmospheric CO2: carboxylases, carbon concentrating mechanisms and carbon oxidation cycles. Phil. Trans. Roy. Soc. London B 367:493–507.CrossRefGoogle ScholarPubMed
Raven, P. H., Evert, R. F., and Eichhorn, S. E. (2005). Biology of Plants (7th edn) Madison, NY: Freeman and Co. Publ., 686 pp. [, , ]Google Scholar
Rayko, E., Maumus, F., Maheswari, U., Jabbari, K., and Bowler, C. (2010). Transcription factor families inferred from genome sequences of photosynthetic stramenopiles. New Phytol. 188:52–66.CrossRefGoogle ScholarPubMed
Reddy, C. R. K., Jha, B., Fujita, Y., and Ohno, M. (2008a). Seaweed protoplasts and their potentials: an overview. J. Appl. Phycol. 20:609–17. [, , ]CrossRefGoogle Scholar
Reddy, C. R. K., Gupta, M. K., Mantri, V. A., and Jha, B. (2008b). Seaweed protoplasts: status, biotechnological perspectives and needs. J. Appl. Phycol. 20:619–52.CrossRefGoogle Scholar
Reed, D. C. (1990a). The effects of variable settlement and early competition on patterns of kelp recruitment. Ecology 71:776–87. [, , ]CrossRefGoogle Scholar
Reed, D. C. (1990b). An experimental evaluation of density dependence in a subtidal algal population. Ecology 71:2286–96.CrossRefGoogle Scholar
Reed, D. C., Laur, D. R., and Ebeling, A. W. (1988). Variation in algal dispersal and recruitment: the importance of episodic events. Ecol. Monogr. 58:321–35.CrossRefGoogle Scholar
Reed, D. C., Amsler, C. D., and Ebeling, A. W. (1992). Dispersal in kelps: factors affecting spore swimming and competency. Ecology 73:1577–85.CrossRefGoogle Scholar
Reed, D. C., Anderson, T. W., Ebeling, A. W., and Anghera, M. (1997). The role of reproductive synchrony in the colonization potential of kelp. Ecology 78:2443–57.CrossRefGoogle Scholar
Reed, D. C., Brzezinski, M. A., Coury, D. A., Graham, W. M., and Petty, R. L. (1999). Neutral lipids in macroalgal spores and their role in swimming. Mar. Biol. 133:737–44.CrossRefGoogle Scholar
Reed, D. C., Kinlan, B. P., Raimondi, P. T., et al. (2006). A metapopulation perspective on the patch dynamics of giant kelp in Southern California. In Kritzer, J., and Sale, P. (eds), Marine Metapopulations (pp. 353–86) Burlington, MA:Elsevier Academic Press.CrossRefGoogle Scholar
Reed, R. H. (1990c). Solute accumulation and osmotic adjustment. In Cole, K. M. and Sheath, R. G. (eds), Biology of Red Algae (pp. 147–70). Cambridge: Cambridge University Press. [, , ]Google Scholar
Reed, R. H., and Collins, J. C. (1980). The ionic relations of Porphyra purpurea (Roth) C. Ag. (Rhodophyta, Bangiales). Plant Cell Environ. 3:399–407.Google Scholar
Reed, R. H., and Moffat, L. (1983). Copper toxicity and copper tolerance in Enteromorpha compressa (L.) Grev. J. Exp. Mar. Biol. Ecol. 63:85–103.CrossRefGoogle Scholar
Reed, R. H., Collins, J. C., and Russell, G. (l980). The effects of salinity upon galactosyl-glycerol content and concentration of the marine red alga Porphyra purpurea (Roth) C. Ag. J. Exp. Bot. 31:1539–54.CrossRefGoogle Scholar
Reed, R. H., Davison, I. R., Chudek, J. A., and Foster, R. (1985). The osmotic role of mannitol in the Phaeophyta: an appraisal. Phycologia 24:35–47.CrossRefGoogle Scholar
Rees, D. A. (1975). Stereochemistry and binding behaviour of carbohydrate chains. In Whelan, W. J. (ed.) Biochemistry of Carbohydrates (pp. 1–42). London: Butterworth.Google Scholar
Rees, T. A. V. (2003). Safety factors and nutrient uptake by seaweeds. Mar. Ecol. Prog. Ser. 263:29–40.CrossRefGoogle Scholar
Rees, T. A. V. (2007). Metabolic and ecological constraints imposed by similar rates of ammonium and nitrate uptake per unit surface area at low substrate concentrations in marine phytoplankton and macroalgae. J. Phycol. 43:197–207.Google Scholar
Rees, T. A. V., Grant, C. M., Harmens, H. E., and Taylor, R. B. (1998). Measuring rates of ammonium assimilation in marine algae: use of the protonophore carbonyl cyanide m-chlorophenylhydrazone to distinguish between uptake and assimilation. J. Phycol. 34:264–72.CrossRefGoogle Scholar
Rees, T. A. V., Dobson, B. C., Bijl, M., and Morelissen, B. (2007). Kinetics of nitrate uptake by New Zealand marine macroalgae and evidence for two nitrate transporters in Ulva intestinalis L. Hydrobiol. 586:135–41.CrossRefGoogle Scholar
Reise, K. (1983). Sewage, green algal mats anchored by lugworms, and the effects on Turbellaria and small Polychaeta. Helgol. Meeresunters. 36:151–62.CrossRefGoogle Scholar
Reiskind, J. B., Seamon, P. T., and Bowes, G. (1988). Alternative methods of photosynthetic carbon assimilation in marine macroalgae. Plant Physiol. 87: 686–92.CrossRefGoogle ScholarPubMed
Reiskind, J. B., Beer, S., and Bowes, G. (1989). Photosynthesis, photorespiration and ecophysiological interaction in marine macroalgae. Aquat. Bot. 34:131–52.CrossRefGoogle Scholar
Reith, M., and Munholland, J. (1995). Complete nucleotide sequence of the Porphyra purpurea choloroplast genome. Plant Molec. Biol. Reporter 13:333–5.CrossRefGoogle Scholar
Ren, G.-Z., Wang, J.-C., and Chen, M.-Q. (1984). Cultivation of Gracilaria by means of low rafts. Hydrobiologia 116/117:72–6.Google Scholar
Rensing, L., and Ruoff, P. (2002). Temperature effect on entrainment, phase shifting, and amplitude of circadian clocks and its molecular bases. Chronobiol. Int. 19:807–64.CrossRefGoogle ScholarPubMed
Revsbech, N. P. (1989). An oxygen microelectrode with a guard cathode. Limnol. Oceanogr. 55:1907–10.Google Scholar
Reyes-Prieto, A., Weber, A. P. M., and Bhattacharya, D. (2007). The origin and establishment of the plastid in algae and plants. Annu. Rev. Genet. 41:147–68.CrossRefGoogle ScholarPubMed
Rhoades, D. F. (1979). Evolution of plant chemical defense against herbivores. In Rosenthal, G. A. and Janzen, D. H. (eds), Herbivores: Their Interaction with Secondary Plant Metabolites (pp. 3–54). New York: Academic Press.Google Scholar
Richoux, N. B., and Froneman, P. W. (2008). Trophic ecology of dominant zooplankton and macrofauna in a temperate, oligotrophic South African estuary: a fatty acid approach. Mar. Ecol. Prog. Ser. 357:121–37.CrossRefGoogle Scholar
Rico, J. M., and Guiry, M. D. (1996). Phototropism in seaweeds: a review. Sci. Mar. 60:273–81.Google Scholar
Ridler, N., Wowchuk, M., Robinson, B., et al. (2007) Integrated multi-trophic aquaculture (IMTA): a potential strategic choice for farmers. Aquacult. Econ. Management 11:99–110.CrossRefGoogle Scholar
Ries, J. B. (2010). Review: geological and experimental evidence for secular variation in seawater Mg/Ca (calcite-aragonite seas) and its effects on marine biological calcification. Biogeosciences 7:2795–848.CrossRefGoogle Scholar
Rietema, H. (1982). Effects of photoperiod and temperature on macrothallus initiation in Dumontia contorta (Rhodophyta). Mar. Ecol. Prog. Ser. 8:187–96.CrossRefGoogle Scholar
Rietema, H. (1984). Development of erect thalli from basal crusts in Dumontia contorta (Gmel.) Rupr. (Rhodophyta, Cryptonemiales). Bot. Mar. 27:29–36.CrossRefGoogle Scholar
Rietema, H., and Breeman, A. M. (1982). The regulation of the life history of Dumontia contorta in comparison to that of several other Dumontiaceae (Rhodophyta). Bot. Mar. 25:569–76.CrossRefGoogle Scholar
Riquelme, C., Rojas, A., Flores, V., and Correa, J. A. (1997). Epiphytic bacteria in a copper-enriched environment in northern Chile. Mar. Poll. Bull. 34:816–20.CrossRefGoogle Scholar
Risk, M. J., Lapointe, B. E., Sherwood, O. A., and Bedford, B. J. (2009). The use of δ15N in assessing sewage stress on corals. Mar. Poll. Bull. 58:793–802.CrossRefGoogle Scholar
Ritchie, R. J., and Larkum, A. W. D. (l987). The ionic relations of small-celled marine algae. Prog. Phycol. Res. 5:179–222.Google Scholar
Ritter, A., Goulitquer, S., Salaün, J. P., et al. (2008). Copper stress induces biosynthesis of octadecanoid and eicosanoid oxygenated derivatives in the brown algal kelpLaminaria digitata. New Phytol. 180:809–21.CrossRefGoogle ScholarPubMed
Rivera, M., and Scrosati, R. (2006). Population dynamics of Sargassum lapazeanum (Fucales, Phaeophyta) from the Gulf of California, Mexico. Phycologia 45:178–89.CrossRefGoogle Scholar
Rivera, M., and Scrosati, R. (2008). Self-thinning and size inequality dynamics in a clonal seaweed (Sargassum lapazeanum, Phaeophyceae). J. Phycol. 44:45–9CrossRefGoogle Scholar
Roberson, L. M., and Coyer, J. A. (2004). Variation in blade morphology of the kelp Eisenia arborea: incipient speciation due to local water motion? Mar. Ecol. Prog. Ser. 282:115–28. [, , ]CrossRefGoogle Scholar
Roberts, D. A., Poore, A. G. B., and Johnston, E. L. (2006). Ecological consequences of copper contamination in macroalgae: effects on epifauna and associated herbivores. Envir. Tox. Chem. 25:2470–9.CrossRefGoogle ScholarPubMed
Roberts, E., and Roberts, A. W. (2009). A cellulose synthase (CESA) gene from the red alga Porphyra yezoensis (Rhodophyta). J. Phycol. 45:203–12.CrossRefGoogle Scholar
Roberts, M., and Ring, F. M. (1972). Preliminary investigations into conditions affecting the growth of the microscopic phase of Scytosiphon lomentarius (Lyngbye) Link. Mem. Soc. Bot. Fr. 1972:117–28.CrossRefGoogle Scholar
Roberts, S. K., Gillot, I., and Brownlee, C. (1994). Cytoplasmic calcium and Fucus egg activation. Development 120:155–63.Google Scholar
Robertson, A. I., and Lucas, J. J. (1983). Food choice, feeding rates, and the turnover of macrophyte biomass by a surf-zone inhabiting amphipod. J. Exp. Mar. Biol. Ecol. 72:99–124.CrossRefGoogle Scholar
Robles, C., and Desharnais, R. (2002). History and current development of a paradigm of predation in rocky intertidal communities. Ecology 83:1521–36.CrossRefGoogle Scholar
Roenneberg, T., and Mittag, M. (1996). The circadian program of algae. Cell Dev. Biol. 7:753–63.CrossRefGoogle Scholar
Rogers, K. M. (2003). Stable carbon and nitrogen isotope signatures indicate recovery of marine biota from pollution at Moa Point, New Zealand. Mar. Poll. Bull. 46:821–7.CrossRefGoogle ScholarPubMed
Rohde, S., and Wahl, M. (2008). Antifeeding defense in Baltic macroalgae: induction by direct grazing versus waterborne cues. J. Phycol. 44:85–90.CrossRefGoogle ScholarPubMed
Rohde, S., Molis, M., and Wahl, M. (2004). Regulation of anti-herbivore defence by Fucus vesiculosus in response to various cues. J. Ecol. 92:1011–18.CrossRefGoogle Scholar
Rohde, S., Hebenthal, C., Wahl, M., Karez, R.,and Bischoff, K. (2008). Decreased depth distribution of Fucus vesiculosus (Phaeophyceae) in the Western Baltic: effects of light deficiency and epibionts on growth and photosynthesis. Eur. J. Phycol. 43:143–50.CrossRefGoogle Scholar
Roleda, M. Y., and Dethleff, D. (2011). Storm-generated sediment deposition on rocky shores: simulating burial effects on the physiology and morphology of Saccharina latissima sporophytes. Mar. Biol. Res. 7:213–23.CrossRefGoogle Scholar
Roleda, M., Wiencke, C., Hanelt, D., Van de Poll, W., and Gruber, A. (2005). Sensitivity of Laminariales zoospores from Helgoland (North Sea) to ultraviolet and photosynthetically active radiation: implications for depth distribution and seasonal reproduction. Plant Cell Envir. 28: 466–79.CrossRefGoogle Scholar
Rosell, K.-G., and Strivastava, L. M. (1985). Seasonal variations in total nitrogen, carbon and amino acids in Macrocystis integrifola and Nereocystis luetkeana (Phaeophyta). J. Phycol. 21:304–9.CrossRefGoogle Scholar
Rosenberg, G., and Paerl, H. W. (1981). Nitrogen fixation by blue-green algae associated with the siphonous green seaweed Codium decorticatum: effects on ammonium uptake. Mar. Biol. 61:151–8.CrossRefGoogle Scholar
Rosenberg, G., and Ramus, J. (1982). Ecological growth strategies in the seaweed Gracilaria foliifera (Rhodophyceae) and Ulva sp. (Chlorophyceae): soluble nitrogen and reserve carbohydrates. Mar. Biol. 66:251–9. [, , ]CrossRefGoogle Scholar
Rosenberg, G., and Ramus, J. (1984). Uptake of inorganic nitrogen and seaweed surface area: volume ratios. Aquat. Bot. 19:65–72.CrossRefGoogle Scholar
Rosenberg, G., Probyn, T. A., and Mann, K. H. (1984). Nutrient uptake and growth kinetics in brown seaweeds: response to continuous and single additions of ammonium. J. Exp. Mar. Biol. Ecol. 80:125–46.CrossRefGoogle Scholar
Rosenberg, G., Littler, D. S., Littler, M. M., and Oliveira, E. C. (1995). Primary production and photosynthetic quotients of seaweeds from Sao Paulo State, Brazil. Bot. Mar. 38:369–77.CrossRefGoogle Scholar
Rosman, J. H., Monismith, S. G., Denny, M. W., and Koseff, J. R. (2010). Currents and turbulence within a kelp forest (Macrocystis pyrifera): insights from a dynamically scaled laboratory model. Limnol. Oceanogr. 55:1145–58.CrossRefGoogle Scholar
Ross, C., and Van Alstyne, K. L. (2007). Intraspecific variation in stress-induced hydrogen peroxide scavenging by the ulvoid macroalga Ulva lactuca. J. Phycol. 43:466–74.CrossRefGoogle Scholar
Ross, C., Vreeland, V., Waite, J. H., and Jacobs, R. S. (2005a). Rapid assembly of a wound plug: stage one of a two-stage wound repair mechanism in the giant unicellular chlorophyte Dasycladus vermicularis (Chlorophyceae). J. Phycol. 41:46–54.CrossRefGoogle Scholar
Ross, C., Küpper, F. C., Vreeland, V., Waite, J. H., and Jacobs, R. S. (2005b). Evidence of a latent oxidative burst in relation to wound repair in the giant unicellular chlorophyte Dasycladus vermicularis. J. Phycol. 41:531–41.CrossRefGoogle Scholar
Ross, C., Küpper, F. C., and Jacobs, R. S. (2006). Involvement of reactive oxygen species and reactive nitrogen species in the wound response of Dasycladus vermicularis. Chem. Biol. 13:353–64.CrossRefGoogle ScholarPubMed
Rossi, F., Olabarria, C., Incera, M., and Garido, J. (2010). The trophic significance of the invasive seaweed Sargassum muticum in sandy beaches. J. Sea Res. 63:52–61.CrossRefGoogle Scholar
Rothäusler, E., Gómez, I., Hinojosa, I. A., et al. (2009). Effect of temperature and grazing on growth and reproduction of floating Macrocystis spp. (Phaeophyceae) along a latitudinal gradient. J. Phycol. 45:547–59.CrossRefGoogle ScholarPubMed
Rothäusler, E., Gutow, L., and Thiel, M. (2012). Floating seaweeds and their communities. In Wiencke, C., and Bischof, K. (eds), Seaweed Biology: Novel Insights Into Ecophysiology, Ecology and Utilization (Ecological Studies, Volume 219, pp. 359–80). Berlin and Heidelberg: Springer.CrossRefGoogle Scholar
Röttgers, R. (2007). Comparison of different variable chlorophyll a fluorescence techniques to determine photosynthetic parameters of natural phytoplankton. Deep Sea Res. I 54:437–51.CrossRefGoogle Scholar
Roughgarden, J., Gaines, S. D., and Possingham, H. (1988). Recruitment dynamics in complex life cycles. Science 241:1460–6.CrossRefGoogle ScholarPubMed
Rovilla, M., Luhan, J., and Sollesta, H. (2010). Growing the reproductive cells (carpospores) of the seaweed, Kappaphycus striatum, in the laboratory until outplanting in the field and maturation to tetrasporophyte. J. Appl. Phycol. 22:579–85.Google Scholar
Rowan, K. S. (1989). Photosynthetic Pigments of Algae. Cambridge: Cambridge University Press.Google Scholar
Rüdiger, W., and López-Figueroa, F. (1992). Photoreceptors in algae. Photochem. Photobiol. 55:949–54.CrossRefGoogle Scholar
Rueness, J. (1973). Pollution effects on littoral algal communities in the inner Oslofjord, with special reference to Ascophyllum nodosum. Helgol. Meeresunters. 24:446–54.CrossRefGoogle Scholar
Ruesink, J. L. (1998). Diatom epiphytes on Odonthalia floccosa: the importance of extent and timing. J. Phycol. 34:29–38.CrossRefGoogle Scholar
Rugg, D. A., and Norton, T. A. (1987). Pelvetia canaliculata, a high shore seaweed that shuns the sea. In Crawford, R. M. M. (ed.), Plant Life in Aquatic and Amphibious Habitats (pp. 347–58). Oxford: Blackwell Scientific.Google Scholar
