Skip to main content Accessibility help
×
Hostname: page-component-586b7cd67f-rcrh6 Total loading time: 0 Render date: 2024-11-29T00:08:36.665Z Has data issue: false hasContentIssue false

Part III - The Future

Published online by Cambridge University Press:  24 October 2024

Mario Giordano
Affiliation:
Università degli Studi di Ancona, Italy
John Beardall
Affiliation:
Monash University, Victoria
John A. Raven
Affiliation:
University of Dundee
Stephen C. Maberly
Affiliation:
UK Centre for Ecology & Hydrology, Lancaster
Get access
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2024

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

References

Arafeh-Dalmau, N., Montaño-Moctezuma, G., Martínez, J. et al. (2019). Extreme marine heatwaves alter kelp forest community near its equatorward distribution limit. Frontiers in Marine Science 6: 499. https://doi.org/10.3389/fmars.2019.00499.CrossRefGoogle Scholar
Bach, L. T., Mackinder, L., Schulz, K. G. et al. (2013). Dissecting the impact of CO2 and pH on the mechanisms of photosynthesis and calcification in the coccolithophore Emiliania huxleyi. New Phytologist 199: 121134.CrossRefGoogle ScholarPubMed
Barton, S., Jenkinson, J., Buckling, A. et al. (2020). Evolutionary temperature compensation of carbon fixation of marine phytoplankton. Ecology Letters 23: 722733.CrossRefGoogle ScholarPubMed
Batten, D. J. & Lister, J. K. (1988). Evidence of freshwater dinoflagellates and other algae in the English Wealden (Early Cretaceous). Cretaceous Research 9: 171179.CrossRefGoogle Scholar
Beardall, J. & Giordano, M. (2002). Ecological implications of microalgal and cyanobacterial CCMs and their regulation. Functional Plant Biology 29: 335347.CrossRefGoogle ScholarPubMed
Beardall, J. & Raven, J. A. (2004). The potential effects of global climate change on microalgal photosynthesis, growth and ecology. Phycologia 43: 2640.CrossRefGoogle Scholar
Beardall, J., Stojkovic, S. & Larsen, S. (2009). Living in a high CO2 world: Impacts of global climate change on marine phytoplankton. Plant Ecology and Diversity 2: 191205.CrossRefGoogle Scholar
Beardall, J., Stojkovic, S. & Gao, K. (2014). Interactive effects of nutrient supply and other environmental factors on the sensitivity of marine primary producers to ultraviolet radiation: Implications for the impacts of global change. Aquatic Biology 22: 523.CrossRefGoogle Scholar
Behrenfeld, M. J., O’Malley, R. T., Siegel, D. A. et al. (2006). Climate-driven trends in contemporary ocean productivity. Nature 444: 752755.CrossRefGoogle ScholarPubMed
Bissinger, J. E., Montagnes, D. J. S., Sharples, J. et al. (2008). Predicting marine phytoplankton maximum growth rates from temperature: Improving on the Eppley curve using quantile regression. Limnology and Oceanography 53: 487493.CrossRefGoogle Scholar
Blank, C. E. & Sánchez-Baracaldo, P. (2010). Timing of morphological and ecological innovations in the cyanobacteria – a key to understanding the rise in atmospheric oxygen. Geobiology 8: 123.CrossRefGoogle ScholarPubMed
Boller, A. J., Thomas, P. J., Cavanaugh, C. M. et al. (2011). Low stable carbon isotope fractionation by coccolithophore RubisCO. Geochimica et Cosmochimica Acta 75: 72007207. https://doi.org/10.1016/J.GCA.2011.08.031.CrossRefGoogle Scholar
Britton, D., Mundy, C. N., McGraw, C. M. et al. (2019). Responses of seaweeds that use CO2 as their sole inorganic carbon source to ocean acidification: Differential effects of fluctuating pH but little benefit of CO2 enrichment. ICES Journal of Marine Science 12: 1234.Google Scholar
Brodie, J., Williamson, C. J., Smale, D. A. et al. (2014). The future of the northeast Atlantic benthic flora in a high CO2 world. Ecology and Evolution 4: 27872798.CrossRefGoogle Scholar
Brown, J. W. & Sorhanus, U. (2010). A molecular genetic timescale for the diversification of autotrophic stramenopiles (Ochrophyta): Substantive underestimation of putative fossil ages. PLOS ONE 5: e12759.CrossRefGoogle ScholarPubMed
Burkhardt, S., Amoroso, G., Riebesell, U. et al. (2001). CO2 and HCO3 uptake in marine diatoms acclimated to different CO2 concentrations. Limnology and Oceanography 46: 13781391.CrossRefGoogle Scholar
Capó-Bauçà, S., Iñiguez, C., Aguiló-Nicolau, P. et al. (2022). Correlative adaptation between Rubisco and CO2-concentrating mechanisms in seagrasses. Nature Plants 8: 706716.CrossRefGoogle Scholar
Carreira, C., Heldal, M. & Bratbak, G. (2012). Effect of increased pCO2 on phytoplankton–virus interactions. Biogeochemistry 113: 391397.Google Scholar
Chave, K. E., Deffeyes, K. S., Weyl, P. K. et al. (1962). Observations on the solubility of skeletal carbonates in aqueous solutions. Science 137: 3334.CrossRefGoogle ScholarPubMed
Chen, S. W., Beardall, J. & Gao, K. S. (2014). A red tide alga grown under ocean acidification up-regulates its tolerance to lower pH by increasing its photophysiological functions. Biogeosciences 11: 48294837.CrossRefGoogle Scholar
Chen, X. & Gao, K. (2004). Photosynthetic utilisation of inorganic carbon and its regulation in the marine diatom Skeletonema costatum. Functional Plant Biology 31: 10271033.CrossRefGoogle ScholarPubMed
Chen, S., Gao, K. & Beardall, J. (2015). Viral attack exacerbates the susceptibility of a bloom-forming alga to ocean acidification. Global Change Biology 21: 629636.CrossRefGoogle ScholarPubMed
Chisholm, S. W. (1992). Phytoplankton size. In: Falkowski, P. G. & Woodhead, A. D. (eds.) Primary Productivity and Biogeochemical Cycles in the Sea. Springer, Boston, MA, pp. 213237.CrossRefGoogle Scholar
Clegg, M. R., Maberly, S. C. & Jones, R. I. (2003). Behavioural responses of freshwater phytoplanktonic flagellates to a temperature gradient European Journal of Phycology 38: 195203.CrossRefGoogle Scholar
Coale, K. H., Johnson, K. S., Fitzwater, S. E. et al. (1996). A massive phytoplankton bloom induced by an ecosystem-scale iron fertilization experiment in the equatorial Pacific Ocean. Nature 383: 495501.CrossRefGoogle ScholarPubMed
Collins, S., Boyd, P. W. & Doblin, M. A. (2020). Evolution, microbes, and changing ocean conditions. Annual Review of Marine Science 12: 181208.CrossRefGoogle ScholarPubMed
Comeau, S., Cornwall, C. E., DeCarlo, T. M. et al. (2018). Similar controls on calcification under ocean acidification across unrelated coral reef taxa. Global Change Biology 24: 48574868.CrossRefGoogle ScholarPubMed
Cornwall, C. E., Hepburn, C. D., McGraw, C. M. et al. (2013). Diurnal fluctuations in seawater pH influence the response of a calcifying macroalga to ocean acidification. Proceedings of the Royal Society B 280: 20132201. https://doi.org/10.1098/rspb.2013.2201.CrossRefGoogle ScholarPubMed
Cornwall, C. E., Hepburn, C. D., Pritchard, D. W. et al. (2012). Carbon-use strategies in macroalgae: Differential responses to lowered pH and implications for ocean acidification. Journal of Phycology 48: 137144.CrossRefGoogle ScholarPubMed
Cornwall, C. E. & Hurd, C. L. (2020). Variability in the benefits of ocean acidification to photosynthetic rates of macroalgae without CO2-concentrating mechanisms. Marine and Freshwater Research 71: 275280.CrossRefGoogle Scholar
Cornwall, C. E., Revill, A. T. & Hurd, C. L. (2015). High prevalence of diffusive uptake of CO2 by macroalgae in a temperate subtidal ecosystem. Photosynthesis Research 124: 181190.CrossRefGoogle Scholar
Crawfurd, K. J., Raven, J. A., Wheeler, G. L. et al. (2011). The response of Thalassiosira pseudonana to long-term exposure to increased CO2 and decreased pH. PLOS ONE 6: e26695.CrossRefGoogle ScholarPubMed
Del Cortona, A., Jackson, C., Bucchini, F. et al. (2020). Neoproterozoic origin and multiple transitions to macroscopic growth in green seaweeds. Proceedings of the National Academy of Sciences USA 117: 23312559.CrossRefGoogle ScholarPubMed
den Hartog, C. (1970). The Sea-Grasses of the World. North-Holland Publishing Company, Amsterdam.Google Scholar
Doney, S. C. (2006). Plankton in a warmer world. Nature 444: 695696.CrossRefGoogle Scholar
Dunstan, P. K., Foster, S. D., King, R. et al. (2018). Global patterns of change and variation in sea surface temperature and chlorophyll a. Scientific Reports 8: article 14624.CrossRefGoogle ScholarPubMed
Eppley, R. W. (1972). Temperature and phytoplankton growth in the sea. Fishery Bulletin 70: 10631085.Google Scholar
Eppley, R. W. & Peterson, B. J. (1979). Particulate organic matter flux and planktonic new production in the deep ocean. Nature 282: 677680.CrossRefGoogle Scholar
Falkowski, P. G. & Raven, J. A. (2007). Aquatic Photosynthesis. Princeton University Press, Princeton, NJ.CrossRefGoogle Scholar
Falkowski, P. G., Katz, M. E., Knoll, A. H. et al. (2004). The evolution of modern eukaryotic phytoplankton. Science 305: 354360.CrossRefGoogle ScholarPubMed
Finkel, Z. V., Beardall, J., Flynn, K. J. et al. (2010). Phytoplankton in a changing world: Cell size and elemental stoichiometry. Journal of Plankton Research 32: 119137.CrossRefGoogle Scholar
Gao, K., Beardall, J., Häder, D. P. et al. (2019). Effects of ocean acidification on marine photosynthetic organisms under the concurrent influences of warming, UV radiation and deoxygenation. Frontiers in Marine Science 6: 322. https://doi.org/10.3389/fmars.2019.00322.CrossRefGoogle Scholar
Gao, K. & Campbell, D. A. (2014). Photophysiological responses of marine diatoms to elevated CO2 and decreased pH: A review. Functional Plant Biology 41: 449459.CrossRefGoogle ScholarPubMed
Gao, K., Helbling, E. W., Häder, D. P. et al. (2012a). Responses of marine primary producers to interactions between ocean acidification, solar radiation, and warming. Marine Ecology Progress Series 470: 167189.CrossRefGoogle Scholar
Gao, K., Xu, J., Gao, G. et al. (2012b). Rising CO2 and increased light exposure synergistically to reduce marine primary productivity. Nature Climate Change 2: 519523.CrossRefGoogle Scholar
Gerten, D. & Adrian, R. (2000). Climate-driven changes in spring plankton dynamics and the sensitivity of shallow polymictic lakes to the North Atlantic Oscillation. Limnology and Oceanography 45: 10581066.CrossRefGoogle Scholar
Godoi, R. H. M., Aerts, K., Harlay, J. et al. (2009). Organic surface coating on coccolithophores – Emiliania huxleyi: Its determination and implication in the marine carbon cycle. Microchemical Journal 91: 265271.CrossRefGoogle Scholar
Goffart, A., Hecq, J. H. & Legendre, L. (2002). Changes in the development of the winter-spring phytoplankton bloom in the Bay of Calvi (NW Mediterranean) over the last two decades: A response to changing climate. Marine Ecology Progress Series 236: 4560.CrossRefGoogle Scholar
Goldman, J. C. & Carpenter, E. J. (1974). A kinetic approach to the effect of temperature on algal growth. Limnology and Oceanography 19: 756766.CrossRefGoogle Scholar
Hall-Spencer, J. M., Rodolfo-Metalpa, R. et al. (2008). Volcanic carbon dioxide vents show ecosystem effects of ocean acidification. Nature 454: 9699.CrossRefGoogle ScholarPubMed
Harley, C. D. G., Anderson, K. M., Demes, K. W. et al. (2012). Effects of climate change on global seaweed communities. Journal of Phycology 48: 10641078.CrossRefGoogle ScholarPubMed
Hayashida, H., Matear, R. J. & Strutton, P. G. (2020). Background nutrient concentration determines phytoplankton bloom response to marine heatwaves. Global Change Biology 26: 48004811.CrossRefGoogle ScholarPubMed
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. Global Change Biology 17: 24882497.CrossRefGoogle Scholar
Heraud, P., Roberts, S., Shelly, K. et al. (2005). Interactions between UVB exposure and phosphorus nutrition: II. Effects on rates of damage and repair. Journal of Phycology 41: 12121218.CrossRefGoogle Scholar
Hervé, V., Derr, J., Douady, S. et al. (2012). Multiparametric analyses reveal the pH-dependence of silicon biomineralization in diatoms. PLOS ONE 7: e46722. https://doi.org/10.1371/JOURNAL.PONE.0046722.CrossRefGoogle ScholarPubMed
Hurd, C. L., Beardall, J., Comeau, S. et al. (2020). Ocean acidification as a multiple driver: How interactions between changing seawater carbonate parameters affect marine life Marine and Freshwater Research 71: 263274.CrossRefGoogle Scholar
Hurd, C. L., Hepburn, C. D., Currie, K. I. et al. (2009). Testing methods of ocean acidification on algal metabolism: Consideration for experimental designs. Journal of Phycology 45: 12361251.CrossRefGoogle ScholarPubMed
Hutchins, D. A. & Fu, F. X. (2017). Microorganisms and ocean global change. Nature Microbiology 2: 17508. https://doi.org/10.1038/nmicrobiol.2017.58.CrossRefGoogle ScholarPubMed
Hutchins, D. A., Walworth, N. G., Webb, E. A. et al. (2015). Irreversibly increased nitrogen fixation in Trichodesmium experimentally adapted to elevated carbon dioxide. Nature Communications 6: 8155. https://doi.org/10.1038/ncomms9155.CrossRefGoogle ScholarPubMed
IPCC (2014). Climate 2014: Synthesis Report. In: Core, Writing Team, Pachauri, R. K. & Meyer, L. A. (eds.) Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. IPCC, Geneva, Switzerland, p. 151.Google Scholar
Iverson, L. L., Winkel, A., Baastrup-Spohr, L. et al. (2019). Catchment properties and the photosynthetic trait composition of freshwater plant communities. Science 366: 878881.CrossRefGoogle Scholar
Ji, X., Verspagen, J. M. H., Stomp, M. et al. (2017). Competition between cyanobacteria and green algae at low versus elevated CO2: Who will win, and why? Journal of Experimental Botany 68: 38153828.CrossRefGoogle ScholarPubMed
Koch, M., Bowes, G., Ross, C. et al. (2013). Climate change and ocean acidification effects on seagrasses and marine macroalgae. Global Change Biology 19: 103132.CrossRefGoogle ScholarPubMed
Kosten, S., Huszar, V. L. M., Bécares, E. et al. (2012). Warmer climates boost cyanobacterial dominance in shallow lakes. Global Change Biology 18: 118−126.CrossRefGoogle Scholar
Kottmeier, D. M., Rokitta, S. D. & Rost, B. (2016). Acidification, not carbonation, is the major regulator of carbon fluxes in the coccolithophore Emiliania huxleyi. New Phytologist 211: 126137.CrossRefGoogle Scholar
Kram, S. L., Price, N. N., Donham, E. M. et al. (2016). Variable responses of temperate calcified and fleshy macroalgae to elevated pCO2 and warming. ICES Journal of Marine Science 73: 693703.CrossRefGoogle Scholar
Krumhardt, K. M., Lovenduski, N. S., Long, M. C. et al. (2019). Coccolithophore growth and calcification in an acidified ocean: Insights from community earth system model simulations. Journal of Advances in Modeling Earth Systems 11: 14181437.CrossRefGoogle Scholar
Krumhansl, K. A. & Scheibling, R. E. (2012). Production and fate of kelp detritus. Marine Ecology Progress Series 467: 281302.CrossRefGoogle Scholar
Kübler, J. E. & Dudgeon, S. R. (2015). Predicting effects of ocean acidification and warming on algae lacking carbon concentrating mechanisms. PLOS ONE 10: e0132806. https://doi.org/10.1371/JOURNAL.PONE.0132806.CrossRefGoogle ScholarPubMed
Kübler, J. E., Johnston, A. M. & Raven, J. A. (1999). The effects of reduced and elevated CO2 and O2 on the seaweed Lomentaria articulata. Plant, Cell and Environment 22: 13031310. https://doi.org/10.1046/J.1365–3040.1999.00492.X.CrossRefGoogle Scholar
Kübler, J. E. & Raven, J. A. (1994). Consequences of light limitation for inorganic carbon acquisition in three rhodophytes. Marine Ecology Progress Series 110: 203209.CrossRefGoogle Scholar
Kübler, J. E. & Raven, J. A. (1995). The interaction between inorganic carbon supply and light supply in Palmaria palmata (Rhodophyta). Journal of Phycology 31: 369375.CrossRefGoogle Scholar
Kuffner, I. B., Andersson, A. J., Jokiel, P. L. et al. (2008). Decreased abundance of crustose coralline algae due to ocean acidification. Nature Geoscience 1: 114117.CrossRefGoogle Scholar
Larkum, A. W. D., Davey, P. A., Kuo, J. et al. (2017). Carbon-concentrating mechanisms in seagrasses. Journal of Experimental Botany 68: 37733784.CrossRefGoogle ScholarPubMed
Larsen, J. B., Larsen, A., Thyrhaug, R. et al. (2008). Response of marine viral populations to a nutrient induced phytoplankton bloom at different pCO2 levels. Biogeosciences 5: 523533.CrossRefGoogle Scholar
Laufkötter, C., Vogt, T., Grüber, N. et al. (2015). Drivers and uncertainties of future global marine production in marine ecosystem models. Biogeosciences 12: 69556984.CrossRefGoogle Scholar
Li, F., Beardall, J., Collins, S. et al. (2017). Decreased photosynthesis and growth with reduced respiration in the model diatom Phaeodactylum tricornutum grown under elevated CO2 over 1800 generations. Global Change Biology 23: 127137.CrossRefGoogle ScholarPubMed
Li, G. & Campbell, D. A. (2013). Rising CO2 interacts with growth light and growth rate to alter photosystem II photoinactivation of the coastal diatom Thalassiosira pseudonana. PLOS ONE 8: e55562.Google ScholarPubMed
Li, W. K. W., McLaughlin, F., Lovejoy, C. et al. (2016). Smallest algae thrive as the Arctic ocean freshens. Science 326: 539. https://doi.org/10.1126/science.1179798.CrossRefGoogle Scholar
Liu, Y.-W., Eagle, R. A., Aciego, S. M. et al. (2018). A coastal coccolithophore maintains pH homeostasis and switches carbon sources in response to ocean acidification. Nature Communications 9: 2857. https://doi.org/10.1038/S41467–018–04463–7.CrossRefGoogle ScholarPubMed
Lohbeck, K. T., Riebesell, U., & Reusch, T. B. H. (2012). Adaptive evolution of a key phytoplankton species to ocean acidification. Nature Geoscience 5: 346351.CrossRefGoogle Scholar
Low-Décarie, E., Fussman, G. F. & Bell, G. (2011). The effect of elevated CO2 on growth and competition in experimental phytoplankton communities. Global Change Biology 17: 25252535.CrossRefGoogle Scholar
Low-Décarie, E., Jewell, M. D., Fussmann, G. F. et al. (2013). Long-term culture at elevated atmospheric CO2 fails to evoke specific adaptation in seven freshwater phytoplankton species. Proceedings of the Royal Society Biology 280: 20122598. http://dx.doi.org/10.1098/rspb.2012.2598.CrossRefGoogle ScholarPubMed
Low–Décarie, E., Bell, G. & Fussmann, G. (2015). CO2 alters community composition and response to nutrient enrichment of freshwater phytoplankton. Oecologia 177: 875883.CrossRefGoogle ScholarPubMed
Lu, J., Vecchi, G. A. & Reichler, T. (2007). Expansion of the Hadley cell under global warming. Geophysical Research Letters 34: L06805. https://doi.org/10.1029/2006GL028443.Google Scholar
Lyman, J. M., Good, S. A., Gourestki, V. V. et al. (2010). Robust warming of the global ocean. Nature 465: 334337.CrossRefGoogle Scholar
Maberly, S.C. (1990). Exogenous sources of inorganic carbon for photosynthesis by marine macroalgae. Journal of Phycology 26: 439449.CrossRefGoogle Scholar
Maberly, S. C. (1996). Diel, episodic and seasonal changes in pH and concentrations of inorganic carbon in a productive lake. Freshwater Biology 35: 579598.CrossRefGoogle Scholar
Maberly, S. C. and Gontero, B. (2017). Ecological imperatives of aquatic CO2-concentrating mechanisms. Journal of Experimental Botany 68: 878881.CrossRefGoogle ScholarPubMed
Maberly, S. C., O’Donnell, R. A., Woolway, R. I. et al. (2020). Global lake thermal regions shift under climate change. Nature Communications 11: article 1232. https://doi.org/10.1038/s.41467–020–15108-z.CrossRefGoogle ScholarPubMed
Malerba, M., Marshall, D., Palacios, M. et al. (2021). Cell size influences inorganic carbon acquisition in artificially selected phytoplankton. New Phytologist 229: 26472659.CrossRefGoogle ScholarPubMed
McCarthy, A., Rogers, S. P., Duffy, S. J. et al. (2012). Elevated carbon dioxide differentially alters the photophysiology of Thalassiosira pseudonana (Bacillariophyceae) and Emiliania huxleyi (Haptophyta). Journal of Phycology 48: 635646.CrossRefGoogle ScholarPubMed
McMinn, A., Müller, M. N., Martin, A. et al. (2014). The response of Antarctic sea ice algae to changes in pH and CO2. PLOS ONE 9: e86984.CrossRefGoogle Scholar
Meyer, M. & Griffiths, H. (2013). Origins and diversity of eukaryotic CO2-concentrating mechanisms: Lessons for the future. Journal of Experimental Botany 64: 769786.CrossRefGoogle ScholarPubMed
Paerl, H. W. & Huisman, J. (2008). Blooms like it hot. Science 320: 5758.CrossRefGoogle ScholarPubMed
Pardew, J., Blanco Pimental, M. & Low-Décarie, E. (2018). Predictable ecological response to rising CO2 of a community of marine phytoplankton. Ecology and Evolution 8: 42924302.CrossRefGoogle ScholarPubMed
Pierangelini, M., Stojkovic, S., Orr, P. T. et al. (2014). Elevated CO2 causes changes in the photosynthetic apparatus of a toxic cyanobacterium, Cylindrospermopsis raciborskii. Journal of Plant Physiology 171: 10911098.CrossRefGoogle ScholarPubMed
Ponce-Toledo, R. I., Deschamps, P., Lopez-Garcia, P. et al. (2017). An early-branching freshwater cyanobacterium at the origin of plastids. Current Biology 27: 386391.CrossRefGoogle ScholarPubMed
Raven, J. A. (2011). Praeger Review: Effects on marine algae of changed seawater chemistry with increasing atmospheric CO2. Biology and Environment 111: 117.CrossRefGoogle Scholar
Raven, J. A. (2017). The possible roles of algae in restricting the increase in atmospheric CO2 and global temperature. European Journal of Phycology 52: 506522.CrossRefGoogle Scholar
Raven, J. A. (2018). Blue carbon: Past present and future, with emphasis on macroalgae. Biology Letters 14: 20180336. https://doi.org/10.1098/vsbl.2018.0036.CrossRefGoogle ScholarPubMed
Raven, J. A. (2023). Distribution and functions of calcium mineral deposits in photosynthetic organisms. In: Lüttge, U., Cánovas, F. M., Risueño, M.-C. et al. (eds.) Progress in Botany, Vol. 84. Springer, Cham, pp. 293326.Google Scholar
Raven, J. A. & Beardall, J. (2020). Energizing the plasmalemma of marine photosynthetic organisms; the role of primary active transport. Journal of the Marine Biological Association of the UK 100: 333346.CrossRefGoogle Scholar