Rui, F., and Boland, W. (2010). Algal pheromone biosynthesis: stereochemical analysis and mechanistic implications in gametes of Ectocarpus siliculosus. J. Org. Chem. 75:3958–64.CrossRefGoogle ScholarPubMed
Ruiz, D. J., and Wolff, M. (2011). The Bolivar Channel Ecosystem of the Galapagos Marine Reserve: Energy flow structure and role of keystone groups. J. Sea Res. 66: 123–34.CrossRefGoogle Scholar
Rumpho, M. E., Dastoor, F. P., Manhart, J. R., and Lee, J. (2006). The kleptoplast. In Wise, R. R., and Hoober, J. K. (eds), The Structure and Function of Plastids (pp. 451–73).Series: Advances in Photosynthesis and Respiration, Vol. 23. Springer.CrossRefGoogle Scholar
Rumpho, M. E., Worful, J. M., Lee, J., et al. (2008). Horizontal gene transfer of the algal nuclear gene psbO to the photosynthetic sea slug Elysia chlorotica. PNAS 105:17,867–71.CrossRefGoogle ScholarPubMed
Rumpho, M. E., Pelletreau, K. N., Moustafa, A., and Bhattacharya, D. (2011). The making of a photosynthetic animal. J. Exp. Biol. 214:303–11.CrossRefGoogle ScholarPubMed
Runcie, J. W., and Larkum, A. W. (2001). Estimating internal phosphorus pools in macroalgae using radioactive phosphorus and trichloroacetic acid extracts. Anal. Biochem. 297:191–2.CrossRefGoogle ScholarPubMed
Runcie, J. W., Ritchie, R. J., and Larkum, A. D. W. (2004). Uptake kinetics and assimilation of phosphorus by Catenella nipae and Ulva lactuca can be used to indicate ambient phosphate availability. J. Appl. Phycol. 16:181–94.CrossRefGoogle Scholar
Runcie, J. W., Gurgel, C. F. D., and McDermid, K. J. (2008) In situ photosynthetic rates of tropical marine macroalgae at their lower depth limit. Eur. J. Phycol. 43:377–88.CrossRefGoogle Scholar
Ruperez, P., Ahrazem, O., and Leal, J. A. (2002). Potential antioxidant capacity of sulfated polysaccharides from the edible marine brown seaweedFucus vesiculosus. J. Agric. Food Chem. 50:840–5.CrossRefGoogle ScholarPubMed
Rusig, A.-M., Le Guyader, H., and Ducreux, G. (1994). Dedifferentiation and microtubule reorganization in the apical cell protoplast of Sphacelaria (Phaeophyceae). Protoplasma 179:83–94.CrossRefGoogle Scholar
Russell, B. D., and Connell, S. D. (2007). Response of grazers to sudden nutrient pulses in oligotrophic versus eutrophic conditions. Mar. Ecol Prog. Ser. 349:73–80.CrossRefGoogle Scholar
Russell, G. (1983). Formation of an ectocarpoid epiflora on blades of Laminaria digitata. Mar. Ecol. Prog. Ser. 11:1817.CrossRefGoogle Scholar
Russell, G. (1986). Variation and natural selection in marine macroalgae. Oceanogr. Mar. Biol. Annu. Rev. 24:309–77.Google Scholar
Russell, G. (1987). Salinity and seaweed vegetation. In Crawford, R. M. M. (ed.), Plant Life in Aquatic and Amphibious Habitats (pp. 35–52). Oxford: Blackwell Scientific.Google Scholar
Russell, G. (1988). The seaweed flora of a young semienclosed sea: the Baltic. Salinity as a possible agent of flora divergence. Helgol Meeresunters 42: 243–50.CrossRefGoogle Scholar
Russell, G. (1991). Vertical distribution. In Mathieson, A. C., and Nienhuis, P. H. (eds), Ecosystems of the World. 24. Intertidal and Littoral Ecosystems (pp. 43–65). New York: Elsevier.Google Scholar
Russell, G., and Bolton, J. J. (1975). Euryhaline ecotypes of Ectocarpus siliculosus (Dillw.)Lyngb. Estu. Cstl. Mar. Sci. 3:91–4.CrossRefGoogle Scholar
Russell, G., and Fielding, A. H. (1974). The competitive properties of marine algae in culture. J. Ecol. 62:689–98.CrossRefGoogle Scholar
Russell, G., and Fielding, A. H. (1981). Individuals, populations and communities. In Lobban, C. S. and Wynne, M. J. (eds), The Biology of Seaweeds (pp. 393–420). Oxford: Blackwell Scientific.Google Scholar
Russell, G., and Veltkamp, C. J. (1984). Epiphyte survival on skin-shedding macrophytes. Mar. Ecol. Prog. Ser. 18:149–53.CrossRefGoogle Scholar
Russell, L. K., Hepburn, C.D., Hurd, C.L., and Stuart, M.D. (2008). The expanding range of Undaria pinnatifida in southern New Zealand: distribution, dispersal mechanisms and the invasion of wave-exposed environments. Biol. Inv. 10:103–15.CrossRefGoogle Scholar
Russell-Hunter, W. D. (1970). Aquatic Productivity. An Introduction to Some Basic Aspects of Biological Oceanography and Limnology. New York: Macmillan.Google Scholar
Ryther, J. H., Goldman, J. C., Gifford, C. E., et al. (1979). Physical models of integrated waste recycling marine polyculture systems. Aquaculture 5:163–77.CrossRefGoogle Scholar
Sacramento, A. T., Garcia-Jiménez, P., Alcázar, R., Tiburcio, A. F., and Robaina, R. R. (2004). Influence of polymines on the sporulation of Grateloupia (Halymeniaceae, Rhodophyta). J. Phycol. 40:887–94.CrossRefGoogle Scholar
Sacramento, A. T., Garcia-Jiménez, P., and Robaina, R. R. (2007). The polyamine spermine induces cystocarp development in the seaweed Grateloupia (Rhodophyta). Plant Growth Regul. 53:147–54.CrossRefGoogle Scholar
Saffo, M. B. (1987). New light on seaweeds. BioSci. 37:654–64.CrossRefGoogle Scholar
Saga, N., Uchida, T., and Sakai, Y. (1978). Clone Laminaria from single isolated cell. Bull. Jpn. Soc. Sci. Fish. 44:87.CrossRefGoogle Scholar
Sagert, S., and Schubert, H. (1995). Acclimation of the photosynthetic aparatus of Palmaria palmata (Rhodophyta) to light qualities that preferentially excite photosystem I or PS II. J. Phycol. 31:547–54.CrossRefGoogle Scholar
Sahoo, D., and Yarish, C. (2005). Mariculture of seaweeds. In Andersen, R. (ed.), Algal Culturing Techniques (pp. 219–37). New York: Academic Press. [, , ]Google Scholar
Saito, A., Mizuta, H., Yasui, H., and Saga, N. (2008). Artificial production of regenerable free cells in the gametophyte of Porphyra pseudolinearis (Bangiales, Rhodophyceae). Aquaculture 281:138–44.CrossRefGoogle Scholar
Salgado, L. T., Viana, N. B., Andrade, L. R., et al. (2008). Intra-cellular storage, transport and exocytosis of halogenated compounds in marine red alga Laurencia obtusa. J. Struct. Biol. 162:345–55.CrossRefGoogle ScholarPubMed
Salisbury, J. L. (2007). A mechanistic view on the evolutionary origin for centrin-based control of centriole duplication. J. Cell. Physiol. 213:420–8.CrossRefGoogle ScholarPubMed
Salles, S., Aguilera, J., and Figueroa, F. L. (1996). Light fields in algal canopies: Changes in spectral light ratios and growth of Porphyra leucosticta Thur. Le Jol. Sci. Mar. 60:29–38.Google Scholar
Salvucci, M. E. (1989). Regulation of Rubisco activity in vivo. Physiol. Plant 77:164–71.CrossRefGoogle Scholar
Sánchez, P. C., Correa, J. A., and Garcia-Reina, G. (1996). Host-specificity of Endophyton ramosum (Chlorophyta), the causative agent of green patch disease in Mazzaella laminarioides (Rhodophyta). Eur. J. Phycol. 31:173–9.CrossRefGoogle Scholar
Sanders, H. L., Grassle, J. F., Hampson, G. R. et al. (1980). Anatomy of an oil spill: long-term effects from the grounding of the barge Florida off West Falmouth, Massachusetts. J. Mar. Res. 38:265–380.Google Scholar
Sand-Jensen, K. (1987). Environmental control of bicarbonate use among freshwater and marine macrophytes. In Crawford, R. M. M. (ed.) Plant Life in Aquatic and Amphibious Habitats (pp. 99–112). Oxford: Blackwell Scientific.Google Scholar
Sanina, N. M., Goncharova, S. N., and Kostetsky, E. Y. (2004). Fatty acid composition of individual polar lipid classes from marine macrophytes. Phytochemistry 65:721–30.CrossRefGoogle ScholarPubMed
Santelices, B. (1990). Patterns of reproduction, dispersal and recruitment in seaweeds. Oceanogr. Mar. Biol. Annu. Rev. 28:177–276. [, , , ]Google Scholar
Santelices, B. (1991). Production ecology of Gelidium. Hydrobiol. 221:31–44.CrossRefGoogle Scholar
Santelices, B. (1999). How many kinds of individual are there? Tree 14:152–5.Google Scholar
Santelices, B. (2002). Recent advances in fertilization ecology of macroalgae. J. Phycol. 38:4–10.CrossRefGoogle Scholar
Santelices, B. (2004a). A comparison of ecological responses among aclonal (unitary), clonal and coalescing macroalgae. J. Exp. Mar. Biol. Ecol. 300:31–64.CrossRefGoogle Scholar
Santelices, B. (2004b). Mosaicism and chimerism as components of intraorganismal genetic heterogeneity. J. Evol. Biol. 17:1187–8.CrossRefGoogle ScholarPubMed
Santelices, B., and Doty, M. S. (1989). A review of Gracilaria farming. Aquaculture 78:95–133.CrossRefGoogle Scholar
Santelices, B., and Martinez, E. (1988). Effects of filter feeders and grazers on algal settlement and growth in mussel beds. J. Exp. Mar. Biol. Ecol. 118:281–306.CrossRefGoogle Scholar
Santelices, B., and Ojeda, F. P. (1984). Recruitment, growth and survival of Lessonia nigrescens (Phaeophyta) at various tidal levels in exposed habitats of central Chile. Mar. Ecol. Prog. Ser. 19:73–82.CrossRefGoogle Scholar
Santelices, B., and Paya, I. (1989). Digestion survival of algae: some ecological comparisons between free spores and propagules in fecal pellets. J. Phycol. 25:693–9.CrossRefGoogle Scholar
Santelices, B., Montalva, S., and Oliger, P. (1981). Competitive algal community organization in exposed intertidal habitats from central Chile. Mar. Ecol. Prog. Ser. 6:267–76.CrossRefGoogle Scholar
Santelices, B., Correa, J., and Avila, M. (1983). Benthic algal spores surviving digestion by sea urchins. J. Exp. Mar. Biol. Ecol. 70:263–9.CrossRefGoogle Scholar
Santelices, B., Hoffmann, A. J., Aedo, D., Bobadilla, M., and Otaiza, R. (1995). A bank of microscopic forms on disturbed boulders and stones in tide pools. Mar. Ecol. Prog. Ser. 129:215–28.CrossRefGoogle Scholar
Santelices, B., Correa, J. A., Aedo, D., Hormazábal, M., and Sànchez, P. (1999). Convergent biological processes in coalescing Rhodophyta. J. Phycol. 35:1127–49.CrossRefGoogle Scholar
Santelices, B., Correa, J. A., Meneses, I., Aedo, D., and Varela, D. (1996). Sporeling coalescence and intraclonal variation in Gracilaria chilensis (Gracilariales, Rhodophyta). J. Phycol. 32:313–22.CrossRefGoogle Scholar
Santelices, B., Bolton, J. J., and Meneses, I. (2009). Marine algal communities. In Witman, J. D., and Roy, K. (eds), Marine Macroecology (pp. 153–92). Chicago: University of Chicago Press.CrossRefGoogle Scholar
Santelices, B., Alvarado, J. L., and Flores, V. (2010). Size increments due to interindividual fusions: how much and for how long? J. Phycol. 46:685–92.CrossRefGoogle Scholar
Santos, R. (1995). Size structure and inequality in a commercial stand of the seaweed Gelidium sesquipedale. Mar. Ecol. Prog. Ser. 119:253–63.CrossRefGoogle Scholar
Saroussi, S., and Beer, S. (2007). Alpha and quantum yield of aquatic plants derived from PAM fluorometry: uses and misuses. Aquat. Bot. 86: 89–92.CrossRefGoogle Scholar
Saunders, G. W. (2005). Applying DNA barcoding to red macroalgae: a preliminary appraisal holds promise for future applications. Phil. Trans. R. Soc. London B 360:1879–88.CrossRefGoogle ScholarPubMed
Saunders, G. W. (2008). A DNA barcode examination of the red algal family Dumontiaceae in Canadian waters reveals substantial cryptic species diversity. 1. The foliose Dilsea-Neodilsea complex and Weeksia. Botany 86:773–89.CrossRefGoogle Scholar
Saunders, G. W. (2009). Routine DNA barcoding of Canadian Gracilariales (Rhodophyta) reveals the invasive species Gracilaria vermiculophylla in British Columbia. Molec. Ecol. Res. 9(s1):140–50.CrossRefGoogle ScholarPubMed
Saunders, G. W., and Druehl, L. D. (1992). Nucleotide sequences of the small–subunit ribosomal RNA genes from selected Laminariales (Phaeophyta): implications for kelp evolution. J. Phycol. 28:544–9.CrossRefGoogle Scholar
Saunders, G.W., and Druehl, L. D. (1993). Revision of the kelp family Alariaceae and the taxonomic affinities of Lessoniopsis Reinke (Laminariales, Phaeophyta). Hydrobiologia 260/261:689–97.CrossRefGoogle Scholar
Saunders, G.W., and Hommersand, M. (2004). Assessing red algal supraordinal diversity and taxonomy in the context of contemporary systematic data. Amer. J. Bot. 91:1494–507.CrossRefGoogle ScholarPubMed
Saunders, G. W., and Kraft, G. T. (1997). A molecular perspective on red algal evolution: focus on the Florideophycidae. Plant Systematics and Evolution (Supplement) 11:115–38.CrossRefGoogle Scholar
Saunders, G.W., and Lehmkuhl, K. V. (2005). Molecular divergence and morphological diversity among four cryptic species of Plocamium (Plocamiales, Florideophyceae) in northern Europe. Eur. J. Phycol. 40: 293–312.CrossRefGoogle Scholar
Saunders, G.W., Bird, C. J., Ragan, M. A., and Rice, E. L. (1995). Phylogenetic relationships of species of uncertain taxonomic position within the Acrochaetiales/Palmariales complex (Rhodophyta): inferences from phenotypic and 18S rDNA sequence data. J. Phycol. 31:601–11.CrossRefGoogle Scholar
Saxena, I. M., and Brown, M. (2005). Cellulose biosynthesis: current views and evolving concepts. Ann. Bot.-London 96:9–21.CrossRefGoogle ScholarPubMed
Scanlan, C. M., and Wilkinson, M. (1987). The use of seaweeds in biocide toxicity testing. Part I. The sensitivity of different stages in the life-history of Fucus, and of other algae, to certain biocides. Mar. Environ. Res. 21:11–29.CrossRefGoogle Scholar
Schaffelke, B. (1995a). Storage carbohydrates and abscisic acid contents in Laminaria hyperborea are entrained by experimental daylengths. Eur. J. Phycol. 30:313–17.CrossRefGoogle Scholar
Schaffelke, B. (1995b). Abscisic acid in sporophytes of 3 Laminaria species (Phaeophyta). J. Plant Physiol. 146:453–8.CrossRefGoogle Scholar
Schaffelke, B. (1999a). Short-term nutrient pulses as tools to assess responses of coral reef macroalgae to enhanced nutrient availability. Mar. Ecol. Prog. Ser. 182:305–10.CrossRefGoogle Scholar
Schaffelke, B. (1999b). Particulate organic matter as an alternative nutrient source for tropical Sargassum species (Fucales, Phaeophyceae). J. Phycol. 35:1150–7.CrossRefGoogle Scholar
Schaffelke, B. (2001). Surface alkaline phosphatase activities of macroalgae on coral reefs of the central Great Barrier Reef, Australia. Coral Reefs 19:310–17.CrossRefGoogle Scholar
Schaffelke, B., and Deane, D. (2005). Desiccation tolerance of the introduced marine green alga Codium fragile ssp. tomentosoides: clues for likely transport vectors? Biol. Invas. 7:557–65.CrossRefGoogle Scholar
Schaffelke, B., and Hewitt, C. L. (2007). Impacts of introduced seaweeds. Bot. Mar. 50:397–417.CrossRefGoogle Scholar
Schaffelke, B., and Klumpp, D. W. (1998a). Nutrient-limited growth of the coral reef macroalga Sargassum baccularia and experimental growth enhancement by nutrient addition in continuous-flow culture. Mar. Ecol. Progr. Ser. 164:199–211.CrossRefGoogle Scholar
Schaffelke, B., and Klumpp, D. W. (1998b). Short-term nutrient pulses enhance growth and photosynthesis of the coral reef macroalga Sargassum baccularia. Mar. Ecol. Progr. Ser. 170:95–105.CrossRefGoogle Scholar
Schaffelke, B., and Lüning, K. (1994). A circannual rhythm controls seasonal growth in the kelps Laminaria hyperborea and L. digitata from Helgoland (North Sea). Eur. J. Phycol. 29:49–56.CrossRefGoogle Scholar
Schaffelke, B., Peters, A. F., and Reusch, T. B. H. (1996). Factors influencing depth distribution of soft bottom inhabiting Laminaria saccharina (L.) Lamour. in Kiel Bay, Western Baltic. Hydrobiologia 326/327:117–23.CrossRefGoogle Scholar
Schaffelke, B., Smith, J. E., and Hewitt, C. L. (2006). Introduced macroalgae: a growing concern. J. Appl. Phycol. 18:529–41.CrossRefGoogle Scholar
Schagerl, M., and Möstl, M. (2011). Drought stress, rain and recovery of the intertidal seaweed Fucus spiralis. Mar. Biol. 158:2471–9.CrossRefGoogle Scholar
Schatz, S. (1980). Degradation of Laminaria saccharina by higher fungi: a preliminary report. Bot. Mar. 23: 617–22.Google Scholar
Scheibling, R. E., and Gagnon, P. (2006). Competitive interactions between the invasive alga Codium fragile ssp. tomentosoides and native canopy-forming seaweeds in Nova Scotia (Canada). Mar. Ecol. Prog. Ser. 325:1–14.CrossRefGoogle Scholar