Raven, J. A., Beardall, J. & Giordano, M. (2014). Energy costs of carbon dioxide concentrating mechanisms in aquatic organisms. Photosynthesis Research 121: 111124.CrossRefGoogle ScholarPubMed
Raven, J. A., Beardall, J. & Sánchez-Baracaldo, P. (2017). The possible evolution and future of CO2 concentrating mechanisms. Journal of Experimental Botany 68: 37013716.CrossRefGoogle ScholarPubMed
Raven, J. A., Caldeira, K., Elderfield, H. et al. (2005). Ocean acidification due to increasing atmospheric carbon dioxide. The Royal Society of London Report 12/05. The Royal Society, London, pp. vii–60. https://royalsociety.org/topics-policy/publications/2005/ocean-acidification/.Google Scholar
Raven, J. A. & Crawfurd, K. (2012). Environmental controls on coccolithophore calcification. Marine Ecology Progress Series 470: 137166.CrossRefGoogle Scholar
Raven, J. A., Gobler, C. J. & Hansen, P. J. (2020a). Dynamic CO2 and pH levels in coastal, estuarine, and inland waters: Theoretical and observed effects on harmful algal blooms. Harmful Algae 91: Article 101594. https://doi.org/10.1016/j.hal.2019.03.012.CrossRefGoogle ScholarPubMed
Raven, J. A. & Johnston, A. M. (1991). Mechanisms of inorganic-carbon acquisition in marine phytoplankton and their implications for the use of other resources. Limnology and Oceanography 36: 17011714.CrossRefGoogle Scholar
Raven, J. A., Suggett, D. J. & Giordano, M. (2020b). Inorganic carbon concentrating mechanisms in free-living and symbiotic diatoms and chromerids. Journal of Phycology 56: 13771397.CrossRefGoogle Scholar
Raymond, P. A., Hartmann, J., Lauerwald, R. et al. (2013). Global carbon dioxide emissions from inland waters. Nature 503: 355359.CrossRefGoogle ScholarPubMed
Reusch, T. B. H. & Boyd, P. W. (2012). Experimental evolution meets marine phytoplankton. Evolution 67: 18491859.CrossRefGoogle Scholar
Riebesell, U. & Tortell, P. D. (2011). Effects of ocean acidification on pelagic organisms and ecosystems. In: Gattuso, J.-P. & Hansson, L. (eds.) Ocean Acidification, Oxford University Press, Oxford, pp. 99121.Google Scholar
Riebesell, U., Zondervan, I., Rost, B. et al. (2000). Reduced calcification of marine plankton in response to increased atmospheric CO2. Nature 407: 364367.CrossRefGoogle ScholarPubMed
Roleda, M. Y., Morris, J. N., McGraw, C. M. et al. (2012). Ocean acidification and seaweed reproduction: Increased CO2 ameliorates the negative effect of lowered pH on meiospore germination in the giant kelp Macrocystis pyrifera (Laminariales, Phaeophyceae). Global Change Biology 18: 854864.CrossRefGoogle Scholar
Rost, B., Zondervan, I. & Wolf-Gladrow, D. (2008). Sensitivity of phytoplankton to future changes in ocean carbonate chemistry: Current knowledge, contradictions and research directions. Marine Ecology Progress Series 373: 227237.CrossRefGoogle Scholar
Sánchez-Baracaldo, P., Raven, J. A., Pisani, D. et al. (2017). Early photosynthetic eukaryotes inhabited low-salinity habitats. Proceeding of the National Academy of Sciences USA 114: E7737E7745.CrossRefGoogle ScholarPubMed
Sarmiento, J. L., Slater, R., Barber, R. et al. (2004). Response of ocean ecosystems to climate warming. Global Biogeochemical Cycles 18: GB3003. https://doi.org/10.1029/2003GB002134.CrossRefGoogle Scholar
Schaum, C. E. & Collins, S. (2014). Plasticity predicts evolution in a marine alga. Proceedings of the Royal Society Biology 281: 20141486. http://dx.doi.org/10.1098/rspb.2014.1486.CrossRefGoogle Scholar
Schaum, C. E., Barton, S., Bestion, E. et al. (2017). Adaptation of phytoplankton to a decade of experimental warming linked to increased photosynthesis. Nature Ecology and Evolution 1: article 0094. https://doi.org/10.1038/s41559–017–0094.CrossRefGoogle ScholarPubMed
Schaum, C. E., Buckling, A., Smirnoff, N. et al. (2018). Environmental fluctuations accelerate molecular evolution of thermal tolerance in a marine diatom. Nature Communications 9: article 179. https://doi.org/10.1038/s41467–08–03906-s.Google Scholar
Schoenrock, K. M., Schram, J. B., Amsler, C. D. et al. (2016). Climate change confers a potential advantage to fleshy Antarctic crustose macroalgae over calcified species. Journal of Experimental Marine Biology and Ecology 474: 5866.CrossRefGoogle Scholar
Semesi, I. S., Kangwe, J. & Björk, M. (2009). Alterations in seawater pH and CO2 affect calcification and photosynthesis in the tropical coralline alga, Hydrolithon sp. (Rhodophyta). Estuarine Coastal and Shelf Science 84: 337341. https://doi.org/10.1016/j.ecss.2009.03.038.CrossRefGoogle Scholar
Sett, S., Schulz, K. G., Bach, L. T. et al. (2018). Shift towards larger diatoms in a natural phytoplankton assemblage under combined high-CO2 and warming conditions. Journal of Plankton Research 40: 391406.CrossRefGoogle Scholar
Shelly, K., Heraud, P. & Beardall, J. (2002). Nitrogen limitation in Dunaliella tertiolecta Butcher (Chlorophyceae) leads to increased susceptibility to damage by ultraviolet-B radiation but also increased repair capacity. Journal of Phycology 38: 18.CrossRefGoogle Scholar
Shelly, K., Roberts, S., Heraud, P. et al. (2005). Interactions between UVB exposure and phosphorus nutrition: I Effects on growth, phosphate-uptake and chlorophyll fluorescence. Journal of Phycology 41: 12041211.CrossRefGoogle Scholar
Siver, P. A., Velez, M., Chivetti, M. et al. (2018). Early freshwater diatoms from the Upper Cretaceous Battle Formation in Western Canada. Paleios 33: 525534.CrossRefGoogle Scholar
Smale, D. A. & Wernberg, T. (2013). Extreme climatic event drives range contraction of a habitat-forming species. Proceedings of the Royal Society London, B 280: 20122829.Google ScholarPubMed
Smol, J. P., Wolfe, A. P., Birks, H. J. B. et al. (2005). Climate-driven regime shifts in the biological communities of arctic lakes. Proceedings of the National Academy of Sciences (USA) 102: 43974402.CrossRefGoogle ScholarPubMed
Sobek, S., Tranvick, L. J. & Cole, J. J. (2005). Temperature independence of carbon dioxide supersaturation in global lakes. Global Biogeochemical Cycles 19: GB2003. https://doi.org/10.1029/2004GB002264.CrossRefGoogle Scholar
Stepien, C. C. (2015). Impacts of geography, taxonomy and functional group on inorganic carbon use patterns in marine macrophytes. Journal of Ecology 103: 13721383.CrossRefGoogle Scholar
Suzuki, Y. & Takahashi, M. (1995). Growth responses of several diatoms isolated from various environments to temperature. Journal of Phycology 31: 880888.CrossRefGoogle Scholar
Taylor, A. R., Brownlee, C. & Wheeler, G. (2017). Coccolithophore cell biology: Chalking up progress. Annual Review of Marine Science 9: 283310.CrossRefGoogle ScholarPubMed
Toggweiler, J. R. & Russell, J. (2008). Ocean circulation in a warming climate. Nature 451: 286288.CrossRefGoogle Scholar
Tortell, P. D., DiTullio, G. R., Digman, D. M. et al. (2002). CO2 effects on taxonomic composition and nutrient utilization in an Equatorial Pacific phytoplankton assemblage. Marine Ecology Progress Series 236: 3743.CrossRefGoogle Scholar
Traving, S. J., Clokie, M. R. & Middelboe, M. (2013). Increased acidification has a profound effect on the interactions between the cyanobacterium Synechococcus sp. WH7803 and its viruses. FEMS Microbiology Ecology 87: 133141.CrossRefGoogle Scholar
Trenberth, K. E., Jones, P. D., Ambenie, P. et al. (2007). Observations: Surface and atmospheric climate change. In: Qin, D., Manning, M., Chen, Z. et al. (eds.) Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, pp. 235336.Google Scholar
Tuya, F., Cacabelos, E., Duarte, P. et al. (2012). Patterns of landscape and assemblage structure along a latitudinal gradient in ocean climate. Marine Ecology Progress Series 466: 919.CrossRefGoogle Scholar
Van de Poll, W., Leeuwe, M. A., Roggeveld, J. et al. (2005). Nutrient limitation and high irradiance acclimation reduce PAR and UV-induced viability loss in Antarctic diatom Chaetoceros brevis (Bacillariophyceae). Journal of Phycology 41: 840−850.CrossRefGoogle Scholar
van der Loos, L. M., Schmid, M., Leal, P. P. et al. (2019). Responses of macroalgae to CO2 enrichment cannot be inferred solely from their inorganic carbon uptake strategy. Ecology and Evolution 9: 125140.CrossRefGoogle ScholarPubMed
Verschoor, A. M., van Dijk, M. A., Huisman, J. et al. (2013). Elevated CO2 concentrations affect the elemental stoichiometry and species composition of an experimental phytoplankton community. Freshwater Biology 58: 597611.CrossRefGoogle Scholar
Wang, Y., Fan, X., Gao, G. et al. (2020). Decreased motility of flagellated microalgae long-term acclimated to CO2-induced acidified waters. Nature Climate Change 10: 561567.CrossRefGoogle Scholar
Wehrli, B. (2013). Conduits of the carbon cycle. Nature 503: 346347.CrossRefGoogle ScholarPubMed
Waycott, M., Procaccini, G., Les, D. H. et al. (2006). Seagrass evolution, ecology and conservation: a genetic perspective. In: Larkum, A. W. D., Orth, R. J. & Duarte, C. (eds.) Seagrasses: Biology, Ecology and Conservation. Springer, Dordrecht, pp. 2550.Google Scholar
Wellman, C. H., Graham, L. E. & Lewis, L. A. (2019). Filamentous green algae from the Early Devonian Rhynie Chert. PalZ 93: 387393.CrossRefGoogle Scholar
Wells, M. L., Trainer, V. L., Smayda, T. J. et al. (2015). Harmful algal blooms (HAB) and climate change; what do we know and where do we go from here? Harmful Algae 49: 6893.CrossRefGoogle Scholar
Wernberg, T., Smale, D. A., Tuya, F. et al. (2013). An extreme climatic event alters marine ecosystem structure in a global biodiversity hotspot. Nature Climate Change 3: 7882.CrossRefGoogle Scholar
Willis, A., Chuang, A. W., Orr, P. T. et al. (2019). Subtropical freshwater phytoplankton show a greater response to increased temperature than to increased pCO2. Harmful Algae 90: article 101705. https:doi.org/10.1016/j.hal.2019.101705.CrossRefGoogle Scholar
Wiltshire, K. H. & Manly, B. F. J. (2004). The warming trend at Helgoland Roads, North Sea: phytoplankton response. Helgoland Marine Research 58: 269273.CrossRefGoogle Scholar
Wohlers, J., Engel, A., Zöllner, E. et al. (2009). Changes in biogenic carbon flow in response to sea surface warming. Proceedings of the National Academy of Sciences USA 106: 70677072.CrossRefGoogle ScholarPubMed
Wong, C.-Y., Teoh, M.-L., Phang, S.-M. et al. (2015). Interactive effects of temperature and UV radiation on photosynthesis of Chlorella spp. from polar, temperate and tropical environments: Differential impacts on repair and damage. PLOS ONE 10: e0139469. https://doi.org/10.1371/journal.pone.0139469.CrossRefGoogle Scholar
Wu, Y., Campbell, D. A., Irwin, A. J. et al. (2014). Ocean acidification enhances the growth rate of larger diatoms. Limnology and Oceanography 59: 10271034.CrossRefGoogle Scholar
Wu, Y., Gao, K. & Riebesell, U. (2010). CO2-induced seawater acidification affects physiological performance of the marine diatom Phaeodactylum tricornutum. Biogeosciences 7: 29152923.CrossRefGoogle Scholar
Zondervan, I. (2007). The effects of light, macronutrients, trace metals and CO2 on the production of calcium carbonate and organic carbon in coccolithophores: A review. Deep-Sea Research II 54: 521537.CrossRefGoogle Scholar

References

Ahn, G. N., Kim, K. N., Cha, S. H. et al. (2007). Antioxidant activities of phlorotannins purified from Ecklonia cava on free radical scavenging using ESR and H2O2-mediated DNA damage. European Food Research and Technology 226: 7179.CrossRefGoogle Scholar
Aitzetmüller, K., Strain, H. H., Svec, W. A. et al. (1969). Loroxanthin, a unique xanthophyll from Scenedesmus obliquus and Chlorella vulgaris. Phytochemistry 8: 17611770.CrossRefGoogle Scholar
Alboresi, A., Ballottari, M., Hienerwadel, R. et al. (2009). Antenna complexes protect Photosystem I from photoinhibition. BMC Plant Biology 9: 71.CrossRefGoogle ScholarPubMed
Asada, K. (1992). Ascorbate peroxidase – A hydrogen peroxide-scavenging enzyme in plants. Physiologia Plantarum 85: 231241.CrossRefGoogle Scholar
Asada, K. (1999). The water cycle in chloroplasts: Scavenging of active oxygens and dissipation of excess photons. Annual Review of Plant Physiology and Plant Molecular Biology 50: 601639.CrossRefGoogle ScholarPubMed
Ayoub, L. M., Hallock, P., Coble, P. G. et al. (2012). MAA-like absorbing substances in Florida Keys phytoplankton vary with distance from shore and CDOM: Implications for coral reefs. Journal of Experimental Marine Biology and Ecology 420–421: 9198.CrossRefGoogle Scholar
Baroli, I. & Niyogi, K. K. (2000). Molecular genetics of xanthophylls-dependent photoprotection in green algae and plants. Philosophical Transactions of the Royal Society B 355: 13851394.CrossRefGoogle ScholarPubMed
Behrenfeld, M. J., O’Malley, R. T., Siegel, D. A. et al. (2006). Climate-driven trends in contemporary ocean productivity. Nature 444: 752755.CrossRefGoogle ScholarPubMed
Bouchard, J. N., Roy, S. & Campbell, D. A. (2006). UVB effects on the photosystem II-D1 protein of phytoplankton and natural phytoplankton communities. Photochemistry and Photobiology 82: 936951.CrossRefGoogle ScholarPubMed
Bray, C. M. & West, C. E. (2005). DNA repair mechanisms in plants: Crucial sensors and effectors for the maintenance of genome integrity. New Phytologist 168: 511528.CrossRefGoogle ScholarPubMed
Buma, A. G. J., Boelen, P. & Jeffrey, W. H. (2003). UVR-induced DNA damage in aquatic organisms. In Helbling, E. W. & Zagarese, H. E. (eds.) UV Effects in Aquatic Organisms and Ecosystems. The Royal Society of Chemistry, Cambridge, pp. 291327.Google Scholar
Buma, A. G. J., Wright, S. W., van den Enden, R. et al. (2006). PAR acclimation and UVBR-induced DNA damage in Antarctic marine microalgae. Marine Ecology Progress Series 315: 3342.CrossRefGoogle Scholar
Buma, A. G. J., Visser, R. J., Van de Poll, W. et al. (2009). Wavelength-dependent xanthophyll cycle activity in marine microalgae exposed to natural ultraviolet radiation. European Journal of Phycology 44: 515524.CrossRefGoogle Scholar
Cabrerizo, M. J., Carrillo, P., Villafañe, V. E. et al. (2014). Current and predicted global change impacts of UVR, temperature and nutrient inputs on photosynthesis and respiration of key marine phytoplankton species. Journal of Experimental Marine Biology and Ecology 461: 371380.CrossRefGoogle Scholar
Carrillo, P., Medina-Sánchez, J. M., Villar-Argaiz, M. et al. (2017). Vulnerability of mixotrophic algae to nutrient pulses and UVR in an oligotrophic Southern and Northern Hemisphere lake. Scientific Reports 7: 6333.CrossRefGoogle Scholar
Chen, H., Guan, W., Zeng, G. et al. (2015). Alleviation of solar ultraviolet radiation (UVR)-induced photoinhibition in diatom Chaetoceros curvisetus by ocean acidification. Journal of the Marine Biological Association of the United Kingdom 95: 661667.CrossRefGoogle Scholar
Chen, S. & Gao, K. (2011). Solar ultraviolet radiation and CO2-induced ocean acidification interacts to influence the photosynthetic performance of the red tide alga Phaeocystis globosa (Prymnesiophyceae). Hydrobiologia 675: 105117.CrossRefGoogle Scholar
Cruces, E., Rautenberger, R., Rojas-Lillo, Y. et al. (2017). Physiological acclimation of Lessonia spicata to diurnal changing PAR and UV radiation: Differential regulation among downregulation of photochemistry, ROS scavenging activity and phlorotannins as major photoprotective mechanisms. Photosynthesis Research 131: 145157.CrossRefGoogle ScholarPubMed
Delgado-Molina, J. A., Carrillo, P., Medina-Sánchez, J. M. et al. (2009). Interactive effects of phosphorus loads and ambient ultraviolet radiation on the algal community in a high-mountain lake. Journal of Plankton Research 31: 619634.CrossRefGoogle Scholar
Demming-Adams, B. & Adams, W. W. (1992). Photoprotection and other responses of plants to high light stress. Annual Review of Plant Physiology 43: 599626.CrossRefGoogle Scholar
Deitrick, R. & Goldblatt, C. (2023) Effects of ozone levels on climate through Earth history. Climate of the Past 19: 12011218.CrossRefGoogle Scholar
Domingues, R. B., Guerra, C. C., Barbosa, A. B. et al. (2014). Effects of ultraviolet radiation and CO2 increase on winter phytoplankton assemblages in a temperate coastal lagoon. Journal of Plankton Research 36: 672684.CrossRefGoogle Scholar
Falkowski, P. G., Scholes, R. J., Boyle, E. et al. (2000). The global carbon cycle: A test of our knowledge of Earth as a system. Science 290: 291296.CrossRefGoogle ScholarPubMed
Farahin, A. W., Yusoff, F. M., Nagao, N. et al. (2016). Phenolic content and antioxidant activity of Tetraselmis tetrathele (West) Butcher 1959 cultured in annular photobioreactor. Journal of Environmental Biology 37: 631639.Google ScholarPubMed
Finkel, Z. V., Beardall, J., Flynn, K. J. et al. (2010). Phytoplankton in a changing world: Cell size and elemental stoichiometry. Journal of Plankton Research 32: 119137.CrossRefGoogle Scholar
Fiorda Giordanino, M. V., Strauch, S. M., Villafane, V. E. et al. (2011). Influence of temperature and UVR on photosynthesis and morphology of four species of cyanobacteria. Journal of Photochemistry and Photobiology B: Biology 103: 6877.CrossRefGoogle Scholar
Gao, K., Li, P., Watanabe, T. et al. (2008). Combined effects of ultraviolet radiation and temperature on morphology, photosynthesis and DNA of Arthrospira (Spirulina) platensis (Cyanophyta). Journal of Phycology 44: 777786.CrossRefGoogle ScholarPubMed
Gao, K., Wu, Y., Li, G. et al. (2007). Solar UV radiation drives CO2 fixation in marine phytoplankton: A double-edged sword. Plant Physiology 144: 5459.CrossRefGoogle ScholarPubMed
Gao, K. S., Ruan, Z., Villafañe, V. E. et al. (2009). Ocean acidification exacerbates the effect of UV radiation on the calcifying phytoplankter Emiliania huxleyi. Limnology and Oceanography 54: 18551862.CrossRefGoogle Scholar
Garcia-Pichel, F. (1994). A model for internal self-shading in planktonic organisms and its implications for the usefulness of ultraviolet sunscreens. Limnology and Oceanography 39: 17041717.CrossRefGoogle Scholar
Grotjohann, I., Jolley, C. & Fromme, P. (2004). Evolution of photosynthesis and oxygen evolution: Implications from the structural comparison of Photosystems I and II. Chemical Physics 6: 47434753.Google Scholar
Gruber, A., Roleda, M. Y., Bartsch, I. et al. (2011). Sporogenesis under UVR in Laminaria digitata (Phaeophyceae) reveals protection of photosensitive meiospores within soral tissue: physiological and anatomical evidence. Journal of Phycology 47: 603614.CrossRefGoogle ScholarPubMed
Grunewald, K., Hirschberg, J. & Hagen, C. (2001). Ketocarotenoid biosynthesis outside of plastids in the unicellular green alga Haematococcus pluvialis. Journal of Biological Chemistry 276: 60236029.CrossRefGoogle ScholarPubMed
Häder, D.-P., Richter, P. R., Villafañe, V. E. et al. (2014a). Influence of light history on the photosynthetic and motility responses of Gymnodinium chlorophorum exposed to UVR and different temperatures. Journal of Photochemistry and Biology B: Biology 138: 273281.CrossRefGoogle ScholarPubMed
Häder, D.-P., Williamson, C. E., Wängberg, S.-A. et al. (2015). Effects of UV radiation on aquatic ecosystems and interactions with other environmental factors. Photochemical and Photobiological Sciences 14: 108126.CrossRefGoogle ScholarPubMed
Häder, D. P., Villafañe, V. E. & Helbling, E. W. (2014b). Productivity of aquatic primary producers under global climate change. Photochemical and Photobiological Sciences 13: 13701392.CrossRefGoogle ScholarPubMed
Halac, S. R., Villafañe, V. E. & Helbling, E. W. (2010). Temperature benefits the photosynthetic performance of the diatoms Chaetoceros gracilis and Thalassiosira weissflogii when exposed to UVR. Journal of Photochemistry and Photobiology, B: Biology 101: 196205.CrossRefGoogle ScholarPubMed
Hanelt, D. (1996). Photoinhibition of photosynthesis in marine macroalgae. Scientia Marina 60: 243248.Google Scholar
Hanelt, D. & Roleda, M. Y. (2009). UVB radiation may ameliorate photoinhibition in specific shallow-water tropical marine macrophytes. Aquatic Botany 91: 612.CrossRefGoogle Scholar
Helbling, E. W., Villafañe, V. E. & Holm-Hansen, O. (1994). Effects of ultraviolet radiation on Antarctic marine phytoplankton photosynthesis with particular attention to the influence of mixing. In Weiler, C. S. & Penhale, P. A. (eds.) Ultraviolet Radiation in Antarctica: Measurements and Biological Effects. American Geophysical Union, Washington, DC, pp. 207227.CrossRefGoogle Scholar
Helbling, E. W., Villafañe, V. E. & Barbieri, E. S. (2001). Sensitivity of winter phytoplankton communities from Andean lakes to ultraviolet-B radiation. Revista Chilena de Historia Natural 74: 273282.CrossRefGoogle Scholar
Helbling, E. W., Banaszak, A. T. & Villafañe, V. E. (2015a). Global change feed-back inhibits cyanobacterial photosynthesis. Scientific Reports 5: 14514. https://doi.org/10.1038/srep14514.CrossRefGoogle Scholar
Helbling, E. W., Banaszak, A. T. & Villafañe, V. E. (2015b). Differential responses of two phytoplankton communities from the Chubut river estuary (Patagonia, Argentina) to the combination of UVR and elevated temperature. Estuaries and Coasts 38: 11341146.CrossRefGoogle Scholar
Helbling, E. W., Villafañe, V. E., Ferrario, M. E. et al. (1992). Impact of natural ultraviolet radiation on rates of photosynthesis and on specific marine phytoplankton species. Marine Ecology Progress Series 80: 89100.CrossRefGoogle Scholar
Helbling, E. W., Pérez, D. E., Medina, C. D. et al. (2010). Phytoplankton distribution and photosynthesis dynamics in the Chubut River estuary (Patagonia, Argentina) throughout tidal cycles. Limnology and Oceanography 55: 5565.CrossRefGoogle Scholar
Helbling, E. W., Buma, A. G. J., Boelen, P. et al. (2011). Increase in Rubisco activity and gene expression due to elevated temperature partially counteracts ultraviolet radiation-induced photoinhibition in the marine diatom Thalassiosira weissflogii. Limnology and Oceanography 56: 13301342.CrossRefGoogle Scholar