Scheibling, R. E., and Hatcher, B. G. (2007). Ecology of Strongylocentrotus droebachiensis. In Lawrence, J. M. (ed.), Edible Sea Urchins: Biology and Ecology (pp. 353–92). New York: Elsevier.CrossRefGoogle Scholar
Schiel, D. R. (2006). Rivets or bolts? When single species count in the function of temperate rocky reef communities. J. Exp. Mar. Biol. Ecol. 338:233–52.CrossRefGoogle Scholar
Schiel, D. R., and Choat, J. H. (1980). Effects of density on monospecific stands of marine algae. Nature 285:324–6.CrossRefGoogle Scholar
Schiel, D. R., and Foster, M. S. (1986). The structure of subtidal algal stands in temperate waters. Oceanogr. Mar. Biol. Annu. Rev. 24:265–307. [, , , , , , ]Google Scholar
Schiel, D. R., and Foster, M. S. (2006). The population biology of large brown seaweeds: ecological consequences of multiphase life histories in dynamic coastal environments. Annu. Rev. Ecol. Evol. Syst. 37:343–72.CrossRefGoogle Scholar
Schiel, D. R., and Lilley, S. A. (2011). Impacts and negative feedbacks in community recovery over eight years following removal of habitat-forming macroalgae. J. Exp. Mar. Biol. Ecol. 407:108–15.CrossRefGoogle Scholar
Schiel, D. R., Steinbeck, J. R., and Foster, M. S. (2004). Ten years of induced ocean warming causes comprehensive changes in marine benthic communities. Ecol. 85:1833–9.CrossRefGoogle Scholar
Schiel, D. R., Wood, S. A., Dunmore, R. A., and Taylor, D. I. (2006). Sediment on rocky intertidal reefs: effects on early post-settlement stages of habitat-forming seaweeds. J. Exp. Mar. Biol. Ecol. 331:158–72.CrossRefGoogle Scholar
Schiff, J. A. (1983). Reduction and other metabolic reactions of sulfate. In Uiuchli, A. and Bieleski, R. L. (eds), Encyclopaedia of Plant Physiology (Vol. 15, pp. 382–99). Berlin:Springer-Verlag.Google Scholar
Schils, T., and Wilson, S. C. (2006). Temperature threshold as a biogeographic barrier in northern Indian Ocean macroalgae. J. Phycol. 42:749–56.CrossRefGoogle Scholar
Schmid, C. E. (1993). Cell–cell-recognition during fertilization in Ectocarpus siliculosus (Phaeophyceae). Hydrobiologia 260/261:437–43.CrossRefGoogle Scholar
Schmid, C. E., Schroer, N., and Müller, D. G. (1994). Female gamete membrane glycoproteins potentially involved in gamete recognition in Ectocarpus siliculosus (Phaeophyceae). Plant Sci. 102:61–7.CrossRefGoogle Scholar
Schmid, R. (1984). Blue light effects on morphogenesis and metabolism in Acetabularia. In Senger, H. (ed.), Blue Light Effects in Biological Systems (pp. 419–32). Berlin: Springer-Verlag.CrossRefGoogle Scholar
Schmid, R., and Dring, M. J. (1996). Blue light and carbon acquisition in brown algae: an overview and recent developments. Sci. Mar. 60:115–24.Google Scholar
Schmid, R., Tunnermann, M., and Idziak, E.-M. (1990). Role of red light in hair-formation induced by blue light inAcetabularia mediterranea. Planta 181:144–7.Google ScholarPubMed
Schmitt, T. M., Lindquist, N., and Hay, M. E. (1998). Seaweed secondary metabolites as antifoulants: effects of Dictyota spp. diterpenes on survivorship, settlement, and development of marine invertebrate larvae. Chemoecology 8:125–31.CrossRefGoogle Scholar
Schmitz, K. (1981). Translocation. In Lobban, C. S. and Wynne, M. J. (eds), The Biology of Seaweeds (pp. 534–58) Oxford: Blackwell Scientific.Google Scholar
Schmitz, K., and Riffarth, W. (1980). Carrier-mediated uptake of L-leucine by the brown alga Giffordia mitchellliae. Z. Pflanzenphysiol. 67:311–24.CrossRefGoogle Scholar
Schmitz, K., and Srivastava, L. M. (l974). Fine structure and development of sieve tubes in Laminaria groenlandica Rosenv. Cytobiologie 10:66–87.Google Scholar
Schneider, C. W. (1976). Spatial and temporal distributions of benthic marine algae on the continental shelf of the Carolinas. Bull. Mar. Sci. 26:133–51.Google Scholar
Schoch, G. C., Menge, B. A., Allison, G., et al. (2006). Fifteen degrees of separation: latitudinal gradients of rocky intertidal biota along the California Current. Limnol. Oceanogr. 51:2564–85.CrossRefGoogle Scholar
Schoenwaelder, M. E. A. (2002). The occurrence and cellular significance of physodes in brown algae. Phycologia 41:125–39.CrossRefGoogle Scholar
Schoenwaelder, M. E. A., and Clayton, M. N. (1999). The role of the cytoskeleton in brown algal physode movement. Eur. J. Phycol. 34:223–9.CrossRefGoogle Scholar
Schofield, O., Evens, T. J., and Millie, D. F. (1998). Photosystem II quantum yields and xanthophyll-cycle pigments of the macroalga Sargassum natans (Phaeophyceae): responses under natural sunlight. J. Phycol. 34: 104–12.CrossRefGoogle Scholar
Schonbeck, M., and Norton, T. A. (1978). Factors controlling the upper limits of fucoid algae on the shore. J. Exp. Mar. Biol. Ecol. 31:303–13. [, , ]CrossRefGoogle Scholar
Schonbeck, M. W., and Norton, T. A. (l979a). An investigation of drought avoidance in intertidal fucoid algae. Bot. Mar. 22:133–44.CrossRefGoogle Scholar
Schonbeck, M. W., and Norton, T. A. (l979b). Drought hardening in the upper shore seaweeds Fucus spiralis and Pelvetia canaliculata. J. Ecol. 67:687–96.CrossRefGoogle Scholar
Schonbeck, M. W., and Norton, T. A. (l979c). The effects of brief periodic submergence on intertidal algae. Estu. Cstl. Mar: Sci. 8:205–11.CrossRefGoogle Scholar
Schonbeck, M. W., and Norton, T. A. (1980a). Factors controlling the lower limits of fucoid algae on the shore. J. Exp. Mar. Biol. Ecol. 43:131–50.CrossRefGoogle Scholar
Schonbeck, M. W., and Norton, T. A. (1980b). The effects on intertidal fucoids of exposure to air under various conditions. Bot. Mar. 23:141–7.CrossRefGoogle Scholar
Schramm, W. (1999). Factors influencing seaweed responses to eutrophication: some results from EU-project EUMAC. J. Appl. Phycol. 11:69–78. [, , ]CrossRefGoogle Scholar
Schreiber, U., Bilger, W., and Neubauer, C. (1994). Chlorophyll fluorescence as a non-intrusive indicator for rapid assessment of in vivo photosynthesis. In Schulze, E. D. and Caldwell, M. M. (eds) Ecophysiology of Photosynthesis (pp. 49–70). Berlin: Springer-Verlag.Google Scholar
Schubert, H., Sagert, S., and Forster, R. M. (2001) Evaluation of the different levels of variability in the underwater light field of a shallow estuary. Helgoland Mar. Res. 55:12–22.CrossRefGoogle Scholar
Schuenhoff, A., Shpigel, M., Lupatsch, I., et al. (2003). A semi-recirculating integrated system for the culture of fish and seaweed. Aquaculture 221:167–81.CrossRefGoogle Scholar
Schulze, T., Prager, K., Dathe, H., et al. (2010). How the green alga Chlamydomonas reinhardtii keeps time. Protoplasma 244:3–14.CrossRefGoogle ScholarPubMed
Schumacher, J. F., Carman, M. L., Estes, T. G., et al. (2007). Engineered antifouling microtopographies – effect of feature size, geometry, and roughness on settlement of zoospores of the green alga Ulva. Biofouling 23:55–62.CrossRefGoogle ScholarPubMed
Schweikert, K., Sutherland, J. E. S., Hurd, C. L., and Burritt, D. J. (2011). UV-B radiation induces changes in polyamine metabolism in the red seaweed Porphyra cinnamomea. Plant Growth Regul. 65:389–99.CrossRefGoogle Scholar
Schwenk, K., Padilla, D. K., Bakken, G. S., and Full, R. J. (2009). Grand challenges in organismal biology. Integr. Comp. Biol. 49:7–14.CrossRefGoogle ScholarPubMed
Scott, F. J., Wetherbee, R., and Kraft, G. T. (1984). The morphology and development of some prominently stalked southern Australian Halymeniaceae (Cryptonemiales, Rhodophyta). II. The sponge-associated genera Thamnoclonium Kuetzing and Codiophyllum Gray. J. Phycol. 20:286–95.CrossRefGoogle Scholar
Scrosati, R. (2002a). An updated definition of genet applicable to clonal seaweeds, bryophytes, and vascular plants. Basic Appl. Ecol. 3:97–9.CrossRefGoogle Scholar
Scrosati, R. (2002b). Morphological plasticity and apparent loss of apical dominance following the natural loss of the main apex in Pterocladiella capillacea (Rhodophyta, Gelidiales) fronds. Phycologia 41:96–8.CrossRefGoogle Scholar
Scrosati, R. (2005). Review of studies on biomass-density relationships (including self-thinning lines) in seaweeds: main contributions and persisting misconceptions. Phycol. Res. 53:224–33.CrossRefGoogle Scholar
Scrosati, R., and Heaven, C. (2008). Trends in abundance of rocky intertidal seaweeds and filter feeders across gradients of elevation, wave exposure, and ice scour in eastern Canada. Hydrobiologia 603:1–14.CrossRefGoogle Scholar
Scrosati, R., and Mudge, B. (2004). Persistence of gametophyte predominance in Chondrus crispus (Rhodophyta, Gigartinaceae) from Nova Scotia after 12 years. Hydrobiologia 519:215–18.CrossRefGoogle Scholar
Scrosati, R., and De Wreede, R. E. (1997). Dynamics of the biomass-density relationship and frond biomass inequality for Mazzaella cornucopiae (Gigartinaceae, Rhodophyta): implications for the understanding of frond interactions. Phycologia 36:506–16.CrossRefGoogle Scholar
Searles, R. B. (1980). Strategy of the red algal life-history. Amer. Nat. 115:113–20.CrossRefGoogle Scholar
Sears, J. R., and Wilce, R. T. (1975). Sublittoral, benthic marine algae of southern Cape Cod and adjacent islands: seasonal periodicity, associations, diversity, and floristic composition. Ecol. Monogr. 45:337–65.CrossRefGoogle Scholar
Seery, C. R., Gunthorpe, L., and Ralph, P. J. (2006). Herbicide impact on Hormosira banksii gametes measured by fluorescence and germination bioassays. Environ Poll. 140:43–51.CrossRefGoogle ScholarPubMed
Semesi, I., Beer, S., and Björk, M. (2009). Seagrass photosynthesis controls rates of calcification and photosynthesis of calcareous macroalgae in a tropical seagrass meadow. Mar. Ecol. Prog. Ser. 382:41–7.CrossRefGoogle Scholar
Serikawa, K. A., and Mandoli, D. F. (1999). Aaknox1, a kn1-like homeobox gene in Acetabularia acetabulum, undergoes developmentally regulated subcellular localization. Plant Mol. Biol. 41:785–93.CrossRefGoogle ScholarPubMed
Serikawa, K. A., Porterfield, D. M., and Mandoli, D. F. (2001). Asymmetric subcellular mRNA distribution correlates with carbonic anhydrase activity in Acetabularia acetabulum. Plant Physiol. 125:900–11.CrossRefGoogle ScholarPubMed
Serrão, E. A., Pearson, G., Kautsky, L., and Brawley, S. H. (1996). Successful external fertilization in turbulent environments. P. Natl. Acad. Sci. USA. 93:5286–90.CrossRefGoogle ScholarPubMed
Serrão, E. A., Alice, L. A., and Brawley, S. H. (1999). Evolution of the Fucaceae (Phaeophyceae) inferred from nrDNA-ITS. J. Phycol. 35:382–94.CrossRefGoogle Scholar
Setchell, W. A., and Gardner, N. L. (1919). The marine algae of the Pacific Coast of North America. Part I. Myxophyceae. Univ. Calif. Publ. Bot. 8:1–138.Google Scholar
Setchell, W. A., and Gardner, N. L. (1925). III. Melanophyceae. The Marine Algae of the Pacific Coast of North America (pp. 383–898). Berkley, CA: Univ. of California Press.Google Scholar
Seymour, R. J., Tegner, M. J., Dayton, P. K., and Parnell, P. E. (1989). Storm wave induced mortality of giant kelp, Macrocystis pyrifera, in southern California. Estu. Cstl. Shelf Sci. 28:277–92. [, , ]CrossRefGoogle Scholar
Sfriso, A., Marcomini, A., and Pavoni, B. (1987). Relationships between macroalgal biomass and nutrient concentrations in a hypertrophic area of the Venice Lagoon. Mar. Environ. Res. 22:297–312.CrossRefGoogle Scholar
Sharp, G. (1987). Ascophyllum nodosum and its harvesting in eastern Canada. In Doty, M. S., Caddy, J. F., and Santelices, B. (eds), Case Studies of Seven Commercial Seaweed Resources (pp. 3–48). FAO Fish. Tech. Pap. 281.Google Scholar
Sharrock, R. A. (2008). The phytochrome red/far-red photoreceptor superfamily. Genome Biol. 9:230.CrossRefGoogle ScholarPubMed
Shaughnessy, F. J., De Wreede, R. E., and Bell, E. C. (1996). Consequences of morphology and tissue strength to blade survivorship of two closely related Rhodophyta species. Mar. Ecol. Prog. Ser. 136:257–66.CrossRefGoogle Scholar
Shears, N. T., and Babcock, R. C. (2003). Continuing trophic cascade effects after 25 years of no-take marine reserve protection. Mar. Ecol. Prog. Ser. 246:1–16.CrossRefGoogle Scholar
Shepherd, V. A., Beilby, M. J., and Bisson, M. A. (2004). When is a cell not a cell? A theory relating coenocytic structure to the unusual electrophysiology of Ventricaria ventricosa (Valonia ventricosa). Protoplasma 223:79–91.CrossRefGoogle Scholar
Shimshock, N., Sennefelder, G., Dueker, M., Thurberg, F., and Yarish, C. (1992). Patterns of metal accumulation in Laminaria longicruris from Long Island Sound (Connecticut). Arch. Environ. Contam. Toxicol. 22:305–12.CrossRefGoogle Scholar
Shivji, M. S. (1985). Interactive effects of light and nitrogen on growth and chemical composition of juvenile Macrocystis pyrifera (L.) C. Ag. (Phaeophyta) sporophytes. J. Exp. Mar. Biol. Ecol. 89:81–96.CrossRefGoogle Scholar
Shivji, M. S. (1991). Organization of the chloroplast genome in the red algaPorphyra yezoensis. Curr. Genet. 19:49–54.CrossRefGoogle Scholar
Sieburth, J. M. (1969). Studies on algal substances in the sea. III. The production of extracellular organic matter by littoral marine algae. J. Exp. Mar. Biol. Ecol. 3: 290–309.CrossRefGoogle Scholar
Sieburth, J. M., and Tootle, J. T. (1981). Seasonality of microbial fouling on Ascophyllum nodosum (L.) Lejol., Fucus vesiculosus L., Polysiphonia lanosa (L.) Tandy and Chondrus crispus Stackh. J. Phycol. 17:57–64.CrossRefGoogle Scholar
Silberfeld, T., Leigh, J. W., Verbruggen, H., et al. (2010). A multi-locus time-calibrated phylogeny of the brown algae (Heterokonta, Ochrophyta, Phaeophyceae): investigating the evolutionary nature of the “brown algal crown radiation”. Mol. Phylogenet. Evol. 56:659–74.CrossRefGoogle ScholarPubMed
Simkiss, K., and Wilbur, K. M. (1989). Biomineralization: Cell Biology and Mineral Deposition. Orlando, FL: Academic Press, 340 pp.Google Scholar
Simpson, C. L., and Stern, D. B. (2002). The treasure trove of algal chloroplast genomes. Surprises in architecture and gene content, and their functional implications. Plant Physiol. 129:957–66.CrossRefGoogle ScholarPubMed
Singh, P, Kumar, P. A., Abroi, Y. P., and Naik, M. S. (1985). Photorespiratory nitrogen cycle – a critical evaluation. Physiol. Plant. 66:169–76.CrossRefGoogle Scholar
Sjøtun, , Fredriksen, K. S., and Rueness, J. (1998). Effect of canopy biomass and wave exposure on growth in Laminaria hyperborea (Laminariaceae, Phaeophya). Eur. J. Phycol. 33:337–43.CrossRefGoogle Scholar
Sjøtun, K., Christie, H., and Fosså, J. H. (2006). The combined effect of canopy shading and sea urchin grazing on recruitment in kelp forests (Laminaria hyperborea). Mar. Biol. Res. 2:24–32.CrossRefGoogle Scholar
Skene, K. H. (2004). Key differences in photosynthetic characteristics of nine species of intertidal macroalgae are related to their position on the shore. Can. J. Bot. 82:177–84.CrossRefGoogle Scholar
Slocum, C. J. (1980). Differential susceptibility to grazers in two phases of an intertidal alga: advantages of heteromorphic generations. J. Exp. Mar. Biol. Ecol. 46:99–110.CrossRefGoogle Scholar
Smetacek, V., von Bodungen, B., von Brödsel, K., and Zeitzschel, B. (1976). The plankton tower.II. Release of nutrients from sediments due to changes in the density of bottom water. Mar. Biol. 34:373–8.CrossRefGoogle Scholar
Smit, A.J. (2004). Medicinal and pharmaceutical uses of seaweed natural products: A review. J. Appl. Phycol. 16:245–62.CrossRefGoogle Scholar
Smit, A. J., Brearley, A., Hyndes, G. A., Lavery, P. S., and Walker, D. I. (2005). Carbon and nitrogen stable isotope analysis of an Amphibolis griffithii seagrass bed. Estu. Cstl. Shelf Sci. 65(3):545–56.CrossRefGoogle Scholar
Smit, A. J., Brearley, A., Hyndes, G. A., Lavery, P. S., and Walker, D. I. (2006). delta N-15 and delta C-13 analysis of a Posidonia sinuosa seagrass bed. Aquat. Bot. 84(3):277–82.CrossRefGoogle Scholar