Helbling, E. W., Carrillo, P., Medina-Sánchez, J. M. et al. (2013). Interactive effects of vertical mixing, nutrients and ultraviolet radiation: In situ photosynthetic responses of phytoplankton from high mountain lakes in Southern Europe. Biogeosciences 10: 10371050.CrossRefGoogle Scholar
Hessen, D. O. (2008). Solar radiation and the evolution of life. In Bjertness, E. (ed.) Solar Radiation and Human Health. The Norwegian Academy of Science and Letters, Oslo, pp. 123136.Google Scholar
Holm-Hansen, O. & Lubin, D. (1994). Solar ultraviolet radiation: Effects on rates of CO2 fixation in marine phytoplankton. In Tolbert, N. E. & Preiss, J. (eds.) Regulation of Atmospheric CO2 and O2 by Photosynthetic Carbon Metabolism. Oxford University Press, New York, pp. 5574.Google Scholar
Holzinger, A., Di Piazza, L., Lütz, C. et al. (2011). Sporogenic and vegetative tissues of Saccharina latissima (Laminariales, Phaeophyceae) exhibit distinctive sensitivity to experimentally enhanced ultraviolet radiation: Photosynthetically active radiation ratio. Phycological Research 59: 221235.CrossRefGoogle Scholar
IPCC (2013). Climate Change 2013. The Physical Science Basis Cambridge University Press, New York, NY, pp. 1535.Google Scholar
Jahns, P. & Holzwarth, A. R. (2012). The role of the xanthophyll cycle and of lutein in photoprotection of photosystem II. Biochimica et Biophysica Acta 1817: 182193.CrossRefGoogle ScholarPubMed
Kieber, D. J., Peake, B. M. & Scully, N. M. (2003). Reactive oxygen species in aquatic ecosystems. In Helbling, E. W. & Zagarese, H. E. (eds.) UV Effects in Aquatic Organisms and Ecosystems. RSC, Cambridge, pp. 251288.Google Scholar
Kitidis, V., Stubbins, A. P., Uher, G. et al. (2006). Variability of chromophoric organic matter in surface waters of the Atlantic Ocean. Deep Sea Research II 53: 16661684.CrossRefGoogle Scholar
Kulk, G., De Vries, P., Van de Poll, W. et al. (2012). Temperature-dependent growth and photophysiology of prokaryotic and eukaryotic oceanic picophytoplankton. Marine Ecology Progress Series 466: 4355.CrossRefGoogle Scholar
Kulk, G., De Vries, P., Van de Poll, W. et al. (2013). Temperature-dependent photoregulation in oceanic picophytoplankton during excessive irradiance exposure. In Dubinsky, Z. (ed.) Photosynthesis. InTech, London, pp. 209228. www.intechopen.com/books/photosynthesis/temperature-dependent-photoregulation-in-oceanic-picophytoplankton-during-excessive-irradiance-exposGoogle Scholar
Lesser, M. P. & Shick, J. M. (1989). Effects of irradiance and ultraviolet radiation on photoadaptation in the zooxanthellae of Aiptasia pallida: Primary production, photoinhibition, and enzymic defenses against oxygen toxicity. Marine Biology 102: 243255.CrossRefGoogle Scholar
Li, G. & Gao, K. (2013). Cell size-dependent effects of solar UV radiation on primary production in coastal waters of the South China Sea. Estuaries and Coasts 36: 728736.CrossRefGoogle Scholar
Li, G., Wu, Y. & Gao, K. (2009). Effects of Typhoon Kaemi on coastal phytoplankton assemblages in the South China Sea, with special reference to the effects of solar UV radiation. Journal of Geophysical Research 114: 04029. https://doi.org/10.1029/2008JG000896.CrossRefGoogle Scholar
Li, G., Gao, K. & Gao, G. (2011). Differential impacts of solar UV radiation on photosynthetic carbon fixation from the coastal to offshore surface waters in the South China Sea. Photochemistry and Photobiology 87: 329334.CrossRefGoogle ScholarPubMed
Litchman, E., Neale, P. J. & Banaszak, A. T. (2002). Increased sensitivity to ultraviolet radiation in nitrogen-limited dinoflagellates: Photoprotection and repair. Limnology and Oceanography 47: 8694.CrossRefGoogle Scholar
Llabrés, M., Agustí, S., Fernández, M. et al. (2013). Impact of elevated UVB radiation on marine biota: A meta-analysis. Global Ecology and Biogeography 22: 131144.CrossRefGoogle Scholar
Lohr, M. & Wilhelm, C. (2001). Xanthophyll synthesis in diatoms: Quantifications of putative intermediate and comparison of pigment conversion kinetics with rate constants derive from a model. Planta 212: 382391.CrossRefGoogle ScholarPubMed
Malpartida, I., Jerez, C. G., Morales, M. M. et al. (2014). Synergistic effect of UV radiation and nutrient limitation on Chlorella fusca (Chlorophyta) cultures grown in outdoor cylindrical photobioreactor. Aquatic Biology 22: 141158.CrossRefGoogle Scholar
Mengelt, B. & Prézelin, B. B. (2005). UVA enhancement of carbon fixation and resilience to UV inhibition in the genus Pseudo-nitzschia may provide a competitive advantage in high UV surface waters. Marine Ecology Progress Series, 301: 8193.CrossRefGoogle Scholar
Mittler, R. (2002). Oxidative stress, antioxidants and stress tolerance. Trends in Plant Science 7: 405410.CrossRefGoogle ScholarPubMed
Mittler, R., Vanderauwera, S., Gollery, M. et al. (2004). Reactive oxygen gene network of plants. Trends in Plant Science 9: 490498.CrossRefGoogle ScholarPubMed
Mizuta, M. & Yasui, H. (2010). Significance of radical oxygen production in sorus development and zoospore germination in Saccharina japonica (Phaeophyceae). Botanica Marina 53: 409416.CrossRefGoogle Scholar
Müller, R., Desel, C., Steinhoff, F. S. et al. (2012). UV-radiation and elevated temperatures induce formation of reactive oxygen species in gametophytes of cold-temperate/Arctic kelps (Laminariales, Phaeophyceae). Phycological Research 60: 2736.CrossRefGoogle Scholar
Neale, P. J. & Thomas, B. C. (2016). Inhibition by ultraviolet and photosynthetically available radiation lowers model estimates of depth-integrated picophytoplankton photosynthesis: Global predictions for Prochlorococcus and Synechococcus. Global Change Biology 23: 293306.CrossRefGoogle ScholarPubMed
Neale, P. J., Helbling, E. W. & Zagarese, H. E. (2003). Modulation of UVR exposure and effects by vertical mixing and advection. In Helbling, E. W. & Zagarese, H. E. (eds.) UV Effects in Aquatic Organisms and Ecosystems. Royal Society of Chemistry, Cambridge, UK, pp. 108134.Google Scholar
Osburn, C. L., O’Sullivan, D. W. & Boyd, T. J. (2009). Increases in the longwave photobleaching of chromophoric dissolved organic matter in coastal waters. Limnology and Oceanography 54: 145159.CrossRefGoogle Scholar
Pakker, H., Martins, R. S. T., Boelen, P. et al. (2000). Effects of temperature on the photoreactivation of ultraviolet-B–induced DNA damage in Palmaria palmata (Rhodophyta). Journal of Phycology 36: 334341.CrossRefGoogle Scholar
Pescheck, F., Lohbeck, K. T., Roleda, M. Y. et al. (2014). UVB-induced DNA and photosystem II damage in two intertidal green macroalgae: Distinct survival strategies in UV-screening and non-screening Chlorophyta. Journal of Photochemistry and Biology B: Biology 132: 8593.CrossRefGoogle ScholarPubMed
Raven, J. A. (1991). Responses of aquatic photosynthetic organisms to increased solar UVB. Journal of Photochemistry and Biology B: Biology 9: 239244.CrossRefGoogle Scholar
Riebesell, U., Schulz, K. G., Bellerby, R. G. J. et al. (2007). Enhanced biological carbon consumption in a high CO2 ocean. Nature 450: 545548.CrossRefGoogle Scholar
Rodrigues, N. D. N., Staniforth, M. & Stavros, V. G. (2016). Photophysics of sunscreen molecules in the gas phase: A stepwise approach towards understanding and developing next-generation sunscreens. Proceeding of the Royal Society A 472: 20160677. http://dx.doi.org/10.1098/rspa.2016.0677.Google ScholarPubMed
Roleda, M. Y., van de Poll, W. H., Hanelt, D. et al. (2004a). PAR and UVBR effects on photosynthesis, viability, growth and DNA in different life stages of two coexisting Gigartinales: Implications for recruitment and zonation pattern. Marine Ecology Progress Series 281: 3750.CrossRefGoogle Scholar
Roleda, M. Y., Hanelt, D., Kräbs, G. et al. (2004b). Morphology, growth, photosynthesis and pigments in Laminaria ochroleuca (Laminariales, Phaeophyta) under ultraviolet radiation. Phycologia 43: 603613.CrossRefGoogle Scholar
Roleda, M. Y., Wiencke, C., Hanelt, D. et al. (2007). Sensitivity of the early life stages of macroalgae from the northern hemisphere to ultraviolet radiation. Photochemistry and Photobiology 83: 851862.CrossRefGoogle ScholarPubMed
Roleda, M. Y., Mohlin, M., Pattanaik, B. et al. (2008a). Photosynthetic response of Nodularia spumigena to UV and photosynthetically active radiation depends on nutrient (N, P) availability. FEMS Microbial Ecology 66: 230242.CrossRefGoogle ScholarPubMed
Roleda, M. Y., Zacher, K., Wulff, A. et al. (2008b). Susceptibility of spores of different ploidy levels from Antarctic Gigartina skottsbergii (Gigartinales, Rhodophyta) to ultraviolet radiation. Phycologia 47: 361370.CrossRefGoogle Scholar
Roleda, M. Y., Campana, G., Wiencke, C. et al. (2009). Sensitivity of Antarctic Urospora penicilliformis (Ulotrichales, Chlorophyta) to ultraviolet radiation is life stage dependent. Journal of Phycology 45: 600609.CrossRefGoogle ScholarPubMed
Rothschild, L. J. (1999). The influence of UV radiation on protistan evolution. Journal of Eukaryotic Microbiology 46: 548555.CrossRefGoogle ScholarPubMed
Roy, S. (2000). Strategies for the minimization of UV-induced damage. In De Mora, S. J., Demers, S. & Vernet, M. (eds.) The Effects of UV Radiation in the Marine Environment. Cambridge University Press, Cambridge, pp. 177205.CrossRefGoogle Scholar
Rozema, J., van de Staaij, J., Björn, L. O. et al. (1997). UV-B as an environmental factor in plant life: Stress and regulation. Trends in Ecology and Evolution 12: 2228.CrossRefGoogle ScholarPubMed
Safafar, H., van Wagenen, J., Møller, P. et al. (2015). Carotenoids, phenolic compounds and tocopherols contribute to the antioxidative properties of some microalgae species grown on industrial wastewater. Marine Drugs 13: 73397356.CrossRefGoogle Scholar
Sancar, A., Lindsey-Boltz, L. A., Unsal-Kacmaz, K. et al. (2004). Molecular mechanisms of mammalian DNA repair and the DNA damage checkpoints. Annual Review of Biochemistry 73: 3985.CrossRefGoogle ScholarPubMed
Santos, L., Pinto, A., Filipe, O. et al. (2016). Insights on the optical properties of estuarine DOM – Hydrological and biological influences. PLOS ONE 11. https://doi.org/10.1371/journal.pone.0154519.CrossRefGoogle ScholarPubMed
Scheller, H. V. & Haldrup, A. (2005). Photoinhibition of photosystem I. Planta 221: 58.CrossRefGoogle ScholarPubMed
Shigeoka, S., Ishikawa, T., Tamoi, M. et al. (2002). Regulation and function of ascorbate peroxidase isoenzymes. Journal of Experimental Botany 53: 13051319.CrossRefGoogle ScholarPubMed
Sicora, C., Máté, Z. & Vass, I. (2003). The interaction of visible and UV-B light during photodamage and repair of photosystem II. Photosynthesis Research 75: 127137.CrossRefGoogle ScholarPubMed