Smith, A. H., Nichols, K., and McLachlan, J. (1984). Cultivation of seamoss (Gracilaria) in St. Lucia, West Indies. Hydrobiologia 116/117:249–51.CrossRefGoogle Scholar
Smith, A. M., Sutherland, J. E., Kregting, L., Farr, T. J., and Winter, D. J. (2012). Phylominerology of the Coralline red algae: Correlation of skeletal mineralogy with molecular phylogeny. Phytochem. 81:97–108.CrossRefGoogle ScholarPubMed
Smith, C. M., and Walters, L. J. (1999). Fragmentation as a strategy for Caulerpa species: Fates of fragments and implications for management of an invasive weed. P. S. Z. N.; Mar. Ecol. 20: 307–19.CrossRefGoogle Scholar
Smith, D. R., Hua, J., Lee, R. W., and Keeling, P. J. (2012). Relative rates of evolution among the three genetic compartments of the red alga Porphyra differ from those of green plants and do not correlate with genome architecture. Mol. Phylogen. Evol. 65:339–44.CrossRefGoogle Scholar
Smith, G. M. (1947). On the reproduction of some Pacific coast species of Ulva. Am. J. Bot. 34:80–7.CrossRefGoogle ScholarPubMed
Smith, J. E., Hunter, C. L., Conklin, E. J., et al. (2004). Ecology of the invasive red alga Gracilaria salicornia (Rhodophyta) on O’ahu, Hawai’i. Pacific Science 58:325–43.CrossRefGoogle Scholar
Smith, J. E., Conklin, E. J., Smith, C. M., and Hunter, C. L. (2008). Fighting algae in Kaneohe Bay (response). Science 319:157–8.Google Scholar
Smith, J., Hunter, C., and Smith, C. (2010). The effects of top–down versus bottom–up control on benthic coral reef community structure. Oecologia 163:497–507.CrossRefGoogle ScholarPubMed
Smith, R. C., and Tyler, J. E. (1974). In Jerlov, N. G. and Steemann-Nielsen, E (eds), Optical Aspects of Oceanography. Orlando, FL: Academic Press.Google Scholar
Smith, R. C., and Tyler, J. E. (1976). Transmission of solar radiation into natural waters. Photochem. Photobiol. Rev. 1:117–55.CrossRefGoogle Scholar
Smith, R. G., Wheeler, W. N., and Srivastava, L. M. (1983). Seasonal photosynthetic performance of Macrocystis integrifolia (Phaeophyceae). J. Phycol. 19:352–9.CrossRefGoogle Scholar
Smith, S. D. A. (2002). Kelp rafts in the Southern Ocean. Global Ecol. Biogeogr. 11:67–9.CrossRefGoogle Scholar
Smith, S. V., Kimmerer, W., Laws, E., Brock, R., and Walsh, T. (1981). Kaneohe Bay sewage diversion experiment: perspectives on ecosystem responses to nutritional perturbation. Pacific Sci. 35:270–395.Google Scholar
Solis, M. J. L., Draeger, S., and dela Cruz, T. E. (2010). Marine-derived fungi from Kappaphycus alvarezii and K. striatum as potential causative agents of ice-ice disease in farmed seaweeds. Bot. Mar. 53:587–94.CrossRefGoogle Scholar
Sorte, C. J. B., Williams, S. L., and Carlton, J. T. (2010). Marine range shifts and species introductions: comparative spread rates and community impacts. Global Ecol. Biogeogr. 19:303–16.CrossRefGoogle Scholar
Sotka, E. E. (2005). Local adaptation in host use among marine invertebrates. Ecol. Lett. 8:448–59.CrossRefGoogle Scholar
Sotka, E. E., and Reynolds, P. L. (2011). Rapid experimental shift in host use traits of a polyphagous marine herbivore reveals fitness costs on alternative hosts. Evol. Ecol. 25:1335–55.CrossRefGoogle Scholar
Sotka, E. E., and Whalen, K. E. (2008). Herbivore offense in the sea: the detoxification and transport of secondary metabolites. In Amsler, C. D. (ed.), Algal Chemical Ecology (pp. 203–28). London: Springer (Limited).CrossRefGoogle Scholar
Sotka, E. E., Hay, M. E., and Thomas, J. D. (1999). Host-plant specialization by a non-herbivorous amphipod: advantages for the amphipod and costs for the seaweed. Oecologia 118:471–82.CrossRefGoogle ScholarPubMed
Sotka, E. E., Taylor, R. B., and Hay, M. E. (2002). Tissue-specific induction of resistance to herbivores in a brown seaweed: the importance of direct grazing versus waterborne signals from grazed neighbors. J. Exp. Mar. Biol. Ecol. 277:1–12.CrossRefGoogle Scholar
Sotka, E. E., Wares, J. P., and Hays, M. E. (2003). Geographic and genetic variation in feeding preference for chemically defended seaweeds. Evolution. 57:2262–767.CrossRefGoogle ScholarPubMed
Sotka, E. E., Forbey, J., Horn, M., et al. (2009). The emerging role of pharmacology in understanding consumer-prey interactions in marine and freshwater systems. Integr. Comp. Biol. 49:291–313.CrossRefGoogle ScholarPubMed
Soulsby, P. G., Lowthion, D., Houston, M., and Montgomery, H. A. C. (1985). The role of sewage effluent in the accumulation of macroalgal mats on intertidal mudflats in two basins in southern England. Neth. J. Sea Res. 19:257–63.CrossRefGoogle Scholar
Sousa, W. P. (1979). Experimental investigation of disturbance and ecological succession in a rocky intertidal algal community. Ecol. Monogr. 49:227–54.CrossRefGoogle Scholar
Sousa, W. P. (2001). Natural disturbance and the dynamics of marine benthic communities. In Bertness, M. D., Gaines, S. D., and Hay, M. E. (eds), Marine Community Ecology (pp. 85–130). Sunderland, MA: Sinauer Assocs. [, , ]Google Scholar
Sousa, W. P., Schroeter, S. C., and Gaines, S. D. (1981). Latitudinal variation in intertidal algal community structure: the influence of grazing and vegetative propagation. Oecologia 48:297–307.CrossRefGoogle ScholarPubMed
Spaargaren, D. H. (1984). On ice formation in sea water and marine animals at subzero temperatures. Mar. Biol. Lett. 5: 203–16.Google Scholar
Spalding, H., Foster, M. S., and Heine, J. N. (2003). Composition, distribution, and abundance of deep-water (>30 m) macroalgae in central California. J. Phycol. 39:273–84.CrossRefGoogle Scholar
Speransky, S. R., Brawley, S. H., and Halteman, W. A. (2000). Gamete release is increased by calm conditions in the coenocytic green alga Bryopsis (Chlorophyta). J. Phycol. 36:730–9.CrossRefGoogle Scholar
Speransky, V. V., Brawley, S. H., and McCully, M. E. (2001). Ion fluxes and modification of the extracellular matrix during gamete release in fucoid algae. J. Phycol. 37:555–73.CrossRefGoogle Scholar
Spilling, K., Titelman, J., Greve, T. M., and Kühl, M. (2010). Microsensor measurements of the external and internal microenvironment of Fucus vesiculosus (Phaeophyceae). J. Phycol. 46:1350–5.CrossRefGoogle Scholar
Springer, Y. P., Hays, C. G., Carr, M. H., and Mackey, M. R. (2010). Toward ecosytem-based management of marine macroalgae – the bull kelp, Nereocystis luetkeana. Oceanogr. Mar. Biol. 48:1–42.Google Scholar
Stachowicz, J. J., and Hay, M. E. (1999). Reducing predation through chemically mediated camouflage: indirect effects of plant defenses on herbivores. Ecology 80:495–509.CrossRefGoogle Scholar
Stachowicz, J. J., and Whitlatch, R. B. (2005). Multiple mutualists provide complementary benefits to their seaweed host. Ecology 86:2418–27.CrossRefGoogle Scholar
Stachowicz, J. J., Bruno, J. F., and Duffy, J. E. (2007). Understanding the effects of marine biodiversity on communities and ecosystems. Annu. Rev. Ecol. Evol. Syst. 38:739–66.CrossRefGoogle Scholar
Stanley, S. M., Ries, J. B., and Hardie, L. A. (2010). Increased production of calcite and slower growth for the major sediment-producing alga Halimeda as the Mg/Ca ratio of seawater is lowered to a “calcite sea” level. J. Sedimentary Res. 80:6–16.CrossRefGoogle Scholar
Stauber, J. L., and Florence, T. M. (1985). Interactions of copper and manganese: a mechanism by which manganese alleviates copper toxicity to the marine diatom, Nitzschia closterium (Ehrenberg) W. Smith. Aquat. Toxic. 7:241–54.CrossRefGoogle Scholar
Stauber, J. L., and Florence, T. M. (1987). Mechanisms of toxicity of ionic copper and copper complexes to algae. Mar. Biol. 94:511–19.CrossRefGoogle Scholar
Steele, R. L., and Hanisak, M. D. (1979). Sensitivity of some brown algal reproductive stages to oil pollution. Proc. Intl. Seaweed Symp. 9:181–91.Google Scholar
Steele, R. L., Thursby, G. B., and van der Meer, J. P. (1986). Genetics of Champia parvula (Rhodymeniales, Rhodophyta): Mendelian inheritance of spontaneous mutants. J. Phycol. 22:538A–42.CrossRefGoogle Scholar
Steemann-Nielsen, E. (1974). Light and primary production. In Jerlov, N. G. (ed.) Optical Aspects of Oceanography (pp. 331–88). New York: Academic Press.Google Scholar
Steen, H. (2004). Interspecific competition between Enteromorpha (Ulvales: Chlorophyceae) and Fucus (Fucales: Phaeophyceae) germlings: effects of nutrient concentration, temperature, and settlement density. Mar. Ecol. Prog. Ser. 278:89–101.CrossRefGoogle Scholar
Steen, H., and Scrosati, R. (2004). Intraspecific competition in Fucus serratus and F. evanescens (Phaeophyceae: Fucales) germlings: effects of settlement density, nutrient concentration, and temperature. Mar. Biol. 14:61–70.Google Scholar
Steinbeck, J. R., Schiel, D. R., and Foster, M. S. (2005). Detecting long-term change in complex communities: a case study from the rocky intertidal zone. Ecol. Appl. 15:1813–32.CrossRefGoogle Scholar
Steinberg, P. D., and Altena, I. (1992). Tolerance of marine invertebrate herbivores to brown algal phlorotannins in temperate Australasia. Ecol. Monogr. 62(2):189–222.CrossRefGoogle Scholar
Steinberg, P. D., and de Nys, R. (2002). Chemical mediation of colonization of seaweed surfaces. J. Phycol. 38:621–9.CrossRefGoogle Scholar
Steinberg, P. D., de Nys, R., and Kjelleberg, S. (2001). Chemical mediation of surface colonization. In McClintock, J. B. and Baker, B. J. (eds), Marine Chemical Ecology (pp. 355–87). Boca Raton, FL: CRC Press.Google Scholar
Steinberg, P. D., de Nys, R., and Kjelleberg, S. (2002). Chemical cues for surface colonization. J. Chem. Ecol. 28:1935–51.CrossRefGoogle ScholarPubMed
Steinhoff, F. S., Wiencke, C., Müller, R., and Bischof, K. (2008). Effects of ultraviolet radiation and temperature on the ultrastructure of zoospores of the brown macroalgaLaminaria hyperborea. Plant Biol. 10: 388–97.CrossRefGoogle ScholarPubMed
Stekoll, M. S., and Deysher, L. (2000). Response of the dominant alga Fucus garneri (Silva) (Phaeophyceae) to the Exxon Valdez oil spill and clean-up. Mar. Poll. Bull. 40:1028–41.CrossRefGoogle Scholar
Steneck, R. S. (1982). A limpet-coralline alga association: adaptations and defenses between a selective herbivore and its prey. Ecology 63:507–22.CrossRefGoogle Scholar
Steneck, R. S. (1992). Plant-herbivore coevolution: a reappraisal from the marine realm and its fossil record. In John, D. M., Hawkins, S. J., and Price, J. H. (eds), Plant–Animal Interactions in the Marine Benthos, Systematics Association Special Volume 46 (pp. 477–91). Oxford: Claredon Press.Google Scholar
Steneck, R. S., and Dethier, M. N. (1994). A functional group approach to the structure of algal-dominated communities. Oikos 69:476–98.CrossRefGoogle Scholar
Steneck, R. S., and Martone, P. T. (2007). Calcified algae. In Denny, M.W. and Gaines, S.D. (eds) Encyclopedia of Tidepools (pp. 21–4) Berkeley, CA: University of California Press.Google Scholar
Steneck, R. S., and Watling, L. (1982). Feeding capabilities and limitation of herbivorous molluscs: a functional approach. Mar. Biol. 68:299–312.CrossRefGoogle Scholar
Steneck, R. S., Hacker, S. D., and Dethier, M. N. (1991). Mechanisms of competitive dominance between crustose coralline algae: an herbivore-mediated competitive reversal. Ecology 72:938–50.CrossRefGoogle Scholar
Steneck, R. S., Graham, M. H., Bourque, B. J., et al. (2002). Kelp forest ecosystems: Biodiversity, stability, resilience and future. Envir. Conserv. 29:436–59. [, , ]CrossRefGoogle Scholar
Stengel, D. B., and Dring, M. J. (1997). Morphology and in situ growth rates of plants of Ascophyllum nodosum (Phaeophyta) from different shore levels and responses of plants to vertical transplantation. Eur. J. Phycol. 32:193–202.CrossRefGoogle Scholar
Stengel, D. B., and Dring, M. J. (2000). Copper and iron concentrations in Ascophyllum nodosum (Fucales, Phaeophyta) from different sites in Ireland and after culture experiments in relation to thallus age and epiphytism. J. Exp. Mar. Biol. Ecol. 246:145–61.CrossRefGoogle ScholarPubMed
Stengel, D. B., Macken, A., Morrison, L., and Morely, N. (2004). Zinc concentrations in marine macroalgae and a lichen from western Ireland in relation to phylogenetic grouping, habitat and morphology. Mar. Poll. Bull. 48:902–9.CrossRefGoogle Scholar
Stepanyan, O. V., and Voskoboinikov, G. M. (2006). Effect of oil and oil products on morphofunctional parameters of marine macrophytes. Russian J. Mar. Biol. 32:S32–9.CrossRefGoogle Scholar
Stephenson, T. A., and Stephenson, A. (1949). The universal features of zonation between tide-marks on rocky coasts. J. Ecol. 38:289–305.CrossRefGoogle Scholar
Stephenson, T. A., and Stephenson, A. (1972). Life Between Tidemarks on Rocky Shores. San Francisco, CA: Freeman, 425 pp. [, , ]Google Scholar
Stevens, C. L., and Hurd, C. L. (1997). Boundary-layers around bladed aquatic macrophytes. Hydrobiologia 346:119–28.CrossRefGoogle Scholar
Stevens, C. L., Hurd, C. L., and Smith, M. J. (2001). Water motion relative to subtidal kelp fronds. Limnol. Oceanogr. 46:669–78.CrossRefGoogle Scholar
Stevens, C. L., Hurd, C. L., and Smith, M. J. (2002). Field measurement of the dynamics of the bull kelp Durvillaea antarctica (Chamisso) Heriot. J. Exp. Mar. Biol. Ecol. 269:147–71.CrossRefGoogle Scholar
Stevens, C. L., Hurd, C. L., and Isachsen, P. E. (2003). Modelling of diffusion boundary-layers in subtidal macroalgal canopies: the response to waves and currents. Aquat. Sci. 65:81–91.CrossRefGoogle Scholar
Stevens, C. L., Taylor, D. I., Delaux, S., Smith, M. J., and Schiel, D. R. (2008). Characterisation of wave-influenced macroalgal propagule settlement. J. Marine Syst. 74:96–107.CrossRefGoogle Scholar
Stewart, H. L. (2004). Hydrodynamic consequences of maintaining an upright posture by different magnitudes of stiffness and buoyancy in the tropical alga Turbinaria ornata. J. Marine Syst. 49:157–67.CrossRefGoogle Scholar
Stewart, H. L. (2006). Morphological variation and phenotypic plasticity of buoyancy in the macroalga Turbinaria ornata across a barrier reef. Mar. Biol. 149:721–30.CrossRefGoogle Scholar
Stewart, H. L., and Carpenter, R. C. (2003). The effects of morphology and water flow on photosynthesis of marine macroalgae. Ecology 84:2999–3012.CrossRefGoogle Scholar
Stewart, J. G. (1982). Anchor species and epiphytes in intertidal algal turf. Pac. Sci. 36:45–59.Google Scholar
Stewart, J. G. (1983). Fluctuations in the quantity of sediments trapped among algal thalli on intertidal rock platforms in southern California. J. Exp. Mar. Biol. Ecol. 73:205–11.CrossRefGoogle Scholar
Stewart, J. G. (1989). Establishment, persistence and dominance of Corallina (Rhodophyta) in algal turf. J. Phycol. 25:436–46.CrossRefGoogle Scholar
Stimson, J., Larned, S. T., and Conklin, E. (2001). Effects of herbivory, nutrient levels, and introduced algae on the distribution and abundance of the invasive macroalga Dictyosphaeria cavernosa in Kaneohe Bay, Hawaii. Coral Reefs 19:343–57.CrossRefGoogle Scholar
Stirk, W. A., Novák, O., Strnad, M., and van Staden, J. (2003). Cytokinins in macroalgae. Plant Growth Regul. 41:13–24.CrossRefGoogle Scholar
Stirk, W. A., Novák, O., Hradecká, V., et al. (2009). Endogenous cytokinins, auxins and abscisic acid in Ulva fasciata (Chlorophyta) and Dictyota humifusa (Phaeophyta): towards understanding their biosynthesis and homeostasis. Eur. J. Phycol. 44:231–40.CrossRefGoogle Scholar
Storz, H., Müller, K., Ehrhart, F., et al. (2009) Physicochemical features of ultra-high viscosity alginates. Carbohydrate Res. 344:985–95.CrossRefGoogle ScholarPubMed
Stratmann, J., Paputsoglu, G., and Oertel, W. (1996). Differentiation of Ulva mutabilis (Chlorophyta) gametangia and gamete release are controlled by extracellular inhibitors. J. Phycol. 32:1009–21.CrossRefGoogle Scholar
Strickland, J. D. H., and Parsons, T. R. (1972). A Practical Handbook of Seawater Analysis, 2nd edn. Fish Res. Bd. Canada Bull. 167, 310 pp.
Strömgren, T. (1979). The effect of zinc on the increase in length of five species of intertidal Fucales. J. Exp. Mar. Biol. Ecol. 40:95–102. [.]CrossRefGoogle Scholar
Strömgren, T. (l980a). The effect of dissolved copper on the increase in length of four species of intertidal fucoid algae. Mar. Environ. Res. 3:5–13.CrossRefGoogle Scholar
Strömgren, T. (l980b). The effect of lead, cadmium, and mercury on the increase in length of five intertidal Fucales. J. Exp. Mar. Biol. Ecol. 43:107–19.CrossRefGoogle Scholar
Strömgren, T. (1980c). Combined effects of Cu, Zn, and Hg on the increase in length of Ascophyllum nodosum (L.) Le Jolis. J. Exp. Mar. Biol. Ecol. 48:225–31.CrossRefGoogle Scholar
Su, H.-N., Xie, B.-B., Zhang, X.-Y., Zhou, B.-C., and Zhang, Y-Z. (2010). The supramolecular architecture, function, and regulation of thylakoid membranes in red algae: an overview. Photosynth. Res. 106:73–87.CrossRefGoogle ScholarPubMed
Suetsugu, N., Mittmann, F., Wagner, G., Hughes, J., and Wada, M. (2005). A chimeric photoreceptor gene, N EOCHROME, has arisen twice during plant evolution. P. Natl. Acad. Sci. USA. 102:13,705–9.CrossRefGoogle Scholar
Sunda, W. G. (2009). Trace element nutrients. In Steele, J. H. (ed.), Encyclopedia of Ocean Science (2nd edn), (pp. 75–86). New York: Springer.Google Scholar
Suple, C. (1999). El Niño/La Niña. Nat. Geog. 4 (3):74–95.Google Scholar
Surif, M. B., and Raven, J. A. (1989). Exogenous inorganic carbon sources for photosynthesis in seawater by members of the Fucales and the Laminariales (Phaeophyta): ecological and taxonomic implications. Oecologia 78:97–105.CrossRefGoogle ScholarPubMed
Surif, M. B., and Raven, J. A. (1990). Photosynthetic gas exchange under emersed conditions in eulittoral and normally submersed members of the Fucales and Laminariales: interpretation in relation to C isotope ratio and N and water use efficiency. Oecologia 82:68–80.CrossRefGoogle Scholar
Sussmann, A. V., and De Wreede, R. E. (2002). Host specificity of the endophytic sporophyte phase of Acrosiphonia (Codiolales, Chlorophyta) in southern British Columbia, Canada. Phycologia 41:169–77.CrossRefGoogle Scholar
Sutherland, J. E., Lindstrom, S. C., Nelson, W. A., et al. (2011). A new look at an ancient order: generic revision of the Bangiales (Rhodophyta). J. Phycol. 47:1131–51. [, , ]CrossRefGoogle Scholar
Suto, S. (1950). Studies on shedding, swimming and fixing of the spores of seaweeds. Bull. Jpn. Soc. Sci. Fish. 16:1–9.CrossRefGoogle Scholar
Suttle, C.A. (2007). Marine viruses–major players in the global ecosystem. Nature Reviews 5: 801–12.Google ScholarPubMed
Suzuki, K., Iwamoto, K., Yokoyama, S., and Ikawa, T. (1991). Glycolate-oxidizing enzymes in algae. J. Phycol. 27:492–8.CrossRefGoogle Scholar
Suzuki, Y., Kuma, K., Kudo, I., and Matsunaga, K. (1995). Iron requirement of the brown macroalgae, Laminaria japonica, Undaria pinnatifida (Phaeophyta) and the crustose coralline alga Lithophyllum yessoense (Rhodophyta), and their competition in the northern Japan Sea. Phycologia 34:201–5.CrossRefGoogle Scholar
Svendsen, H., Beszczynska-Møller, A., Hagen, J. O., et al. (2002). The physical environment of Kongsfjorden–Krossfjorden, an Arctic fjord system in Svalbard. Polar Res. 21:133–66.Google Scholar
Svensson, J. R., Lindegarth, M., Siccha, M., et al. (2007). Maximum species richness at intermediate frequencies of disturbance: consistency among levels of productivity. Ecology 88:830–8.CrossRefGoogle ScholarPubMed
Svensson, J. R., Lindegarth, M., and Pavia, H. (2010). Physical and biological disturbances interact differently with productivity: effects on floral and faunal richness. Ecology 91:3069–80.CrossRefGoogle ScholarPubMed
Svensson, J. R., Lindegarth, M., Jonsson, P. R., and Pavia, H. (2012). Disturbance-diversity models: what do they really predict and how are they tested? P. Roy. Soc. Lond. B. Bio. 279:2163–70.CrossRefGoogle ScholarPubMed
Swanson, A. K., and Druehl, L. D. (2002). Induction, exudation and the UV protective role of kelp phlorotannins. Aquat. Bot. 73:241–53.CrossRefGoogle Scholar
Sweeney, B. M. (1974). A physiological model for circadian rhythms derived from the Acetabularia rhythm paradoxes. Int. J. Chronobiol. 2: 25–33.Google ScholarPubMed
Sweeney, B. M., and Prezelin, B. B. (1978). Circadian rhythms. Photochem. Photobiol. 27: 841–7.CrossRefGoogle ScholarPubMed
Swinbanks, D. D. (1982). Intertidal exposure zones: a way to subdivide the shore. J. Exp. Mar. Biol. Ecol. 62:69–86.CrossRefGoogle Scholar
Syrett, P. J. (1981). Nitrogen metabolism of microalgae. Can. Bull. Fish. Aquat. Sci. 210:182–210.Google Scholar
Szymanski, D. B., and Cosgrove, D. J. (2009). Dynamic coordination of cytoskeletal and cell wall systems during plant cell morphogenesis. Curr. Biol. 19:R800–11.CrossRefGoogle ScholarPubMed
Tabita, F. R., Satagopan, S., Hanson, T. E., Kreel, N. E., and Scott, S. S. (2008) Distinct form I, II, III, and IV Rubisco proteins from the three kingdoms of life provide clues about Rubisco evolution and structure/function relationships. J. Exp. Bot. 59:1515–24.CrossRefGoogle Scholar
Taiz, L., and Zeiger, E. (eds) (2010). Plant Physiology (5th edn). Sunderland, MA: Sinauer Assoc., Inc., 782 pp. [, , ]Google Scholar
Takahashi, F., Yamagata, D., Ishikawa, M., et al. (2007). AUREOCHROME, a photoreceptor required for photomorphogenesis in stramenopiles. PNAS 104:19,625–30.CrossRefGoogle ScholarPubMed
Talarico, L. (1990). R-phycoerythrin from Audouinella saviana (Nemaliales, Rhodophyta). Ultrastructural and biochemical analysis of aggregates and subunits. Phycologia 29:292–302.CrossRefGoogle Scholar
Talarico, L., and Maranzana, G. (2000). Light and adaptive responses in red macroalgae: an overview. J. Photochem. Photobiol. B: Biol. 56: 1–11.CrossRefGoogle ScholarPubMed
Tanaka, A., Nagasato, C., Uwai, S., Motomura, T., and Kawai, H. (2007). Re-examination of ultrastructures of the stellate chloroplast organization in brown algae: structure and development of pyrenoids. Phycol. Res. 55:203–13.CrossRefGoogle Scholar
Tang, Y. Z., and Gobler, C. J. (2011). The green macroalga, Ulva lactuca, inhibits the growth of seven common harmful algal bloom species via allelopathy. Harmful Algae 10:480–8.CrossRefGoogle Scholar
Tanner, C. E. (1986). Investigations of the taxonomy and morphological variation of Ulva (Chlorophyta): Ulva californica Wille. Phycologia 25:510–20.CrossRefGoogle Scholar
Tarakhovskaya, E. R., Maslov, Yu. I., and Shishova, M. F. (2007). Phytohormones in algae. Russ. J. Plant Physl. 54:163–70.CrossRefGoogle Scholar
Targett, N. M., and Arnold, T. M. (2001). Effects of secondary metabolites on digestion in marine herbivores. In McClintock, J. B., and Baker, B. J., (eds), Marine Chemical Ecology (pp. 391–411). Boca Raton, FL: CRC Press.CrossRefGoogle Scholar
Tatewaki, M. (1970). Culture studies on the life history of some species of the genusMonostroma. Sci. Pap. Inst. Algol. Res., Fac. Sci., Hokkaido U. 6(1):1–56.Google Scholar
Tatewaki, M., Provasoli, L., and Pintner, I. J. (1983). Morphogenesis of Monostroma oxyspermum (Kiitz.) Doty (Chlorophyceae) in axenic culture, especially in bialgal culture. J. Phycol. 19:409–16.CrossRefGoogle Scholar
Taylor, D. I., and Schiel, D. R. (2010). Algal populations controlled by fish herbivory across a wave exposure gradient on southern temperate shores. Ecology 91:201–11.CrossRefGoogle ScholarPubMed
Taylor, D., Delaux, S., Stevens, C., Nokes, R., and Schiel, D. (2010). Settlement rates of macroalgal propagules: cross-species comparisons in a turbulent environment. Limnol. Oceanogr. 55:66–76.CrossRefGoogle Scholar
Taylor, M. W.Taylor, R. B., and Rees, T.A. V. (1999). Allometric evidence for the dominant role of surface cells in ammonium metabolism and photosynthesis in northeastern New Zealand seaweeds. Mar. Ecol. Prog. Ser. 184:73–81. [, , ]CrossRefGoogle Scholar
Taylor, M. W., and Rees, T. A. V. (1999). Kinetics of ammonium assimilation in two seaweeds, Enteromorpha sp. (Chlorophyta) and Osmundaria colensoi (Rhodophyceae). J. Phycol. 35:740–6.CrossRefGoogle Scholar
Taylor, M. W., Barr, N. G., Grant, C. M., and Rees, T. A. V. (2006). Changes in amino acid composition of Ulva intestinalis (Chlorophyceae) following addition of ammonium or nitrate. Phycologia 45:270–6.CrossRefGoogle Scholar
Taylor, P. R., and Littler, M. M. (1982). The roles of compensatory mortality, physical disturbance, and substrate retention in the development and organization of a sand-influenced rocky intertidal community. Ecology 63: 135–46.CrossRefGoogle Scholar
Taylor, R. B., and Rees, T. A. V. (1998). Excretory products of mobile epifauna as a nitrogen source for seaweeds. Limnol. Oceanogr. 43:600–6.CrossRefGoogle Scholar
Taylor, R. B., Peek, J. T. A., and Rees, T. A. V. (1998). Scaling of ammonium uptake by seaweeds to surface area:volume ratio: geographical variation and the role of uptake by passive diffusion. Mar. Ecol. Prog. Ser. 169:143–8. [, , ]CrossRefGoogle Scholar
Taylor, R. B., Sotka, E., and Hay, M. E. (2002). Tissue-specific induction of herbivore resistance: seaweed response to amphipod grazing. Oecologia 132: 68–76.CrossRefGoogle ScholarPubMed
Taylor, W. R. (1957). Marine Algae of the Northeastern Coast of North America. Ann Arbor, MI: University of Michigan Press.Google Scholar
Taylor, W. R. (1960). Marine Algae of the Eastern Tropical and Subtropical Coasts of the Americas. Ann Arbor, MI: University of Michigan Press.Google Scholar
Tegner, M. J., and Dayton, P. K. (1987). El Niño effects on Southern California kelp forest communities. Adv. Ecol. Res. 17:243–79.CrossRefGoogle Scholar
Teichberg, M., Fox, S., Aguila, C., Olsen, Y., and Valiela, I. (2008). Macroalgal responses to experimental nutrient enrichment in shallow coastal waters: growth, internal nutrient pools, and isotopic signatures. Mar. Ecol. Prog. Ser. 368:117–26.CrossRefGoogle Scholar
Teichberg, M., Fox, S., Aguila, C., et al. (2010). Eutrophication and macroalgal blooms in temperate and tropical coastal waters: nutrient enrichment experiments with Ulva spp. Global Change Biol. 16:2624–37.Google Scholar
Telfer, A. (2002). What is β-carotene doing in the photosystem II reaction centre? Phil. Trans. Roy. Soc. London B Biol. Sci. 357:1431–40.CrossRefGoogle ScholarPubMed
Terumoto, I. (1964). Frost resistance in some marine algae from the winter intertidal zone. Low Temp. Sci. (Ser. B) 22:19–28.Google Scholar
Tewari, A., and Joshi, H. V. (1988). Effect of domestic sewage and industrial effluents on biomass and species diversity of seaweeds. Bot. Mar. 31:389–97.CrossRefGoogle Scholar
The Royal Society (2005). Ocean acidification due to increasing atmospheric CO2. Policy Document 12/05. The Royal Society, London.Google Scholar
Thiel, M., and Gutow, L. (2004). The ecology of rafting in the marine environment. I. The floating substrata. Oceanogr. Mar. Biol.: Ann. Rev. 42:181–264.CrossRefGoogle Scholar
Thiel, M., and Gutow, L. (2005). The ecology of rafting in the marine environment. II. The rafting organisms and community. Oceanogr. Mar. Biol. 43:279–418.CrossRefGoogle Scholar
Thiel, M., Macaya, E. C., Acuña, E., et al. (2007). The Humboldt current system of northern-central Chile: oceanographic processes, ecological interactions and socio-economic feedback. Oceanogr. Mar. Biol. Ann. Rev. 45:195–345.Google Scholar
Thomas, D. N., and Wiencke, C. (1991). Photosynthesis, dark respiration and light independent carbon fixation of endemic Antarctic macroalgae. Polar Biol. 11:329–37.CrossRefGoogle Scholar
Thomas, D. N., Fogg, G. E., Convey, P., et al. (2008). The Biology of Polar Regions. Oxford: Oxford University Press, 416 pp.CrossRefGoogle Scholar
Thomas, F., Cosse, A., Goulitquer, S., et al. (2011). Waterborne signaling primes the expression of elicitor-induced genes and buffers the oxidative responses in the brown alga Laminaria digitata. PloS ONE 6:e21475.CrossRefGoogle ScholarPubMed
Thomas, M. L. H. (1986). A physically derived exposure index for marine shorelines. Ophelia 25:1–13. [8.14]CrossRefGoogle Scholar
Thomas, T. E., and Harrison, P. J. (1985). Effects of nitrogen supply on nitrogen uptake, accumulation and assimilation in Porphyra perforata (Rhodophyta). Mar. Biol. 85:269–78.CrossRefGoogle Scholar
Thomas, T. E., and Harrison, P. J. (1987). Rapid ammonium uptake and field conditions. J. Exp. Mar. Biol. Ecol. 107:1–8.CrossRefGoogle Scholar
Thomas, T. E., and Harrison, P. J. (1988). A comparison of in vitro and in vivo nitrate reductase assays in three intertidal seaweeds. Bot Mar. 31:101–7.CrossRefGoogle Scholar
Thomas, T. E., and Turpin, D. H. (1980). Desiccation enhanced nutrient uptake rates in the intertidal alga Fucus distichus. Bot. Mar. 23:479–81.Google Scholar
Thomas, T. E., Harrison, P. J., and Taylor, E. B. (1985). Nitrogen uptake and growth of the germlings and mature thalli of Fucus distichus. Mar. Biol. 84:267–74.CrossRefGoogle Scholar
Thomas, T. E., Harrison, P. J., and Turpin, D. H. (1987a). Adaptations of Gracilaria pacifica (Rhodophyta) to nitrogen procurement at different intertidal locations. Mar. Biol. 93:569–80.CrossRefGoogle Scholar
Thomas, T. E., Turpin, D. H., and Harrison, P. J. (l987b). Desiccation enhanced nitrogen uptake rates in intertidal seaweeds. Mar. Biol. 94:293–8.CrossRefGoogle Scholar
Thompson, S. E. M., Callow, J. A., Callow, M. E., et al. (2007). Membrane recycling and calcium dynamics during settlement and adhesion of zoospores of the green alga Ulva linza. Plant Cell Environ. 30:733–44.CrossRefGoogle ScholarPubMed
Thompson, S. E. M., Callow, M. E., and Callow, J. A. (2010). The effects of nitric oxide in settlement and adhesion of zoospores of the green alga Ulva. Biofouling 26:167–78.CrossRefGoogle ScholarPubMed
Thomsen, M. S., Wemberg, T., and Kendrick, G. A. (2004). The effect of thallus size, life stage, aggregation, wave exposure and substratum conditions on the forces required to break or dislodge the small kelp Ecklonia radiata. Bot. Mar. 47:454–60.CrossRefGoogle Scholar
Thomsen, M. S., Wernberg, T., Altieri, A., et al. (2010). Habitat cascades: the conceptual context and global relevance of facilitation cascades via habitat formation and modification. Integr. Compar. Biol. 50:158–75.CrossRefGoogle ScholarPubMed
Thornber, C. S. (2006). Functional properties of the isomorphic biphasic algal life cycle. Integr. Comp. Biol. 46:605–14.CrossRefGoogle ScholarPubMed
Thornber, C. S., and Gaines, S. D. (2004). Population demographics in species with biphasic life cycles. Ecology 85:1661–74.CrossRefGoogle Scholar
Thornber, C., Stachowicz, J. J., and Gaines, S.. (2006). Tissue type matters: selective herbivory on different life history stages of an isomorphic alga. Ecology 87:2255–63.CrossRefGoogle ScholarPubMed
Thurman, H. V., and Trujillo, A. P. (eds) (2004). Introductory Oceanography (10th edition). 608 pp. Upper Saddle River, NJ: Prentice Hall. [8.0]Google Scholar
Thursby, G. (1984). Development of toxicity test procedures for the marine algaChampia parvula. USEPA 60019844.Google Scholar
Titlyanov, E. A. (1976). Adaptation of benthic plants to light. Role of light in distribution of attached marine algae. Biol. Morya 1:3–12.Google Scholar
Titlyanov, E. A., and Titlyanova, T. V. (2010). Seaweed cultivation: methods and problems. Russian J. Mar. Biol. 36:227–42.CrossRefGoogle Scholar
Togashi, T., and Cox, P. A. (2001). Tidal-linked synchrony of gamete release in the marine green alga, Monostroma angicava Kjellman. J. Exp. Mar. Biol. Ecol. 264:117–31.CrossRefGoogle Scholar
Togashi, T., Nagisa, M., Miyazaki, T., et al. (2006). Gamete behaviours and the evolution of “marked anisogamy”: reproductive strategies and sexual dimorphism in Bryopsidales marine green algae. Evol. Ecol. Res. 8:617–28.Google Scholar
tom Dieck (Bartsch), I. (1991). Circannual growth rhythm and photoperiodic sorus induction in the kelp Laminaria setchellii (Phaeophyta). J. Phycol. 27:341–50.Google Scholar
Toohey, B. D., and Kendrick, G. A. (2007). Survival of juvenile Ecklonia radiata sporophytes after canopy loss. J. Exp. Mar. Biol. Ecol. 349:170–82.CrossRefGoogle Scholar
Toohey, B. D., Kendrick, G. A., and Harvey, E. S. (2007). Disturbance and reef topography maintain high local diversity in Ecklonia radiata kelp forests. Oikos 116:1618–30.CrossRefGoogle Scholar
Toole, C. M., and Allnutt, F. C. T. (2003). Red, cryptomonade and glaucocystophyte algal phycobiliproteins. In Larkum, A. W. D., Douglas, S. E., and Raven, J. A. (eds) Photosynthesis in Algae. Advances in Photosynthesis and Respiration, Vol. 14. (pp. 305–34) Dordrecht, The Netherlands: Kluwer Academic Publishers.CrossRefGoogle Scholar
Topinka, J. A. (1978). Nitrogen uptake by Fucus spiralis (Phaeophyceae). J. Phycol. 14:241–7.CrossRefGoogle Scholar
Torres, J., Rivera, A., Clark, G., and Roux, S. J. (2008). Participation of extracellular nucleotides in the wound response of Dasycladus vermicularis and Acetabularia acetabulum (Dasycladales, Chlorophyta). J. Phycol. 44:1504–11.CrossRefGoogle Scholar
Toth, G. B., and Pavia, H. (2000a). Water-borne cues induce chemical defense in a marine alga (Ascophyllum nodosum). P. Nat. Acad. Sci. USA 97:14,418–20.CrossRefGoogle Scholar
Toth, G., and Pavia, H. (2000b). Lack of phlorotannin induction in the brown seaweed Ascophyllum nodosum in response to increased copper concentrations. Mar. Ecol. Prog. Ser. 192:119–26.CrossRefGoogle Scholar
Toth, G. B., and Pavia, H. (2007). Induced herbivore resistance in seaweeds: a meta-analysis. J. Ecol. 95:425–34.CrossRefGoogle Scholar
Toth, G. B., Langhamer, O., and Pavia, H. (2005). Inducible and constitutive defenses of valuable seaweed tissues: consequences for herbivore fitness. Ecology 86:612–18.CrossRefGoogle Scholar
Trautman, D. A., Hinde, R., and Borowitzka, M. A. (2000). Population dynamics of an association between a coral reef sponge and a red macroalga. J. Exp. Mar. Biol. Ecol. 244:87–105.CrossRefGoogle Scholar
Troell, M. (2009). Integrated marine and brackish water aquaculture in tropical regions: research implementation and prospects. In Soto, D. (ed.) Integrated Mariculture: A Global Perspective. FAO Fisheries and Aquaculture Technical Paper No. 529:47–131.Google Scholar
Troell, M., Rönnbäck, P., Halling, C., Kautsky, N., and Buschmann, A. (1999). Ecological engineering in aquaculture: use of seaweeds for removing nutrients from intensive mariculture. J. Appl. Phycol. 11:89–97.CrossRefGoogle Scholar
Troell, M., Halling, C., Neori, A., et al. (2003). Integrated mariculture: asking the right questions. Aquaculture 226:69–90.CrossRefGoogle Scholar
Troell, M., Joyce, A., Chopin, T., et al. (2009). Ecological engineering in aquaculture: potential for integrated multi-trophic aquaculture (IMTA) in marine offshore systems. Aquaculture 297:1–9.CrossRefGoogle Scholar
Trowbridge, C. D., and Todd, C. D. (2001). Host-plant change in marine specialist herbivores: Ascoglossan sea slugs on introduced macroalgae. Ecol. Monogr. 71:219–43.CrossRefGoogle Scholar
Trujillo, A. P., and Thurman, H. V. (2010). Essentials of Oceanography. Upper Saddle River, NJ: Prentice Hall, 551 pp.Google Scholar
Tsekos, I. (1999). The sites of cellulose synthesis in algae: diversity and evolution of cellulose-synthesizing enzyme complexes. J. Phycol. 35:635–55.CrossRefGoogle Scholar
Tsekos, I., and Reiss, H.-D. (1994). Tip cell growth and the frequency and distribution of cellulose microfibril-synthesizing complexes in the plasma membrane of apical shoot cells of the red alga Porphyra yezoensis. J. Phycol. 30:300–10.CrossRefGoogle Scholar
Tseng, C. K. (1981). Commercial cultivation. In Lobban, C. S. and Wynne, M. J. (eds), The Biology of Seaweeds (pp. 680–725). Oxford: Blackwell Scientific. [, , , , , , , ]Google Scholar
Tseng, C. K. (l987). Laminaria mariculture in China. In Doty, M. S., Caddy, J. F., and Santelices, B. (eds), Case Studies of Seven Commercial Seaweed Resources (pp. 239–63). FAO Fish. Tech. Pap. 281. [, , ]Google Scholar
Tsuda, R. T. (1965). Marine algae from Laysan Island with additional notes on the vascular flora. Atoll Res. Bull. 110, 31 pp.CrossRefGoogle Scholar
Tsuda, R. T., Larson, H. K., and Lujan, R. J. (1972). Algal growth on beaks of live parrotfishes. Pac. Sci. 26:20–3.Google Scholar
Tucker, J., Sheats, N., Giblin, A. E., Hopkinson, C. S., and Montoya, J. P (1999). Using stable isotopes to trace sewage-derived material through Boston Harbor and Massachusetts Bay. Mar. Environ. Res. 48:353–75.CrossRefGoogle Scholar
Turner, A. (2010). Marine pollution from antifouling paint particles. Mar. Poll. Bull. 60:159–71.CrossRefGoogle ScholarPubMed
Turner, A., Pollock, H., and Brown, M. T. (2009). Accumulation of Cu and Zn from antifouling paint particles by the marine macroalga, Ulva lactuca. Environ. Poll. 157:2314–19.CrossRefGoogle Scholar
Turner, D. R., Whitfield, M., and Dickson, G. A. (1981). The equilibrium speciation of dissolved components in freshwater and seawater at 25ºC and I atm pressure. Geochim. Cosmochim. Acta 45:855–81.CrossRefGoogle Scholar
Turpin, D. H. (1980). Processes in nutrient based phytoplankton ecology. Ph.D. dissertation, University of British Columbia, Vancouver.