Smith, R. C., Prézelin, B. B., Baker, K. S. et al. (1992). Ozone depletion: Ultraviolet radiation and phytoplankton biology in Antarctic waters. Science 255: 952959.CrossRefGoogle ScholarPubMed
Smyth, T. J. (2011). Penetration of UV irradiance into the global ocean. Journal of Geophysical Research 116: C11020. https://doi.org/11010.11029/12011JC007183.CrossRefGoogle Scholar
Sobrino, C. & Neale, P. J. (2007). Short-term and long-term effects of temperature on photosynthesis in the diatom Thalassiosira pseudonana under UVR exposures. Journal of Phycology 43: 426436.CrossRefGoogle Scholar
Sommaruga, R. (2001). The role of solar UV radiation in the ecology of alpine lakes. Journal of Photochemistry and Photobiology B: Biology 62: 3542.CrossRefGoogle ScholarPubMed
Stamenković, M. & Hanelt, D. (2013). Protection strategies of Cosmarium strains (Zygnematophyceae, Streptophyta) isolated from various geographic regions against excessive photosynthetically active radiation. Photochemistry and Photobiology 89: 900910.CrossRefGoogle ScholarPubMed
Stamenković, M. & Hanelt, D. (2017). Geographic distribution and ecophysiological adaptations of desmids (Zygnematophyceae, Streptophyta) in relation to PAR, UV radiation and temperature: A review. Hydrobiologia 787: 126.CrossRefGoogle Scholar
Szilárd, A., Sass, L., Deák, Z. et al. (2017). The sensitivity of photosystem II to damage by UV-B radiation depends on the oxidation state of the water-splitting complex. Biochimica et Biophysica Acta 1767: 876882.CrossRefGoogle Scholar
Takaichi, S. (2011). Carotenoids in algae: Distributions, biosyntheses and functions. Marine Drugs 9: 11011118.CrossRefGoogle ScholarPubMed
Thomas, N. V. & Kim, S. K. (2011). Potential pharmacological applications of polyphenolic derivatives from marine brown algae. Environmental Toxicology and Pharmacology 32: 325335.CrossRefGoogle ScholarPubMed
UNEP (2017). Environmental effects of ozone depletion and its interactions with climate change: Progress report 2016. Photochemical and Photobiological Sciences 16: 107145.CrossRefGoogle Scholar
van de Poll, W., Visser, R. J. & Buma, A. G. J. (2007). Acclimation to a dynamic irradiance regime changes excessive irradiance sensitivity of Emiliania huxleyi and Thalassiosira weissflogii. Limnology and Oceanography 52: 14301438.CrossRefGoogle Scholar
Van de Poll, W. H. & Buma, A. G. J. (2009). Does ultraviolet radiation affect the xanthophyll cycle in marine phytoplankton? Photochemical and Photobiological Sciences 8: 12951301.CrossRefGoogle ScholarPubMed
van de Poll, W. H., Eggert, A., Buma, A. G. J. et al. (2001). Effects of UV-B induced DNA damage and photoinhibition on growth of temperate marine red macrophytes: Habitat related differences in ultraviolet-B tolerance. Journal of Phycology 37: 3037.CrossRefGoogle Scholar
Van de Poll, W. H., Eggert, A., Buma, A. G. J. et al. (2002a). Temperature dependence of UV radiation effects in Arctic and temperate isolates of three red macrophytes. European Journal of Phycology 37: 5968.CrossRefGoogle Scholar
van de Poll, W. H., van Leeuwe, M. A., Roggeveld, J. et al. (2005). Nutrient limitation and high irradiance acclimation reduce PAR and UV-induced viability loss in the Antarctic diatom Chaetoceros brevis (Bacillariophyceae). Journal of Phycology 41: 840850.CrossRefGoogle Scholar
van de Poll, W. H., Hanelt, D., Hoyer, K. et al. (2002b). Ultraviolet-B-induced cyclobutane-pyrimidine dimer formation and repair in Arctic marine macrophytes. Photochemistry and Photobiology 76: 493500.2.0.CO;2>CrossRefGoogle ScholarPubMed
van de Poll, W. H., Alderkamp, A.-C., Janknegt, P. J. et al. (2006). Photoacclimation modulates excessive photosynthetically active and ultraviolet radiation effects in a temperate and an Antarctic marine diatom. Limnology and Oceanography 51: 12391248.CrossRefGoogle Scholar
Vasilkov, A. P., Krotkov, N., Haffner, D., Fasnacht, Z., & Joiner, J. (2022) Estimates of hyperspectral surface and underwater UV planar and scalar irradiances from OMI measurements and radiative transfer computations. Remote Sensing 14: 2278. https://doi.org/10.3390/rs14092278CrossRefGoogle Scholar
Villafañe, V. E., Sundbäck, K., Figueroa, F. L. et al. (2003). Photosynthesis in the aquatic environment as affected by UVR. In Helbling, E. W. & Zagarese, H. E. (eds.) UV Effects in Aquatic Organisms and Ecosystems. Royal Society of Chemistry, Cambridge, UK, pp. 357397.Google Scholar
Villafañe, V. E., Buma, A. G. J., Boelen, P. et al. (2004). Solar UVR-induced DNA damage and inhibition of photosynthesis in phytoplankton from Andean lakes of Argentina. Archiv für Hydrobiologie 161: 245266.CrossRefGoogle Scholar
Villafañe, V. E., Valiñas, M. S., Cabrerizo, M. J. et al. (2015). Physio-ecological responses of Patagonian coastal marine phytoplankton in a scenario of global change: Role of acidification, nutrients and solar UVR. Marine Chemistry 177: 411420.CrossRefGoogle Scholar
Villafañe, V. E., Cabrerizo, M. J., Erzinger, G. S. et al. (2017). Photosynthesis and growth of temperate and sub-tropical estuarine phytoplankton in a scenario of nutrient enrichment under solar ultraviolet radiation exposure. Estuaries and Coasts 40: 842855.CrossRefGoogle Scholar
Vlcček, D., Ševčovičová, A., Sviežená, B., Gálová, E. et al. (2008). Chlamydomonas reinhardtii: A convenient model system for the study of DNA repair in photoautotrophic eukaryotes. Current Genetics 53: 122.CrossRefGoogle Scholar
Wada, N., Sakamoto, T. & Matsugo, S. (2015). Mycosporine-like amino acids and their derivatives as natural antioxidants. Antioxidants 4: 603646.CrossRefGoogle ScholarPubMed
Williamson, C. E. & Rose, K. C. (2010). When UV meets freshwater. Science 329: 637639.CrossRefGoogle Scholar
Williamson, C. E., Zepp, R. G., Lucas, R. M. et al. (2014). Solar ultraviolet radiation in a changing climate. Nature Climate Change 4: 434441.CrossRefGoogle Scholar
Wu, H., Abasova, L., Cheregi, O. et al. (2011). D1 protein turnover is involved in protection of Photosystem II against UV-B induced damage in the cyanobacterium Arthrospira (Spirulina) platensis. Journal of Photochemistry and Biology B: Biology 104: 320325.CrossRefGoogle ScholarPubMed
Wu, Y., Gao, K., Li, G. et al. (2010). Seasonal impacts of solar UV radiation on photosynthesis of phytoplankton assemblages in the coastal waters of the South China Sea. Photochemistry and Photobiology 86: 586592.CrossRefGoogle ScholarPubMed
Wulff, A., Wängberg, S.-Å., Sundbäck, K. et al. (2000). Effects of UVB radiation on a marine microphytobenthic community growing on a sand-substratum under different nutrient conditions. Limnology and Oceanography 45: 11441152.CrossRefGoogle Scholar
Xenopoulos, M. A., Frost, P. C. et al. (2002). Joint effects of UV radiation and phosphorus supply on algal growth rate and elemental composition. Ecology 83: 423435.CrossRefGoogle Scholar
Zheng, Y. & Gao, K. (2009). Impacts of solar UV radiation on the photosynthesis, growth, and UV-absorbing compounds in Gracilaria lemaneiformis (Rhodophyta) grown at different nitrate concentrations. Journal of Phycology 45: 314323.CrossRefGoogle ScholarPubMed

References

Aaronson, S. (1974). The biology and ultrastructure of phagotrophy in Ochromonas danica (Chrysophyceae: Chrysomonadida). Journal of General Microbiology 83: 2129.CrossRefGoogle Scholar
Albert, J. S., Destouni, G., Duke-Sylvester, S. M. et al. (2021). Scientists’ warning to humanity on the freshwater biodiversity crisis. Ambio 50: 8594.CrossRefGoogle ScholarPubMed
Anderson, D. M., Cembella, A. D. & Hallegraeff, G. M. (2012). Progress in understanding harmful algal blooms: Paradigm shifts and new technologies for research, monitoring, and management. Annual Review of Marine Science 4: 143176.CrossRefGoogle ScholarPubMed
Anderson, D. M., Glibert, P. M. & Burkholder, J. M. (2002). Harmful algal blooms and eutrophication: Nutrient sources, composition, and consequences. Estuaries 25: 704726.CrossRefGoogle Scholar
Andresen, E., Peiter, E. & Küpper, H. (2018). Trace metal metabolism in plants. Journal of Experimental Botany 69: 909954.CrossRefGoogle ScholarPubMed
Attayde, J. L. & Hansson, L. A. (1999). Effects of nutrient recycling by zooplankton and fish on phytoplankton communities. Oecologia 121: 4754.CrossRefGoogle ScholarPubMed
Azam, F., Fenchel, T., Field, J. G. et al. (1983). The ecological role of water-column microbes in the sea. Marine Ecology Progress Series 10: 257263.CrossRefGoogle Scholar
Barko, J. W., Gunnison, D. & Carpenter, S. R. (1991). Sediment interactions with submersed macrophyte growth and community dynamics. Aquatic Botany 41: 4165.CrossRefGoogle Scholar
Belzile, C., Vincent, W. F., Gibson, J. A. et al. (2001). Bio-optical characteristics of the snow, ice, and water column of a perennially ice-covered lake in the High Arctic. Canadian Journal of Fisheries and Aquatic Sciences 58: 24052418.CrossRefGoogle Scholar
Bergstrom, A. K. & Jansson, M. (2006). Atmospheric nitrogen deposition has caused nitrogen enrichment and eutrophication of lakes in the northern hemisphere. Global Change Biology 12: 635643.CrossRefGoogle Scholar
Beusen, A. H. W. & Bouwman, A. F. (2022). Future projections of river nutrient export to the global coastal ocean show persisting nitrogen and phosphorus distortion. Frontiers in Water 4: 893585.CrossRefGoogle Scholar
Beusen, A. H. W., Bouwman, A. F., Van Beek, L. P. H. et al. (2016). Global riverine N and P transport to ocean increased during the 20th century despite increased retention along the aquatic continuum. Biogeosciences 13: 24412451.CrossRefGoogle Scholar
Beusen, A. H. W., Doelman, J. C., Van Beek, L. P. H. et al. (2022). Exploring river nitrogen and phosphorus loading and export to global coastal waters in the shared socio-economic pathways. Global Environmental Change 72: 102426.CrossRefGoogle Scholar
Boyd, P. W., Jickells, T., Law, C. S. et al. (2007). Mesoscale iron enrichment experiments 1993–2005: Synthesis and future directions. Science 315: 612617.CrossRefGoogle ScholarPubMed
Brown, C. J., Saunders, M. I., Possingham, H. P. et al. (2014). Interactions between global and local stressors of ecosystems determine management effectiveness in cumulative impact mapping. Diversity and Distributions 20: 538546.CrossRefGoogle Scholar
Brownlie, W., Sutton, M. A., Heal, K. V. et al. (2022). The Our Phosphorus Future Report. https://doi.org/10.13140/RG.2.2.17834.08645.CrossRefGoogle Scholar
Brzezinski, M. A. (1985). The Si:C:N ratio of marine diatoms: Interspecific variability and the effect of some environmental variables. Journal of Phycology 21: 347357.CrossRefGoogle Scholar
Burger, D. F., Hamilton, D. P., Pilditch, C. A. et al. (2007). Benthic nutrient fluxes in a eutrophic, polymictic lake. Hydrobiologia 584: 1325.CrossRefGoogle Scholar
Burkholder, J. M., Tomasko, D. A. & Touchette, B. W. (2007). Seagrasses and eutrophication. Journal of Experimental Marine Biology and Ecology 350: 4672.CrossRefGoogle Scholar