Turpin, D. H., and Huppe, H. C. (1994) Integration of carbon and nitrogen metabolism in plant and algal cells. Annu. Rev. Plant Physiol. Plant Mol. Biol. 45:577–607. [, , ]CrossRefGoogle Scholar
Turvey, J. R. (1978). Biochemistry of algal polysaccharides. Int. Rev. Biochem. 16: 151–77.Google Scholar
Tyler, A. C., and McGlathery, K. J. (2006). Uptake and release of nitrogen in the macroalgae Gracilaria vermiculophylla (Rhodophyta). J. Phycol. 42:515–25.CrossRefGoogle Scholar
Tyrrell, T. (2011). Anthropogenic modification of the oceans. Phil. Trans. Roy. Soc. A 309:887–908.CrossRefGoogle Scholar
Ueki, C., Nagasato, C., Motomura, T., and Saga, N. (2008). Reexamination of the pit plugs and the characteristic membranous structures in Porphyra yezoensis (Bangiales, Rhodophyta). Phycologia 47:5–11.CrossRefGoogle Scholar
Ueki, C., Nagasato, C., Motomura, T., and Saga, N. (2009). Ultrastructure of mitosis and cytokinesis during spermatogenesis in Porphyra yezoensis (Bangiales, Rhodophyta). Bot. Mar. 52:129–39.CrossRefGoogle Scholar
Uhrmacher, S., Hanelt, D., and Nultsch, W. (1995). Zeaxanthin content and the degree of photoinhibition are linearly correlated in the brown alga Dictyota dichotoma. Mar. Biol. 123:159–65.CrossRefGoogle Scholar
Umar, M. J., McCook, L. J., and Price, I. R. (1998). Effects of sediment deposition on the seaweed Sargassum on a fringing coral reef. Coral Reefs 17:169–77.CrossRefGoogle Scholar
Underwood, A. J. (1978). A refutation of critical tidal levels as determinants of the structure of intertidal communities on British shores. J. Exp. Mar. Biol. Ecol. 33:261–76.CrossRefGoogle Scholar
Underwood, A. J. (1980). The effects of grazing by gastropods and physical factors on the upper limits of distribution of intertidal macroalgae. Oecologia 46:201–13.CrossRefGoogle ScholarPubMed
Underwood, A. J. (1986). The analysis of competition by field experiments. In: Kikkawa, J. and Anderson, D. J. (eds) Community Ecology: Pattern and Processes (pp. 240–68). Oxford: Blackwell Scientific Publications.Google Scholar
Underwood, A.J. (1992). Beyond BACI: The detection of environmental impacts on population in the real, but variable world. J. Exp. Mar. Biol. Ecol. 161:145–78.CrossRefGoogle Scholar
Underwood, A.J. (1994). On beyond BACI: Sampling designs that might reliably detect environmental disturbances. Ecol. Appl. 4:3–15.CrossRefGoogle Scholar
Underwood, A. J. (1997). Experiments in Ecology: Their Logical Design and Interpretation Using Analysis of Variance. Cambridge: Cambridge University Press, 504 pp.Google Scholar
Underwood, A. J. (2006). Why overgrowth of intertidal encrusting algae does not always cause competitive exclusion. J. Exp. Mar. Biol. Ecol. 330:448–54.CrossRefGoogle Scholar
Underwood, A. J., and Denley, E. J. (1984). Paradigms, explanations, and generalizations in models for the structure of intertidal communities on rocky shores. In Strong, J. D.R., Simberloff, D., Abele, L. G., and Thistle, A. B., (eds). Ecological Communities: Conceptual Issues and the Evidence (pp. 151–80). Princeton, NJ: Princeton University Press.Google Scholar
Underwood, A. J., and Keough, M. J. (2001). Supply-side ecology: the nature and consequences of variations in recruitment of intertidal organisms. In Bertness, M. D., Gaines, S. D., and Hay, M. E. (eds), Marine Community Ecology (pp. 183–200). Sunderland, MA: Sinauer Assocs.Google Scholar
Underwood, A. J., and Jernakoff, P. (1981). Effects of interactions between algae and grazing gastropods on the structure of a low-shore intertidal algal community. Oecologia 48:221–33.CrossRefGoogle ScholarPubMed
Uppalapati, S. R., and Fujita, Y. (2001). The relative resistance of Porphyra species (Bangiales, Rhodophyta) to infection by Pythium porphyrae (Peronosporales, Oomycota). Bot. Mar. 44:1–7.CrossRefGoogle Scholar
Uthicke, S., Schaffelke, B., and Byrne, M. (2009). A boom-bust phylum? Ecological and evolutionary consequences of density variations in echinoderms. Ecol. Monogr. 79:3–24.CrossRefGoogle Scholar
Vadas, R. L. (1977). Preferential feeding: an optimization strategy in sea urchins. Ecol. Monogr. 47:337–71.CrossRefGoogle Scholar
Vadas, R. L. (1985). Herbivory. In Littler, M. M. and Littler, D. S. (eds), Handbook of Phycological Methods: Ecological Field Methods: Macroalgae (pp. 531–72). Cambridge: Cambridge University Press.Google Scholar
Vadas, R. L., and Steneck, R. S. (1988). Zonation of deep-water benthic algae in the Gulf of Maine. J. Phycol. 24:338–46.CrossRefGoogle Scholar
Vadas, R. L., Keser, M., and Larson, B. (1978). Effects of reduced temperatures on previously stressed populations of an intertidal alga. In J. H. Thorp and J. W. Gibbons (eds), (pp. 431–51). DOE Symposium Series 48 (CONF-721114). Washington DC: U.S. Government Printing Office.
Vadas, R. L., Wright, W. A., and Miller, S. L. (1990). Recruitment of Ascophyllum nodosum: wave action as a source of mortality. Mar. Ecol. Prog. Ser. 61:263–72.CrossRefGoogle Scholar
Vadas, R. L.., Johnson, S., and Norton, T. A. (1992). Recruitment and mortality of early post-settlement stages of benthic algae. Brit. Phycol. J. 27:331–51.CrossRefGoogle Scholar
Valdivia, N., Scrosati, R. A., Molis, M., and Knox, A. S. (2011). Variation in community structure across vertical intertidal stress gradients: how does it compare with horizontal variation at different scales? PLoS ONE 6: e24062, doi:.CrossRefGoogle ScholarPubMed
Valentine, J. P., and Johnson, C. R. (2004). Establishment of the introduced kelp Undaria pinnatifida following dieback of the native macroalga Phyllospora comosa in Tasmania, Australia. Mar. Freshwater Res. 55:223–30.CrossRefGoogle Scholar
Valiela, I., and Cole, M. L. (2002). Comparative evidence that salt marshes and mangroves may protect seagrass meadows from land-derived nitrogen loads. Ecosystems 5:92–102.CrossRefGoogle Scholar
Valiella, I., McClelland, J., Hauxwell, J., et al. (1997). Macroalgal blooms in shallow estuaries: Controls and ecophysiological and ecosystem consequences. Limnol. Oceanogr. 42:1105–18.CrossRefGoogle Scholar
van Alystyne, K. L., Ehlig, J. M., and Whitman, S. L. (1999). Feeding preferences for juvenile and adult algae depend on algal stage and herbivore species. Mar. Ecol. Prog. Ser. 180:179–85.CrossRefGoogle Scholar
van Alstyne, K. L., Whitman, S. L., and Ehlig, J. M. (2001). Differences in herbivore preferences, phlorotannin production, and nutritional quality between juvenile and adult tissues from marine brown algae. Mar. Biol. 139:201–10.CrossRefGoogle Scholar
Van Assche, F., and Clijsters, H. (1990). Effects of metals on enzyme activity in plants. Plant Cell Environ. 13:195–206.CrossRefGoogle Scholar
van de Poll, W. H., Eggert, A., Buma, A. G. J., and Breeman, A. M. (2001). Effects of UV-B-induced DNA damage and photoinhibition on growth of temperate marine red macrophytes: Habitat-related differences in UV-B tolerance. J. Phycol. 37:30–7.CrossRefGoogle Scholar
van den Hoek, C. (1982). Phytogeographic distribution groups of benthic marine algae in the North Atlantic Ocean. A review of experimental evidence from life history studies. Helgol. Meeresunters 35:153–214.CrossRefGoogle Scholar
van den Hoek, C., Stam, W. T, and Olsen, J. L. (1988). The emergence of a new chlorophytan system, and Dr. Kornmann’s contribution thereto. Helgol. Meeresunters 42:339–83. [,]CrossRefGoogle Scholar
van den Hoek, C., Mann, D. G., and Jahns, H. M. (1995). Algae: An Introduction to Phycology. Cambridge: Cambridge University Press, 627 pp. [, , , , , , , ]Google Scholar
van der Meer, J. P. (1978). Genetics of Gracilaria sp. (Rhodophyceae, Gigartinales). III. Non-Mendelian gene transmission. Phycologia 17:314–18.CrossRefGoogle Scholar
van der Meer, J. P. (1986a). Genetic contributions to research on seaweeds. Prog. Phycol. Res. 4:1–38.Google Scholar
van der Meer, J. P. (l986b). Genetics of Gracilaria tikvahiae (Rhodophyceae). XI. Further characterization of a bisexual mutant. J. Phycol. 22:151–8.Google Scholar
van der Meer, J. P. (1990). Genetics. In Cole, K. M. and Sheath, R. G. (eds), Biology of the Red Algae (pp. 103–21). Cambridge: Cambridge University Press.Google Scholar
van der Meer, J. P., and Bird, N. L. (1977). Genetics of Gracilaria sp. (Rhodophyceae, Gigartinales). I. Mendelian inheritance of two spontaneous green variants. Phycologia 16: 159–61.CrossRefGoogle Scholar
van der Meer, J. P., and Todd, E. R. (1977). Genetics of Gracilaria sp. (Rhodophyceae, Gigartinales). IV. Mitotic recombination and its relationship to mixed phases in the life history. Can. J. Bot. 55:2810–17.CrossRefGoogle Scholar
van der Meer, J. P., and Todd, E. R. (1980). The life history of Palmaria palmata in culture. A new type for the Rhodophyta. Can. J. Bot. 58:1250–6. [, , ]CrossRefGoogle Scholar
van der Meer, J. P., and Zhang, X. (1988). Similar unstable mutations in three species of Gracilaria (Rhodophyta). J. Phycol. 24:198–202.Google Scholar
van der Meer, J. P, Patwary, M. U., and Bird, C. J. (1984). Genetics of Gracilaria tikvahiae (Rhodophyceae). X. Studies on a bisexual clone. J. Phycol. 20:42–6.CrossRefGoogle Scholar
van der Strate, H., Boele-Bos, S., Olsen, J., van de Zande, L., and Stam, W. (2002a). Phylogeographic studies in the tropical seaweed Chladophoropsis membranacea (Chlorophyta, Ulvophyceae) reveal a cryptic species complex. J. Phycol. 38:572–82.CrossRefGoogle Scholar
van der Strate, H. J., van de Zande, L., Stam, W. T., and Olsen, J. L. (2002b). The contribution of haploids, diploids and clones to fine-scale population structure in the seaweed Cladophoropsis membranacea (Chlorophyta). Mol. Ecol. 11:329–45.CrossRefGoogle Scholar
van Oppen, M. J. H., Olsen, J. L., Stam, W. T., Van der Hoek, C., and Wiencke, C. (1993). Arctic–Antarctic disjunction in the benthic seaweeds Acrosiphonia arcta (Chlorophyta) and Desmarestia viridis (Phaeophyta) are of recent origin. Mar. Biol. 115:381–6.CrossRefGoogle Scholar
van Oppen, M. J. H., Diekmann, O. E., Wiencke, C., Stam, W. T., and Olsen, J. L. (1994). Tracking dispersal routes: phylogeography of the Arctic–Antarctic disjunct seaweed Acrosiphonia arcta (Chlorophyta). J. Phycol. 30:67–80.CrossRefGoogle Scholar
van Tamelen, P. G. (1996). Algal zonation in tidepools: experimental evaluation of the roles of physical disturbance, herbivory and competition. J. Exp. Mar. Biol. Ecol. 201:197–231.CrossRefGoogle Scholar
van Tussenbroek, B. I., and Barba Santos, M.G. (2011). Demography of Halimeda incrassata (Bryopsidales, Chlorophyta) in a Caribbean reef lagoon. Mar. Biol. 158:1461–71.CrossRefGoogle Scholar
Vanderklift, M. A., and Ponsard, S. (2003). Sources of variation in consumer-diet delta N-15 enrichment: a meta-analysis. Oecologia 136(2):169–82.CrossRefGoogle Scholar
Vanderklift, M. A., and Wernberg, T. (2008). Detached kelps from distant sources are a food subsidy for sea urchins. Oecologia 157(2): 327–35.CrossRefGoogle ScholarPubMed
Varvarigos, V., Katsaros, C., and Galatis, B. (2004). Radial F-actin configurations are involved in polarization during protoplast germination and thallus branching of Macrocystis pyrifera (Phaeophyceae, Laminariales). Phycologia 43:693–702.CrossRefGoogle Scholar
Vasconcelos, M. T., and Leal, M. F. C. (2001). Seasonal variability in the kinetics of Cu, Pb, Cd, and Hg accumulation by macroalgae. Mar. Chem. 74:65–85.CrossRefGoogle Scholar
Vásquez, J. A. (2008). Production, use and fate of Chilean brown seaweeds: resources for a sustainable fishery. J. Appl. Phycol. 20:457–67.CrossRefGoogle Scholar
Vásquez, J. A., Véliz, D., and Pardo, L. M. (2001). Vida bajo las grandes algas pardas. In Alveal, K. and Antezana, T. (eds), Sustentabilidad de la Biodiversidad. Un Problema Actual (pp. 615–34). Concepción, Chile: Base Científico de Concepdión.Google Scholar
Vass, I. (1997). Adverse effects of UV-B light on the structure and function of the photosynthetic apparatus. In Pessarakli, M. (ed.) Handbook of Photosynthesis (pp. 931–49). New York: Marcel Dekker Inc.Google Scholar
Vayda, M. E., and Yuan, M. L. (1994) The heat shock response of an Antarctic alga is evident at 5ºC. Plant Mol. Biol. 24:229–33.CrossRefGoogle Scholar
Venegas, M., Matsuhiro, B., and Edding, M. E. (1993). Alginate composition of Lessonia trabeculata (Phaeophyta: Laminariales) growing in exposed and sheltered habitats. Bot. Mar. 36:47–51.CrossRefGoogle Scholar
Verges, A., Paul, N. A., and Steinberg, P. D. (2008). Sex and life-history stage alter herbivore responses to a chemically defended red alga. Ecology 89:1334–43.CrossRefGoogle ScholarPubMed
Verhaeghe, E. F., Fraysse, A., Guerquin-Kern, J-L., et al. (2008). Microchemical imaging of iodine distribution in the brown alga Laminaria digitata suggests a new mechanism for its accumulation. J. Biol Inorg. Chem. 13:257–69.CrossRefGoogle ScholarPubMed
Vermeij, M. J. A., Dailer, M. L., and Smith, C. M. (2011). Crustose coralline algae can suppress macroalgal growth and recruitment on Hawaiian coral reefs. Mar. Ecol. Prog. Ser. 422:1–7.CrossRefGoogle Scholar
Viano, Y., Bonhomme, D., Camps, M., et al. (2009). Diterpenoids from the Mediterranean brown alga Dictyota sp. evaluated as antifouling substances against a marine bacterial biofilm. J. Nat. Prod. 72:1299–1304.CrossRefGoogle ScholarPubMed
Vidondo, B., and Duarte, C. M. (1998). Population structure, dynamics, and production of the Mediterranean macroalga Codium bursa (Chlorophyceae). J. Phycol. 34:918–24.CrossRefGoogle Scholar
Villares, R., Puente, X., and Carballeira, A. (2001). Ulva and Enteromorpha as indicators of heavy metal pollution. Hydrobiol. 462:221–32.CrossRefGoogle Scholar
Villares, R., Carral, E., Puente, X., and Carballeira, A. (2005). Metal levels in estuarine macrophytes: differences among species. Estuaries 28:948–56.CrossRefGoogle Scholar
Vincensini, L., Blisnick, T., and Bastin, P. (2011). 1001 model organisms to study cilia and flagella. Biol. Cell 103:109–30.CrossRefGoogle ScholarPubMed
Vogel, H., Grieninger, G. E., and Zetsche, K. H. (2002). Differential messenger RNA gradients in the unicellular alga Acetabularia acetabulum. Plant Physiol. 129:1407–16.CrossRefGoogle ScholarPubMed
Vogel, S. (1994). Life in Moving Fluids: The Physical Biology of Flow (2nd edn) Princeton, NJ: Princeton University Press, 467 pp. [8.0, , ]Google Scholar
Vogel, S., and Loudon, C. (1985). Fluid mechanics of the thallus of an intertidal red alga, Halosaccion glandiforme. Biol. Bull. 168:161–74.CrossRefGoogle Scholar
Vreeland, V., and Kloareg, B. (2000). Cell wall biology in red algae: divide and conquer. J. Phycol. 36:793–7.CrossRefGoogle Scholar
Vreeland, V., and Laetsch, W. M. (1989). Identification of associating carbohydrate sequences with labeled oligosaccharides. Localization of alginate-gelling subunits in walls of a brown alga. Planta 177:423–34.CrossRefGoogle Scholar
Vreeland, V., Zablackis, E., and Laetsch, W. M. (1992). Monoclonal-antibodies as molecular markers for the intracellular and cell wall distribution of carrageenan epitopes in Kappaphycus (Rhodophyta) during tissue-development. J. Phycol. 28:328–42.CrossRefGoogle Scholar
Vreeland, V., Waite, J. H., and Epstein, L. (1998). Polyphenols and oxidases in substratum adhesion by marine algae and mussels. J. Phycol. 34:1–8.CrossRefGoogle Scholar
Vreeland, Y., Zablackis, E., Doboszewski, B., and Laetsch, W. M. (1987). Molecular markers for marine algal polysaccharides. Hydrobiologia 151/152:155–60.CrossRefGoogle Scholar
Waaland, J. R., Dickson, L. G., and Carrier, J. E. (1987). Conchocelis growth and photoperiodic control of conchospore release in Porphyra torta (Rhodophyta). J. Phycol. 23:399–406.CrossRefGoogle Scholar
Waaland, J. R., Stiller, J. W., and Cheney, D. P. (2004). Macroalgal candidates for genomics. J. Phycol. 40:26–33.CrossRefGoogle Scholar
Waaland, S. D. (1980). Development in red algae: elongation and cell fusion. In Gantt, E. (ed.), Handbook of Phycological Methods. Developmental and Cytological Methods (pp. 85–93). Cambridge: Cambridge University Press.Google Scholar
Waaland, S. D. (1989). Cellular morphogenesis in the filaments of the red alga Griffithsia. In Coleman, A. W., Goff, L. J., and Stein-Taylor, J. R. (eds), Algae as Experimental Systems (pp. 121–34). New York: Alan R. Liss.Google Scholar
Waaland, S. D. (1990). Development. In Cole, K. M. and Sheath, R. G. (eds), Biology of the Red Algae (pp. 259–73). Cambridge: Cambridge University Press.Google Scholar
Waaland, S. D., and Cleland, R. (1972). Development of the red alga Griffithsia pacifica: control by internal and external factors. Planta 105:196–204.CrossRefGoogle ScholarPubMed
Waaland, S. D., and Cleland, R. (1974). Cell repair through cell fusion in the red algaGriffithsia pacifica. Protoplasma 79:185–96.CrossRefGoogle ScholarPubMed
Waaland, S. D., and Waaland, J. R. (1975). Analysis of cell elongation in red algae by fluorescent labeling. Planta 126:127–38.CrossRefGoogle Scholar
Wada, S., Aoki, M. N., Tsuchiya, Y., et al. (2007) Quantitative and qualitative analyses of dissolved organic matter released from Ecklonia cava Kjellman, in Oura Bay, Shimoda, Izu Peninsula, Japan. J. Exp. Mar. Biol. Ecol. 349:344–58.CrossRefGoogle Scholar
Wada, S., Aoki, M. N., Mikami, A., T et al. (2008) Bioavailability of macroalgal dissolved organic matter in seawater. Mar. Ecol. Prog. Ser. 370:33–44.CrossRefGoogle Scholar
Wagner, F., and Falkner, G. (2001). Phosphate limitation. In Rai, L.C. and Gaur, J.P. (eds). Algal Adaptation to Environmental Stresses (pp. 65–110). New York: Springer.CrossRefGoogle Scholar
Wahl, M. (2008). Ecological lever and interface ecology: epibiosis modulates the interactions between host and environment. Biofouling 24:427–38.CrossRefGoogle Scholar
Wai, T.-C., and Williams, G. A. (2005). The relative importance of herbivore-induced effects on productivity of crustose coralline algae: sea urchin grazing and nitrogen excretion. J. Exp. Mar. Biol. Ecol. 324:141–56.CrossRefGoogle Scholar
Walker, C. H., Hopkin, S. P., Sibly, R. M., and Peakall, D. B. (2006). Principles of Ecotoxicology (3rd edn). New York: Taylor and Francis, 315 pp.Google Scholar
Walker, G., Dorrell, R. G., Schlacht, A., and Dacks, J. B. (2011). Eukaryotic systematics: a user’s guide for cell biologists and parasitologists. Parasitology 138:1638–63.CrossRefGoogle ScholarPubMed
Wallentinus, I. (1984). Comparisons of nutrient uptake rates for Baltic macroalgae with different thallus morphologies. Mar. Biol. 80:215–25.CrossRefGoogle Scholar
Walsh, R. S., and Hunter, K. A. (1992). Influence of phosphorus storage on the uptake of cadmium by the marine alga Macrocystis pyrifera. Linmol. Oceanogr. 37:1361–9.CrossRefGoogle Scholar
Walters, L. J., and Smith, C. M. (1994). Rapid rhizoid production in Halimeda discoidea decaisne (Chlorophyta, Caulerpales) fragments: a mechanism for survival after separation from adult thalli. J. Exp. Mar. Biol. Ecol. 175:105–20.CrossRefGoogle Scholar
Walters, L. J., Smith, C. M., Coyer, J. A., et al. (2002). Asexual propagation in the coral reef macroalga Halimeda (Chlorophyta, Bryopsidales): production, dispersal and attachment of small fragments. J. Exp. Mar. Biol. Ecol. 278:47–65.CrossRefGoogle Scholar
Warwick, R. M., Clarke, K. R., and Suharsono, (1990). A statistic-analysis of coral community response to the El Niño in the Thousand Islands, Indonesia. Coral Reefs 8:171–9.CrossRefGoogle Scholar
Watanabe, S., Metaxas, A., and Scheibling, R. E. (2009). Dispersal potential of the invasive green alga Codium fragile ssp. fragile. J. Exp. Mar. Biol. Ecol. 381:114–25.CrossRefGoogle Scholar
Watson, B. A., and Waaland, S. D. (1983). Partial purification and characterization of a glycoprotein cell fusion hormone from Griffithsia pacifica, a red alga. Plant Physiol. 71:327–32.CrossRefGoogle ScholarPubMed
Watson, B. A., and Waaland, S. D. (1986). Further biochemical characterization of a cell fusion hormone from the red algaGriffithsia pacifica. Plant Cell Physiol. 27:1043–50.Google Scholar
Webb, W. L., Newton, M., and Starr, D. (1974). Carbon dioxide exchange ofAlnus rubra. Oecologia 17:281–91.CrossRefGoogle Scholar
Weber, A. P. M., and Osteryoung, K. W. (2010). From endosymbiosis to synthetic photosynthetic life. Plant Physiol. 154:593–7.CrossRefGoogle ScholarPubMed
Wehr, J. D., and Sheath, R. G. (eds) (2003). Freshwater Algae of North America: Ecology and Classification. New York: Academic Press. 918 pp.Google Scholar
Weich, R. G., and Graneli, E. (1989). Extracellular alkaline phosphatase activity in Ulva lactuca L. J. Exp. Mar. Biol. Ecol. 129:33–44.CrossRefGoogle Scholar
Weinberger, F., and Friedlander, M. (2000). Response of Gracilaria conferta (Rhodophyta) to oligoagars results in defense against agar-degrading epiphytes. J. Phycol. 36:1079–86.CrossRefGoogle Scholar
Weinberger, F., Coquempot, B., Forner, S., et al. (2007). Different regulation of haloperoxidation during agar oligosaccharide-activated defence mechanisms in two related red algae, Gracilaria sp. and Gracilaria chilensis. J. Exper. Bot. 58:4365–72.CrossRefGoogle ScholarPubMed
Weinberger, F., Guillemin, M.-L., Destombe, C., et al. (2010). Defense evolution in the Gracilariaceae (Rhodophyta): substrate-regulated oxidation of agar oligosaccharides is more ancient than the oligoagar-activated oxidative burst. J. Phycol. 46:958–68.CrossRefGoogle Scholar
Weiner, S. (1986). Organization of extracellularly mineralized tissues: a comparative study of biological crystal growth. Crit. Rev. Biochem. 20:365–408.CrossRefGoogle ScholarPubMed
Weissflog, J., Adolph, S., Wiesemeier, T., and Pohnert, G. (2008). Reduction of herbivory through wound-activated protein cross-linking by the invasive macroalga Caulerpa taxifolia. ChemBioChem. 9:29–32.CrossRefGoogle ScholarPubMed
Welling, M., Pohnert, G., Küpper, F. C., and Ross, C. (2009). Rapid biopolymerisation during wound plug formation in green algae. J. Adhesion 85:825–38.CrossRefGoogle Scholar
Wernberg, T. (2005). Holdfast aggregation in relation to morphology, age, attachment and drag for the kelp Ecklonia radiata. Aquat. Bot. 82:168–80.CrossRefGoogle Scholar
Wernberg, T., and Vanderklift, M. A. (2010). Contribution of temporal and spatial components to morphological variation in the kelp Ecklonia (Laminariales). J. Phycol. 46:153–61.CrossRefGoogle Scholar
Wernberg, T., Kendrick, G. A., and Phillips, J. C. (2003). Regional differences in kelp-associated algal assemblages on temperate limestone reefs in south-western Australia. Diversity and Distributions 9:427–41.CrossRefGoogle Scholar
Wernberg, T., Vanderklift, M. A., How, J., and Lavery, P. S. (2006). Export of detached macroalgae from reefs to adjacent seagrass beds. Oecologia 147(4):692–701.CrossRefGoogle ScholarPubMed
Wernberg, T., Thomsen, M. S., Tuya, F., G. A., et al. (2010). Decreasing resilience of kelp beds along a latitudinal temperature gradient: potential implications for a warmer future. Ecol. Lett. 13:685–94.CrossRefGoogle ScholarPubMed
Wernberg, T., Russell, B. D., Thomsen, M. S., et al. (2011a). Seaweed communities in retreat from ocean warming. Curr. Biol. 21:1828–32.CrossRefGoogle ScholarPubMed
Wernberg, T., Thomsen, M. S., Tuya, F., and Kendrick, G. A. (2011b). Biogenic habitat structure of seaweeds change along a latitudinal gradient in ocean temperature. J. Exp. Mar. Biol. Ecol. 400:264–71.CrossRefGoogle Scholar
West, J. A. (1972). Environmental regulation of reproduction in Rhodochorton purpureum. In Abbott, I. A. and Kurogi, M. (eds), Contributions to the Systematics of the Benthic Marine Algae of the North Pacific (pp. 213–30). Kobe: Japan Soc. Phycol.Google Scholar
Westermeier, R., and Gómez, I. (1996). Biomass, energy contents and major organic compounds in the brown alga Lessonia nigrescens (Laminariales, Phaeophyceae) from Mehuin, South Chile. Bot. Mar. 39:553–9.CrossRefGoogle Scholar
Weykam, G., Gómez, I., Wiencke, C., Iken, K., and Klöser, H. (1996). Photosynthetic characteristics and C:N ratios of macroalgae from King George Island (Antarctica). J. Exp. Mar. Biol. Ecol. 204:1–22.CrossRefGoogle Scholar
Weykam, G., Thomas, D. N., and Wiencke, C. (1997). Growth and photosynthesis of the Antarctic red algae Palmaria decipiens (Palmariales) and Iridaea cordata (Gigartinales) during and following extended periods of darkness. Phycologia 36:395–405.CrossRefGoogle Scholar
Wheeler, P. A. (1979). Uptake of methylamine (an ammonium analogue) by Macrocystis pyrifera (Phaeophyta). J. Phycol. 15:12–17.CrossRefGoogle Scholar
Wheeler, P. A. (1985). Nutrients. In Littler, M. M. and Littler, D. S. (eds), Handbook of Phycological Methods: Ecological Field Methods: Macroalgae (pp. 493–508). Cambridge: Cambridge University Press.Google Scholar
Wheeler, P. A., and Björnsater, B. R. (1992). Seasonal fluctuations in tissue nitrogen, phosphorus, and N: P for five macroalgal species common to the Pacific northwest coast. J. Phycol. 28:1–6.CrossRefGoogle Scholar
Wheeler, P. A., and North, W. J. (1980). Effect of nitrogen supply on nitrogen content and growth rate of juvenile Macrocystis pyrifera (Phaeophyta) sporophytes. J. Phycol. 16:577–82.CrossRefGoogle Scholar
Wheeler, P. A., and North, W. J. (1981). Nitrogen supply, tissue composition and frond growth rates for Macrocystis pyrifera off the coast of southern California. Mar. Biol. 64:59–69.CrossRefGoogle Scholar
Wheeler, W. N. (1980). Effect of boundary layer transport on the fixation of carbon by the giant kelp Macrocystis pyrifera. Mar. Biol. 56:103–10. [, , ]CrossRefGoogle Scholar
Wheeler, W. N. (1982). Nitrogen nutrition of Macrocystis. In Srivastiva, L. M. (ed.), Synthetic and Degradative Processes in Marine Macrophytes (pp. 121–37). Berlin:Walter de Gruyter.Google Scholar
Wheeler, W. N. (1988). Algal productivity and hydrodynamics-a synthesis. Prog. Phycol. Res. 6:23–58.Google Scholar
Wheeler, W. N., and Srivastava, L. M. (1984). Seasonal nitrate physiology of Macrocystis integrifolia Bory. J. Exp. Mar. Biol. Ecol. 76:35–50.CrossRefGoogle Scholar
Wheeler, W. N., and Weidner, M. (1983). Effects of external inorganic nitrogen concentration on metabolism, growth and activities of key carbon and nitrogen assimilating enzymes of Laminaria saccharina (Phaeophyceae) in culture. J. Phycol. 19:92–6.CrossRefGoogle Scholar
White, H. H. (1984). Concepts in Marine Pollution Measurements. College Park, MD: Maryland Sea Grant College Program, University of Maryland, 743 pp.Google Scholar
White, P. J., and Broadley, M. R. (2003). Calcium in plants. Annals Bot. 92:487–511.CrossRefGoogle ScholarPubMed
Whitton, B. A., and Potts, M. (1982). Marine littoral. In Carr, N. G. and Whitton, B. A. (eds), The Biology of Cyanobacteria (pp. 515–42). Oxford: Blackwell Scientific.Google Scholar
Wichard, T., and Oertel, W. (2010). Gametogenesis and gamete release of Ulva mutabilis and Ulva lactuca (Chlorophyta): regulatory effects and chemical characterization of the “swarming inhibitor”. J. Phycol. 46:248–59.CrossRefGoogle Scholar
Wiencke, C. (1990a). Seasonality of brown macroalgae from Antarctica – a long-term culture study under fluctuating Antarctic daylengths. Polar Biol. 10:589–600.CrossRefGoogle Scholar
Wiencke, C. (1990b). Seasonality of red and green macroalgae from Antarctica – a long-term culture study under fluctuating Antarctic daylengths. Polar Biol. 10:60–7.Google Scholar
Wiencke, C. (1996). Recent advances in the investigation of Antarctic macroalgae. Polar Biol. 16:231–40.CrossRefGoogle Scholar
Wiencke, C., and Clayton, M. N. (1990). Sexual reproduction, life history, and early development in culture of the Antarctic brown alga Himantothallus grandifolius (Desmarestiales, Phaeophyceae). Phycologia 29:9–18.CrossRefGoogle Scholar
Wiencke, C., and Clayton, M. (2002). Antarctic Seaweeds. Synopses of the Antarctic Benthos. Lichtenstein: A.R.G. Gantner Verlag, 239 pp.Google Scholar
Wiencke, C., and Clayton, M. N. (2009). Biology of polar benthic algae. Bot. Mar. 52:479–81.Google Scholar
Wiencke, C., and Davenport, J. (1987). Respiration and photosynthesis in the intertidal alga Cladophora rupestris (L.) Kütz. under fluctuating salinity regimes. J. Exp. Mar. Biol. Ecol. 114:183–97.CrossRefGoogle Scholar
Wiencke, C., and Läuchli, A. (1981). Inorganic ions and floridoside as osmostic solutes inPorphyra umbilicalis. Z. Pflanzenphysiol. 103:247–58.CrossRefGoogle Scholar
Wiencke, C., and tom Dieck, I. (1989). Temperature requirements for growth and temperature tolerance of macroalgae endemic to the Antarctic region. Mar. Ecol. Prog. Ser. 54:189–97. [, , ]CrossRefGoogle Scholar
Wiencke, C., Rahmel, J., Karsten, U., Weykam, G., and Kirst, G. (1993). Photosynthesis of marine macroalgae from Antarctica: light and temperature requirements. Bot. Acta 106:78–87.CrossRefGoogle Scholar
Wiencke, C., Bartsch, I., Bischoff, B., Peters, A. F., and Breeman, A. M. (1994). Temperature requirements and biogeography of Antarctic, Arctic and amphiequatorial seaweeds. Bot. Mar. 37:247–59.CrossRefGoogle Scholar
Wiencke, C., Gómez, I., Pakker, H., et al. (2000). Impact of UV-radiation on viability, photosynthetic characteristics and DNA on brown algal zoospores: Implications for depth zonation. Mar. Ecol. Progr. Ser. 197:217–29.CrossRefGoogle Scholar
Wiencke, C., Roleda, M., Gruber, A., Clayton, M., and Bischof, K. (2006). Susceptibility of zoospores to UV radiation determines upper depth distribution limit of Arctic kelps: evidence through field experiments. J Ecol 94:455–63. [, , ]CrossRefGoogle Scholar
Wiencke, C., Gómez, I., and Dunton, K. (2009). Phenology and seasonal physiological performance of polar seaweeds. Bot. Mar. 52:585–92.CrossRefGoogle Scholar
Wiens, J. A. (1989). Spatial scaling in ecology. Func. Ecol. 3:385–97.CrossRefGoogle Scholar
Wiesemeier, T., Hay, M. E., and Pohnert, G. (2007). The potential role of wound-activated volatile release in the chemical defence of the brown alga Dictyota dichotoma: blend recognition by marine herbivores. Aquat. Sci. 69:403–12.CrossRefGoogle Scholar
Wiesemeier, T., Jahn, K., and Pohnert, G. (2008). No evidence for the induction of brown algal chemical defense by the phytohormones jasmonic acid and methyl jasmonate. J. Chem. Ecol. 34:1523–31.CrossRefGoogle ScholarPubMed
Wilce, R. T. (1990). Role of the Arctic Ocean as a bridge between the Atlantic and the Pacific Ocean: fact and hypothesis. In Garbary, D. J. and South, G. R. (eds), Evolutionary Biogeography of the Marine Algae of the North Atlantic. NATO ASI. Ser. G. Ecol. Sci. (vol. 22, pp. 323–47). Berlin: Springer-Verlag.CrossRefGoogle Scholar
Wilce, R. T., and Davis, A. N. (1984). Development of Dumontia contorta (Dumontiaceae, Cryptonemiales) compared with that of other higher red algae. J. Phycol. 20:336–51.CrossRefGoogle Scholar
Wilkinson, C. (2008). Status of Coral Reefs of the World: 2008. Townsville, Australia: Global Coral Reef Monitoring Network and Reef and Rainforest Research Centre.Google Scholar
Williams, I. D., Polunin, V. C., and Hendrick, V. J. (2001). Limits to grazing by herbivorous fishes and the impact of low coral cover on macroalgal abundance on a coral reef in Belize. Mar. Ecol. Prog. Ser. 222:187–96.CrossRefGoogle Scholar
Williams, S. L. (1984). Uptake of sediment ammonium and translocation in a marine green macroalga Caulerpa cupressoides. Limnol. Oceanogr. 29:374–9.CrossRefGoogle Scholar