Carey, C. (2023). Causes and consequences of changing oxygen availability in lakes Inland Waters 13: 316–326.CrossRefGoogle Scholar
Carey, C. C., Hanson, P. C., Thomas, R. Q. et al. (2022). Anoxia decreases the magnitude of the carbon, nitrogen, and phosphorus sink in freshwaters. Global Change Biology 28: 48614881.CrossRefGoogle ScholarPubMed
Carpenter, S. R., Caraco, N. F., Correll, D. L. et al. (1998). Nonpoint pollution of surface waters with phosphorus and nitrogen. Ecological Applications 8: 559568.CrossRefGoogle Scholar
Carpenter, S. R., Stanley, E. H. & Vander Zanden, M. J. (2011). State of the world’s freshwater ecosystems: Physical, chemical, and biological changes. Annual Review of Environment and Resources 36: 7599.CrossRefGoogle Scholar
Chambers, P. A., Prepas, E. E., Bothwell, M. L. et al. (1989). Roots versus shoots in nutrient uptake by aquatic macrophytes in flowing waters. Canadian Journal of Fisheries and Aquatic Sciences 46: 435439.CrossRefGoogle Scholar
Charette, M. A., Gonneea, M. E., Morris, P. J. et al. (2007). Radium isotopes as tracers of iron sources fueling a Southern Ocean phytoplankton bloom. Deep Sea Research Part II: Topical Studies in Oceanography 54: 19891998.CrossRefGoogle Scholar
Cloern, J. (2001). Our evolving conceptual model of the coastal eutrophication problem. Marine Ecology Progress Series 210: 223253.CrossRefGoogle Scholar
Coale, K. H., Johnson, K. S., Fitzwater, S. E. et al. (1996). A massive phytoplankton bloom induced by an ecosystem-scale iron fertilization experiment in the equatorial Pacific Ocean. Nature 383: 495501.CrossRefGoogle ScholarPubMed
Conley, D. J., Kilham, S. S. & Theriot, E. (1989). Differences in silica content between marine and fresh-water diatoms. Limnology and Oceanography 34: 205213.CrossRefGoogle Scholar
Cordell, D. & White, S. (2011). Peak phosphorus: Clarifying the key issues of a vigorous debate about long-term phosphorus security. Sustainability 3: 20272049.CrossRefGoogle Scholar
Currie, H. A. & Perry, C. C. (2007). Silica in plants: Biological, biochemical and chemical studies. Annals of Botany 100: 13831389.CrossRefGoogle ScholarPubMed
Dai, Y., Yang, S., Zhao, D. et al. (2023). Coastal phytoplankton blooms expand and intensify in the 21st century. Nature 615: 280284.CrossRefGoogle ScholarPubMed
de-Bashan, L. E. & Bashan, Y. (2004). Recent advances in removing phosphorus from wastewater and its future use as fertilizer (1997–2003). Water Research 38: 42224246.CrossRefGoogle ScholarPubMed
De’ath, G. & Fabricius, K. (2010). Water quality as a regional driver of coral biodiversity and macroalgae on the Great Barrier Reef. Ecological Applications 20: 840850.CrossRefGoogle ScholarPubMed
Diaz, R. J. & Rosenberg, R. (2008). Spreading dead zones and consequences for marine ecosystems. Science 321: 926929.CrossRefGoogle ScholarPubMed
Dillon, P. J. & Rigler, F. H. (1974). Phosphorus-chlorophyll relationship in lakes. Limnology and Oceanography 19: 767773.CrossRefGoogle Scholar
Dong, X. (2010). Using diatoms to understand climate-nutrient interactions in Esthwaite Water, England: evidence from observational and palaeolimnological records, University College London Retrieved from https://discovery.ucl.ac.uk/id/eprint/763095/1/763095.pdfGoogle Scholar
Dong, X., Bennion, H., Battarbee, R. W. et al. (2012a). A multiproxy palaeolimnological study of climate and nutrient impacts on Esthwaite Water, England over the past 1200 years. The Holocene 22: 107118.CrossRefGoogle Scholar
Dong, X., Bennion, H., Maberly, S. C. et al. (2012b). Nutrients exert a stronger control than climate on recent diatom communities in Esthwaite Water: Evidence from monitoring and palaeolimnological records. Freshwater Biology 57: 20442056.CrossRefGoogle Scholar
Drake, J. C. & Heaney, S. I. (1987). Occurrence of phosphorus and its potential remobilization in the littoral sediments of a productive English lake. Freshwater Biology 17: 513523.CrossRefGoogle Scholar
Elliott, J. A. (2010). The seasonal sensitivity of cyanobacteria and other phytoplankton to changes in flushing rate and water temperature. Global Change Biology 16: 864876.CrossRefGoogle Scholar
Elser, J. J., Andersen, T., Baron, J. S., et al. (2009a). Shifts in lake N:P stoichiometry and nutrient limitation driven by atmospheric nitrogen deposition. Science 326: 835837.CrossRefGoogle Scholar
Elser, J. J., Bracken, M. E. S., Cleland, E. E. et al. (2007). Global analysis of nitrogen and phosphorus limitation of primary producers in freshwater, marine and terrestrial ecosystems. Ecology Letters 10: 11351142.CrossRefGoogle ScholarPubMed
Elser, J. J., Kyle, M., Steger, L. et al. (2009b). Nutrient availability and phytoplankton nutrient limitation across a gradient of atmospheric nitrogen deposition. Ecology 90: 30623073.CrossRefGoogle ScholarPubMed
Eom, H., Borgatti, D., Paerl, H. W. et al. (2017). Formation of low-molecular-weight dissolved organic nitrogen in predenitrification biological nutrient removal systems and its impact on eutrophication in coastal waters. Environmental Science and Technology 51: 37763783.CrossRefGoogle ScholarPubMed
Erisman, J. W., Sutton, M. A., Galloway, J. et al. (2008). How a century of ammonia synthesis changed the world. Nature Geoscience 1: 636639.CrossRefGoogle Scholar
Evans, C. D., Monteith, D. T. & Harriman, R. (2001). Long-term variability in the deposition of marine ions at west coast sites in the UK Acid Waters Monitoring Network: Impacts on surface water chemistry and significance for trend determination. Science of the Total Environment 265: 115129.CrossRefGoogle ScholarPubMed
Fabricius, K. E. (2005). Effects of terrestrial runoff on the ecology of corals and coral reefs: review and synthesis. Marine Pollution Bulletin 50: 125146.CrossRefGoogle ScholarPubMed
Fee, E. J., Hecky, R. E., Regehr, G. W. et al. (1994). Effects of lake size on nutrient availability in the mixed layer during summer stratification. Canadian Journal of Fisheries and Aquatic Sciences 51: 27562768.CrossRefGoogle Scholar
Ferber, L. R., Levine, S. N., Lini, A. et al. (2004). Do cyanobacteria dominate in eutrophic lakes because they fix atmospheric nitrogen? Freshwater Biology 49: 690708.CrossRefGoogle Scholar
Filstrup, C. T. & Downing, J. A. (2017). Relationship of chlorophyll to phosphorus and nitrogen in nutrient-rich lakes. Inland Waters 7: 385400.CrossRefGoogle Scholar
Floener, L. & Bothe, H. (1980). Nitrogen fixation in Rhopalodia gibba, a diatom containing blue-greenish inclusions symbiotically. In: Schenk, H. E. A. & Schwemmler, W. (eds.) Endosymbiosis and Cell Biology, De Gruyter, Berlin, pp. 541552.CrossRefGoogle Scholar
Foley, B., Jones, I. D., Maberly, S. C. et al. (2012). Long-term changes in oxygen depletion in a small temperate lake: Effects of climate change and eutrophication. Freshwater Biology 57: 278289.CrossRefGoogle Scholar
Francoeur, S. N., Smith, R. A. & Lowe, R. L. (1999). Nutrient limitation of algal biomass accrual in streams: Seasonal patterns and a comparison of methods. Journal of the North American Benthological Society 18: 242260.CrossRefGoogle Scholar
Gächter, R. & Wehrli, B. (1998). Ten years of artificial mixing and oxygenation: No effect on the internal phosphorus loading of two eutrophic lakes. Environmental Science and Technology 32: 36593665.CrossRefGoogle Scholar
Galloway, J. N., Dentener, F. J., Capone, D. G. et al. (2004). Nitrogen cycles: Past, present, and future. Biogeochemistry 70: 153226.CrossRefGoogle Scholar
Giordano, M., Norici, A. & Hell, R. (2005). Sulfur and phytoplankton: Acquisition, metabolism and impact on the environment. New Phytologist 166: 371382.CrossRefGoogle ScholarPubMed
Glass, J. B., Axler, R. P., Chandra, S. et al. (2012). Molybdenum limitation of microbial nitrogen assimilation in aquatic ecosystems and pure cultures. Frontiers in Microbiology 3: Article 331.CrossRefGoogle ScholarPubMed
Godfray, H. C. J., Beddington, J. R., Crute, I. R. et al. (2010). Food security: The challenge of feeding 9 billion people. Science 327: 812818.CrossRefGoogle ScholarPubMed
Guilizzoni, P. (1991). The role of heavy metals and toxic amterials in the physiological ecology of submersed macrophytes. Aquatic Botany 41: 87109.CrossRefGoogle Scholar
Hamilton, D. P., Salmaso, N. & Paerl, H. W. (2016). Mitigating harmful cyanobacterial blooms: Strategies for control of nitrogen and phosphorus loads. Aquatic Ecology 50: 351366.CrossRefGoogle Scholar
Hampton, S. E., Izmest’Eva, L. R., Moore, M. V. et al. (2008). Sixty years of environmental change in the world’s largest freshwater lake – Lake Baikal, Siberia: Warming of the world’s largest freshwater lake. Global Change Biology 14: 19471958.CrossRefGoogle Scholar
Hanson, J. M. & Leggett, W. C. (1982). Empirical prediction of fish biomass and yield. Canadian Journal of Fisheries and Aquatic Sciences 39: 257263.CrossRefGoogle Scholar
Harke, M. J., Steffen, M. M., Gobler, C. J. et al. (2016). A review of the global ecology, genomics, and biogeography of the toxic cyanobacterium, Microcystis spp. Harmful Algae 54: 420.CrossRefGoogle ScholarPubMed
Hein, L. & Leemans, R. (2012). The impact of first-generation biofuels on the depletion of the global phosphorus reserve. AMBIO 41: 341349.CrossRefGoogle ScholarPubMed
Holm-Nielsen, J. B., Al Seadi, T. & Oleskowicz-Popiel, P. (2009). The future of anaerobic digestion and biogas utilization. Bioresource Technology 100: 54785484.CrossRefGoogle ScholarPubMed
Hu, Z., Anderson, N. J., Yang, X. et al. (2014). Catchment-mediated atmospheric nitrogen deposition drives ecological change in two alpine lakes in SE Tibet. Global Change Biology 20: 16141628.CrossRefGoogle Scholar
Huisman, J., Codd, G. A., Paerl, H. W. et al. (2018). Cyanobacterial blooms. Nature Reviews Microbiology 16: 471483.CrossRefGoogle ScholarPubMed
Ishangulyyev, R., Kim, S. & Lee, S. (2019). Understanding food loss and waste – why are we losing and wasting food? Foods 8: 297.CrossRefGoogle ScholarPubMed
James, C., Fisher, J. & Moss, B. (2003). Nitrogen driven lakes: The Shropshire and Cheshire meres? Archiv Fur Hydrobiologie 158: 249266.CrossRefGoogle Scholar
Jane, S. F., Hansen, G. J. A., Kraemer, B. M. et al. (2021). Widespread deoxygenation of temperate lakes. Nature 594: 6670.CrossRefGoogle ScholarPubMed
Jankowski, T., Livingstone, D. M., Bührer, H. et al. (2006). Consequences of the 2003 European heat wave for lake temperature profiles, thermal stability, and hypolimnetic oxygen depletion: Implications for a warmer world. Limnology and Oceanography 51: 815819.CrossRefGoogle Scholar
Janse, J. H., Kuiper, J. J., Weijters, M. J. et al. (2015). GLOBIO-Aquatic, a global model of human impact on the biodiversity of inland aquatic ecosystems. Environmental Science and Policy 48: 99114.CrossRefGoogle Scholar
Jansson, M., Berggren, M., Laudon, H., Jonsson, A., 2012. Bioavailable phosphorus in humic headwater streams in boreal Sweden. Limnology & Oceanography 57: 11611170.CrossRefGoogle Scholar