Williams, S. L. (2007). Introduced species in seagrass ecosystems: status and concerns. J. Exp. Mar. Biol. Ecol. 350:89–110.CrossRefGoogle Scholar
Williams, S. L., and Carpenter, R. C. (1990). Photosynthesis/ photon flux density relationships among components of coral reef algal turfs. J. Phycol. 26:36–40. [, , , ]CrossRefGoogle Scholar
Williams, S. L., and Dethier, M. N. (2005). High and dry: variation in net photosynthesis of the intertidal seaweed Fucus gardneri. Ecology 86:2373–9.CrossRefGoogle Scholar
Williams, S. L., and Fisher, T. R. (1985). Kinetics of nitrogen-15 labeled ammonium uptake by Caulerpa cupressoides (Chlorophyta). J. Phycol. 21:287–96.CrossRefGoogle Scholar
Williams, S. L., and Heck, K. L. (2001). Seagrass community ecology. In Bertness, M. D., Gaines, S. D., and Hay, M. E. (eds), Marine Community Ecology (pp. 317–37). Sunderland, MA: Sinauer Associates, Inc.Google Scholar
Williams, S. L., and Herbert, S. K. (1989). Transient photosynthetic responses of nitrogen-deprived Petalonia fascia and Laminaria saccharina (Phaeophyta) to ammonium resupply. J. Phycol. 25:515–22.CrossRefGoogle Scholar
Williams, S. L., and Schroeder, S. L. (2004). Eradication of the invasive seaweed Caulerpa taxifolia by chlorine bleach. Mar. Ecol. Prog. Ser. 272:69–76.CrossRefGoogle Scholar
Williams, S. L., and Smith, J. E. (2007). A global review of the distribution, taxonomy, and impacts of introduced seaweeds. Annu. Rev. Ecol. Evol. Syst. 38:327–59.CrossRefGoogle Scholar
Williams, S. L., Breda, V. A., Anderson, T. W., and Nyden, B. B (1985). Growth and sediment disturbances of Caulerpa spp. (Chlorophyta) in a submarine canyon. Mar. Ecol. Prog. Ser. 21:275–81.CrossRefGoogle Scholar
Wilmotte, A., Goffart, A., and Demoulin, V. (1988). Studies of marine epiphytic algae, Calvi, Corsica. I. Determination of minimal sampling areas for microscopic algal epiphytes. Br. Phycol. J. 23:251–8.CrossRefGoogle Scholar
Wilson, J. B. (1999). Assembly rules in plant communities. In Weiher, E. and Keddy, P. (eds), Ecological Assembly Rules: Perspectives, Advances and Retreats (pp. 130–64). Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Wilson, S. M., West, J. A., and Pickett-Heaps, J. D. (2003). Time-lapse videomicroscopy of fertilization and the actin cytoskeleton in Murrayella periclados (Rhodomelaceae, Rhodophyta). Phycologia 42:638–45.CrossRefGoogle Scholar
Wiltens, J., Schreiber, U., and Vidaver, W. (1978). Chlorophyll fluorescence induction: an indicator of photosynthetic activity in marine algae undergoing desiccation. Can. J. Bot. 56: 2787–94.CrossRefGoogle Scholar
Wing, S. R., and Patterson, M. R. (1993). Effects of wave-induced light flecks in the intertidal zone on photosynthesis in the macroalgae Postelsia palmaeformis and Hedophyllum sessile (Phaeophyceae). Mar. Biol. 116:519–25. [, , , , ]CrossRefGoogle Scholar
Wing, S. R., Leichter, J.J., Perrin, C. et al. (2007). Topographic shading and wave exposure influence morphology and ecophysiology of Ecklonia radiata (C. Aghard 1817) in Fiordland, New Zealand. Limnol. Oceanogr. 52:1853–64.CrossRefGoogle Scholar
Wise, R. R. (2007). The diversity of plastid form and function. In Wise, R. R. and Hoober, J. K. (eds), The Structure and Function of Plastids (pp. 3–26). Dordrecht, The Netherlands: Springer.CrossRefGoogle Scholar
Witman, J. D., and Dayton, P. K. (2001). Rocky subtidal communities. In Bertness, M. D., Gaines, S. D., and Hay, M. E. (eds), Marine Community Ecology (pp. 339–66). Sunderland, MA: Sinauer Assocs. [, , ]Google Scholar
Witman, J. D., and Roy, K. (2009). Experimental marine macroecology: progress and prospects. In Witman, J. D., and Roy, K. (eds), Marine Macroecology (pp. 341–56). Chicago: University of Chicago Press.CrossRefGoogle Scholar
Witman, J. D., Brandt, M., and Smith, F. (2010). Coupling between subtidal prey and consumers along a mesoscale upwelling gradient in the Galapagos Islands. Ecol. Monogr. 80:153–77.CrossRefGoogle Scholar
Wolanski, E., and Hammer, W. M. (1988). Topographically controlled fronts in the ocean and their biological influence. Science 241:177–81.CrossRefGoogle ScholarPubMed
Wolcott, B. D. (2007). Mechanical size limitation and life-history strategy of an intertidal seaweed. Mar. Ecol. Prog. Ser. 338:1–10.CrossRefGoogle Scholar
Wolfe, J. M., and Harlin, M. M. (l988a). Tidepools in southern Rhode Island, USA. I. Distribution and seasonality of macroalgae. Bot. Mar. 31:525–36.CrossRefGoogle Scholar
Wolfe, J. M., and Harlin, M. M. (l988b). Tidepools in southern Rhode Island, USA. II. Species diversity and similarity analysis of macroalgal communities. Bot. Mar. 31:537–46.CrossRefGoogle Scholar
Wolff, M. (1987). Population dynamics of the Peruvian scallop Argopecten purpuratus during the El Nino Phenomenon 1983. Can. J. Fish. Aquat. Sci. 44:1684–91.CrossRefGoogle Scholar
Womersley, H. B. S. (1971). Palmoclathrus, a new deep-water genus of Chlorophyta. Phycologia 10:229–33.CrossRefGoogle Scholar
Wong, K. F., and Craigie, J. S. (1978). Sulfohydrolase activity and carrageenan biosynthesis in Chondrus crispus (Rhodophyceae). Plant Physiol. 61:663–6.CrossRefGoogle Scholar
Wong, P.-F., Tan, L.-J., Nawi, H., and AbuBakar, S. (2006). Proteomics of the red alga, Gracilaria changii (Gracilariales, Rhodophyta). J. Phycol. 42:113–20.CrossRefGoogle Scholar
Wong, T. K.-M., Ho, C.-L., Lee, W.-W., Rahim, R. A., and Phang, S.-M. (2007). Analyses of expressed sequence tags from Sargassum binderi (Phaeophyta). J. Phycol. 43:528–34.CrossRefGoogle Scholar
Wood, S. A., Lilley, S. A., Schiel, D. R., and Shurin, J. B. (2010). Organismal traits are more important than environment for species interactions in the intertidal zone. Ecol. Lett. 13:1160–71.CrossRefGoogle ScholarPubMed
Woodley, J. D., Chornesky, E. A., Clifford, P. A., et al. (1981). Hurricane Allen’s impact on Jamaican coral reefs. Science 214:749–55.CrossRefGoogle ScholarPubMed
Worm, B., and Lotze, H. K. (2006). Effects of eutrophication, grazing and algal blooms on rocky shores. Limnol. Oceanogr. 51:569–79.CrossRefGoogle Scholar
Worm, B., Lotze, H. K., and Sommer, U. (2001). Algal propagule banks modify competition, consumer and resource control on Baltic rocky shores. Oecologia 128:281–93.CrossRefGoogle ScholarPubMed
Worm, B., Barbier, E. B., Beaumont, N., et al. (2006). Impacts of biodiversity loss on ocean ecosystem services. Science 314:787–90.CrossRefGoogle ScholarPubMed
Wright, D. G., Pawlowicz, R., McDougall, T. J., Feistel, R., and Marion, G. M. (2010). Absolute salinity, “density salinity” and the reference-composition salinity scale: present and future use in the seawater standard TEOS-10. Ocean Sci. Discuss. 7:1559–625.CrossRefGoogle Scholar
Wright, J. T., Dworjanyn, S. A., Steinberg, C. N., Williamson, J. E., and Poore, A. G. B. (2005). Density-dependent sea urchin grazing: differential removal of species, changes in community composition and alternative community states. Mar. Ecol. Prog. Ser. 298:143–56.CrossRefGoogle Scholar
Wright, P. J., Chudek, J. A., Foster, R., Davison, I. R., and Reed, R. H. (1985). The occurrence of altritol in the brown algaHimanthalia elongata. Br. Phycol. J. 20:191–2.Google Scholar
Wright, P. J., Green, J. R., and Callow, J. A. (1995a). The Fucus (Phaeophyceae) sperm receptor for eggs. 1. Development and characteristics of a binding assay. J. Phycol. 31:584–91.CrossRefGoogle Scholar
Wright, P. J., Callow, J. A., and Green, J. R. (1995b). The Fucus (Phaeophyceae) sperm receptor for eggs. 2. Isolation of a binding-protein which partially activates eggs. J. Phycol. 31:592–600.CrossRefGoogle Scholar
Wulff, A., Iken, K., Quartino, M. L., et al. (2009). Biodiversity, biogeography and zonation of marine benthic micro- and macroalgae in the Arctic and Antarctic. Bot. Mar. 52:491–507.CrossRefGoogle Scholar
Xu, B., Zhang, Q. S., Qu, S. C., Cong, Y. Z., and Tang, X. X. (2009). Introduction of a seedling production method using vegetative gametophytes to the commercial farming of Laminaria in China. J. Appl. Phycol. 21:171–8.CrossRefGoogle Scholar
Xu, H., Deckert, R. J., and Garbary, D. J. (2008). Ascophyllum and its symbionts. X. Ultrastructure of the interaction between A. nodosum (Phaeophyceae) and Mycophycias ascophylli (Ascomycetes). Botany 86:185–93.CrossRefGoogle Scholar
Yamada, T., Ikawa, K., and Nisizawa, K. (1979). Circadian rhythm of the enzymes participating in the CO2 photoassimilation of a brown alga, Spatoglossum pacificum. Bot. Mar. 22:203–9.Google Scholar
Yan, X.-H., and Huang, M. (2010). Identification of Porphyra haitanensis (Bangiales, Rhodophyta) meiosis by simple sequence repeat markers. J. Phycol. 46:982–6.CrossRefGoogle Scholar
Yarish, C., and Pereira, R. (2008). Mass production of marine macroalgae. In Jørgensen, S. E. and Fath, B. D. (eds), Ecological Engineering. Vol. 3 of Encyclopedia of Ecology, (pp. 2236–47). Oxford: Elsevier. [, , , , ]Google Scholar
Yarish, C., Edwards, P., and Casey, S. (1979). Acclimation responses to salinity of three estuarine red algae from New Jersey. Mar. Biol. 51:289–94.CrossRefGoogle Scholar
Yarish, C., Edwards, P., and Casey, S. (1980). The effects of salinity, and calcium and potassium variation on the growth of two estuarine red algae. J. Exp. Mar. Biol. Ecol. 47:235–49.CrossRefGoogle Scholar
Yarish, C., Breeman, A. M., and van den Hoek, C. (1984). Temperature, light, and photoperiod responses of some northeast American and west European endemic rhodophytes in relation to their geographical distribution. Helgol. Meeresunters 38:273–304.CrossRefGoogle Scholar
Yarish, C., Kirkman, H., and Lüning, K. (1987). Lethal exposure times and preconditioning to upper temperature limits of some temperate North Atlantic red algae. Helgol. Meeresunters 41: 323–7.CrossRefGoogle Scholar
Yarish, C., Brinkhuis, B. H., Egan, B., and Garcia-Ezquivel, Z. (1990). Morphological and physiological bases for Laminaria selection protocols in Long Island Sound. In Yarish, C., Penniman, C. A., and Van Patten, P. (eds), Economically Important Plants of the Atlantic: Their Biology and Cultivation (pp. 53–94). Groton, CT: Connecticut Sea Grant College Program.Google Scholar
Yellowlees, D., Rees, T. A. V., and Leggat, W. (2008). Metabolic interactions between algal symbionts and invertebrate hosts. Plant Cell Environ. 31:679–94.CrossRefGoogle ScholarPubMed
Yiou, F., Raisbeck, G. M., Christensen, G. C., and Holm, E. (2002). 129I/127I, 129I/137Cs and 129I/99Tc in the Norwegian coastal current from 1980 to 1998. J. Environ. Radioact. 60:61–1.CrossRefGoogle ScholarPubMed
Yokohama, Y., and Misonou, T. (1980). Chlorophyll a:b ratios in marine benthic algae. Jap. J. Phycol. 28: 219–23.Google Scholar
Yoon, H. S., Hackett, J. D., Ciniglia, C., et al. (2004). A molecular timeline for the origin of photosynthetic eukaryotes. Mol. Biol. Evol. 21:809–18.CrossRefGoogle Scholar
Yoon, H. S., Zuccarello, G. C., and Battacharya, D. (2010). Evolutionary history and taxonomy of red algae. In Seckbach, J., and Chapman, D. J. (eds) Red Algae in the Genomic Age (pp. 25–44). Dordrecht, The Netherlands: Springer.CrossRefGoogle Scholar
Yoon, K. S., Lee, K. P., Klochkova, T. A., and Kim, G. H. (2008). Molecular characterization of the lectin, bryohealin, involved in protoplast regeneration of the marine alga Bryopsis plumosa (Chlorophyta). J. Phycol. 44:103–12.CrossRefGoogle Scholar
Young, E. B., Lavery, P. S., van Elven, B., Dring, M. J., and Berges, J. A. (2005). Nitrate reductase activity in macroalgae and its vertical distribution in macroalgal epiphytes of seagrasses. Mar. Ecol. Prog. Ser. 288:103–14.CrossRefGoogle Scholar
Young, E. B., Dring, M. J., Savidge, G., Birkett, D. A., and Berges, J. A. (2007). Seasonal variations in nitrate reductase activity and internal N pools in intertidal brown algae are correlated with ambient nitrate concentrations. Plant Cell Envir. 30:764–74.CrossRefGoogle ScholarPubMed
Young, E. B., Berges, J. A., and Dring, M. J. (2009). Physiological responses of intertidal marine brown algae to nitrogen deprivation and resupply of nitrate and ammonium. Physiol. Plant. 135:400–11.CrossRefGoogle ScholarPubMed
Young-Sook, O., Sim, D. S., and Kim, S. J. (2001). Effects of nutrients on crude oil biodegradation in the upper intertidal zone. Mar. Poll. Bull. 42:1367–72.Google Scholar
Ytreberg, E., Karlsson, J., Hoppe, S., Eklund, B., and Ndungu, K. (2011a). Effects of organic complexation on copper accumulation and toxicity to the estuarine red macroalga Ceramium tenuicorne: A test of the free ion activity model. Environ. Sci. Technol. 45:3145–53.CrossRefGoogle ScholarPubMed
Ytreberg, E., Karlsson, J., Ndungu, K., et al. (2011b). Influence of salinity and organic matter on the toxicity of Cu to a brackish water and marine clone of the red macroalgaCeramium tenuicorne. Ecotox. Environ. Safety 74:636–42.CrossRefGoogle ScholarPubMed
Zacher, K., Wulff, A., Molis, M., Hanelt, D., and Wiencke, C. (2007). Ultraviolet radiation and consumer effects on a field-grown intertidal macroalgal assemblage in Antarctica. Global Change Biol. 13:1201–15.CrossRefGoogle Scholar
Zacher, K., Rautenberger, R., Hanelt, D., Wulff, A., and Wiencke, C. (2009). The abiotic environment of polar marine benthic algae. Bot. Mar. 52:483–90.CrossRefGoogle Scholar
Zbikowski, R., Szefer, P., and Latala, A. (2007). Comparison of green algae Cladophora sp. and Enteromorpha sp. as potential biomonitors of chemical elements in the southern Baltic. Sci. Total Environ. 387:320–32.CrossRefGoogle ScholarPubMed
Zechman, F. W., Verbruggen, H., Leliaert, F., et al. (2010). An unrecognized ancient lineage of green plants persists in deep marine waters. J. Phycol. 46:1288–95.CrossRefGoogle Scholar
Zhang, Q. S., Tang, X. X., Cong, Y. Z., Qu, S. C., and Yang, G. P. (2007). Breeding of an elite Laminaria variety 90–1 through inter-specific gametophyte crossing. J. Appl. Phycol. 19:303–11.CrossRefGoogle Scholar
Zhang, A. Q., Leung, K. M. Y., Kwok, K. W. H., Bao, V. W. W., and Lam, M. H. W. (2008a). Toxicities of antifouling biocide Irgarol 1051 and its major degraded product to marine primary producers. Mar. Poll. Bull. 57:575–86.CrossRefGoogle ScholarPubMed
Zhang, Q. S., Qu, S. C., Cong, Y. Z., Luo, S. J, and Tang, X. X. (2008b). High throughput culture and gametogenesis induction of Laminaria japonica gametophyte clones. J. Appl. Phycol. 20:205–11.CrossRefGoogle Scholar
Zhang, X., and van der Meer, J. P. (1988). Polyploid gametophytes of Gracilaria tikvahiae (Gigartinales, Rhodophyta). Phycologia 27:312–18.CrossRefGoogle Scholar
Zhuang, S. H. (2006). Species richness, biomass and diversity of macroalgal assemblages in tidepools of different sizes. Mar. Ecol. Prog. Ser. 309:67–73.CrossRefGoogle Scholar
Zimmerman, R. C., and Robertson, D. L. (1985). Effects of El Niño on local hydrography and growth of the giant kelp, Macrocystis pyrifera, at Santa Catalina Island, California. Limnol. Oceanogr. 30: 1298–1302.CrossRefGoogle Scholar
Zuccarello, G. C., Moon, D., and Goff, L. J. (2004). A phylogenetic study of parasitic genera placed in the family Choreocolacaceae (Rhodophyta). J. Phycol. 40:937–45.CrossRefGoogle Scholar
Zuccaro, A., and Mitchell, J. I. (2005). Fungal communities of seaweeds. In Dighton, J., White, J. F., and Oudemans, P. (eds), The Fungal Community: Its Organization and Role in the Ecosystem (3rd edn) (pp. 533–79). Boca Raton, FL: CRC Press.CrossRefGoogle Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

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

Available formats
×

Save book to Google Drive

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

Available formats
×