Jarvie, H. P., Sharpley, A. N., Flaten, D. et al. (2015). The pivotal role of phosphorus in a resilient water-energy-food security nexus. Journal of Environmental Quality 44: 10491062.CrossRefGoogle Scholar
Jarvie, H. P., Sharpley, A. N., Spears, B. et al. (2013). Water quality remediation faces unprecedented challenges from ‘Legacy Phosphorus’. Environmental Science and Technology 47: 89978998.CrossRefGoogle ScholarPubMed
Jensen, E. L., Clement, R., Kosta, A. et al. (2019). A new widespread subclass of carbonic anhydrase in marine phytoplankton. The ISME Journal 13: 20942106.CrossRefGoogle ScholarPubMed
Jeppesen, E., Peder Jensen, J., Søndergaard, M. et al. (2000). Trophic structure, species richness and biodiversity in Danish lakes: Changes along a phosphorus gradient: A detailed study of Danish lakes along a phosphorus gradient. Freshwater Biology 45: 201218.CrossRefGoogle Scholar
Jeppesen, E., Pekcan-Hekim, Z., Lauridsen, T. L. et al. (2006). Habitat distribution of fish in late summer: Changes along a nutrient gradient in Danish lakes. Ecology of Freshwater Fish 15: 180190.CrossRefGoogle Scholar
Jiang, S., Hua, H., Sheng, H. et al. (2019). Phosphorus footprint in China over the 1961–2050 period: Historical perspective and future prospect. Science of the Total Environment 650: 687695.CrossRefGoogle Scholar
Jing, X., Lin, S., Zhang, H. et al. (2017). Utilization of urea and expression profiles of related genes in the dinoflagellate Prorocentrum donghaiense. PLOS ONE 12: e0187837.CrossRefGoogle ScholarPubMed
Jones, I., George, G. & Reynolds, C. (2005). Quantifying effects of phytoplankton on the heat budgets of two large limnetic enclosures. Freshwater Biology 50: 12391247.CrossRefGoogle Scholar
Jones, R. I., Young, J. M., Hartley, A. M. et al. (1996). Light limitation of phytoplankton development in an oligotrophic lake – Loch Ness, Scotland. Freshwater Biology 35: 533543.CrossRefGoogle Scholar
Ju, X.-T., Xing, G.-X., Chen, X.-P. et al. (2009). Reducing environmental risk by improving N management in intensive Chinese agricultural systems. Proceedings of the National Academy of Sciences 106: 30413046.CrossRefGoogle ScholarPubMed
Karl, D., Michaels, A., Bergman, B. et al. (2002). Dinitrogen fixation in the world’s oceans. Biogeochemistry 57: 4798.CrossRefGoogle Scholar
Karlsson, J., Bystrom, P., Ask, J. et al. (2009). Light limitation of nutrient-poor lake ecosystems. Nature 460: 506509.CrossRefGoogle ScholarPubMed
Kellogg, R. M., Moosburner, M. A., Cohen, N. R. et al. (2022). Adaptive responses of marine diatoms to zinc scarcity and ecological implications. Nature Communications 13: 1995.CrossRefGoogle ScholarPubMed
Kibriya, S. & Jones, J. I. (2007). Nutrient availability and the carnivorous habit in Utricularia vulgaris. Freshwater Biology 52: 500509.CrossRefGoogle Scholar
Kim, H., Takayama, K., Hirose, N. et al. (2019). Biological modulation in the seasonal variation of dissolved oxygen concentration in the upper Japan Sea. Journal of Oceanography 75: 257271.CrossRefGoogle Scholar
Kirk, J. T. O. (2010). Light and Photosynthesis in Aquatic Environments, 3rd ed. Cambridge University Press, Cambridge, UK.CrossRefGoogle Scholar
Kleeberg, A. & Kohl, J.-G. (1999). Assessment of the long-term effectiveness of sediment dredging to reduce benthic phosphorus release in shallow Lake Müggelsee (Germany). Hydrobiologia 394: 153161.CrossRefGoogle Scholar
Krupke, A., Lavik, G., Halm, H. et al. (2014). Distribution of a consortium between unicellular algae and the N2 fixing cyanobacterium UCYN-A in the North Atlantic Ocean: Distribution of a consortium between UCYN-A and Haptophyta. Environmental Microbiology 16: 31533167.CrossRefGoogle Scholar
Lapointe, B. E., West, L. E., Sutton, T. T. et al. (2014). Ryther revisited: Nutrient excretions by fishes enhance productivity of pelagic Sargassum in the western North Atlantic Ocean. Journal of Experimental Marine Biology and Ecology 458: 4656.CrossRefGoogle Scholar
Larned, S. T. (1998). Nitrogen- versus phosphorus-limited growth and sources of nutrients for coral reef macroalgae. Marine Biology 132: 409421.CrossRefGoogle Scholar
Lassaletta, L., Billen, G., Garnier, J. et al. (2016). Nitrogen use in the global food system: Past trends and future trajectories of agronomic performance, pollution, trade, and dietary demand. Environmental Research Letters 11: 095007.CrossRefGoogle Scholar
Le Corre, K. S., Valsami-Jones, E., Hobbs, P. et al. (2009). Phosphorus recovery from wastewater by struvite crystallization: A review. Critical Reviews in Environmental Science and Technology 39: 433477.CrossRefGoogle Scholar
Lee, K.-S., Park, S. R. & Kim, Y. K. (2007). Effects of irradiance, temperature, and nutrients on growth dynamics of seagrasses: A review. Journal of Experimental Marine Biology and Ecology 350: 144175.CrossRefGoogle Scholar
Leles, S. G., Mitra, A., Flynn, K. J. et al. (2017). Oceanic protists with different forms of acquired phototrophy display contrasting biogeographies and abundance. Proceedings of the Royal Society B: Biological Sciences 284: 20170664.CrossRefGoogle ScholarPubMed
Lund, J. W. G. (1950). Studies on Asterionella formosa Hass. II. Nutrient depletion and the spring maximum. Journal of Ecology 38: 114.CrossRefGoogle Scholar
Lürling, M. (2021). Grazing resistance in phytoplankton. Hydrobiologia 848: 237249.CrossRefGoogle Scholar
Lürling, M. & Mucci, M. (2020). Mitigating eutrophication nuisance: In-lake measures are becoming inevitable in eutrophic waters in the Netherlands. Hydrobiologia 847: 44474467.CrossRefGoogle Scholar
Maavara, T., Parsons, C. T., Ridenour, C. et al. (2015). Global phosphorus retention by river damming. Proceedings of the National Academy of Sciences USA 112: 1560315608.CrossRefGoogle ScholarPubMed
Maberly, S. C., Barker, P. A., Stott, A. W. et al. (2013). Catchment productivity controls CO2 emissions from lakes. Nature Climate Change 3: 391394.CrossRefGoogle Scholar
Maberly, S. C., De Ville, M. M., Thackeray, S. J. et al. (2016). A survey of the status of the lakes of the English Lake District: The 2015 Lakes Tour. Report to United Utilities No. LA/NEC05369/1.Google Scholar
Maberly, S. C., King, L., Dent, M. M. et al. (2002). Nutrient limitation of phytoplankton and periphyton growth in upland lakes. Freshwater Biology 47: 21362152.CrossRefGoogle Scholar
Maberly, S. C. & Madsen, T. V. (1990). Contribution of air and water to the carbon balance of Fucus spiralis. Marine Ecology Progress Series 62: 175183.CrossRefGoogle Scholar
Maberly, S. C., Pitt, J.-A., Davies, P. S. et al. (2020). Nitrogen and phosphorus limitation and the management of small productive lakes. Inland Waters 10: 159172.CrossRefGoogle Scholar
Maberly, S. C., Van de Waal, D. B. & Raven, J. A. (2022). Phytoplankton growth and nutrients. In Encyclopedia of Inland Waters, 2nd ed. Elsevier, Amsterdam, pp. 130138.CrossRefGoogle Scholar
Machovina, B., Feeley, K. J. & Ripple, W. J. (2015). Biodiversity conservation: The key is reducing meat consumption. Science of the Total Environment 536: 419431.CrossRefGoogle ScholarPubMed
Mackay, E. B., Feuchtmayr, H., De Ville, M. M. et al. (2020). Dissolved organic nutrient uptake by riverine phytoplankton varies along a gradient of nutrient enrichment. Science of the Total Environment 722: 137837.CrossRefGoogle ScholarPubMed
Madsen, T. V. & Cedergreen, N. (2002). Sources of nutrients to rooted submerged macrophytes growing in a nutrient-rich stream. Freshwater Biology 47: 283291.CrossRefGoogle Scholar
Malone, T. C. & Newton, A. (2020). The globalization of cultural eutrophication in the coastal ocean: Causes and consequences. Frontiers in Marine Science 7: 670.CrossRefGoogle Scholar
Martin, J. H., Gordon, M. & Fitzwater, S. E. (1991). The case for iron. Limnology and Oceanography 36: 17931802.CrossRefGoogle Scholar
Martin, J. H. & Gordon, M. R. (1988). Northeast Pacific iron distributions in relation to phytoplankton productivity. Deep Sea Research Part A. Oceanographic Research Papers 35: 177196.CrossRefGoogle Scholar
Martiny, A. C., Pham, C. T. A., Primeau, F. W. et al. (2013). Strong latitudinal patterns in the elemental ratios of marine plankton and organic matter. Nature Geoscience 6: 279283.CrossRefGoogle Scholar
Matthijs, H. C. P., Visser, P. M., Reeze, B. et al. (2012). Selective suppression of harmful cyanobacteria in an entire lake with hydrogen peroxide. Water Research 46: 14601472.CrossRefGoogle Scholar
McCrackin, M. L., Jones, H. P., Jones, P. C. et al. (2017). Recovery of lakes and coastal marine ecosystems from eutrophication: A global meta-analysis. Limnology and Oceanography 62: 507518.CrossRefGoogle Scholar
McGowan, S., Barker, P., Haworth, E. Y. et al. (2012). Humans and climate as drivers of algal community change in Windermere since 1850. Freshwater Biology 57: 260277.Google Scholar
Meis, S., Spears, B. M., Maberly, S. C. et al. (2012). Sediment amendment with Phoslock (R) in Clatto Reservoir (Dundee, UK): Investigating changes in sediment elemental composition and phosphorus fractionation. Journal of Environmental Management 93: 185193.CrossRefGoogle Scholar
Metson, G. S., Brownlie, W. J. & Spears, B. M. (2022). Towards net-zero phosphorus cities. Npj Urban Sustainability 2: 30. https://doi.org/10.1038/s42949–022–00076–8.CrossRefGoogle Scholar
Metson, G. S., Cordell, D. & Ridoutt, B. (2016). Potential impact of dietary choices on phosphorus recycling and global phosphorus footprints: The case of the average Australian city. Frontiers in Nutrition 3: 35. https://doi.org/10.3389/fnut.2016.00035.CrossRefGoogle ScholarPubMed
Michelou, V. K., Lomas, M. W. & Kirchman, D. L. (2011). Phosphate and adenosine-5’-triphosphate uptake by cyanobacteria and heterotrophic bacteria in the Sargasso Sea. Limnology and Oceanography 56: 323332.CrossRefGoogle Scholar
Millennium Ecosystem Assessment (2005). Ecosystems and Human Well-being: Synthesis; A Report of the Millennium Ecosystem Assessment. Island Press, Washington, DC.Google Scholar
Mogollón, J. M., Bouwman, A. F., Beusen, A. H. W. et al. (2021). More efficient phosphorus use can avoid cropland expansion. Nature Food 2: 509518.CrossRefGoogle ScholarPubMed
Molinari, B., Stewart-Koster, B., Adame, M. F. et al. (2021). Relationships between algal primary productivity and environmental variables in tropical floodplain wetlands. Inland Waters 11: 180190.CrossRefGoogle Scholar
Moore, C. M., Mills, M. M., Arrigo, K. R. et al. (2013). Processes and patterns of oceanic nutrient limitation. Nature Geoscience 6: 701710.CrossRefGoogle Scholar
Morel, F. M. M., Lam, P. J. & Saito, M. A. (2020). Trace metal substitution in marine phytoplankton. Annual Review of Earth and Planetary Sciences 48: 491517.CrossRefGoogle Scholar
Mortimer, C. (1941). The exchange of dissolved substances between mud and water in lakes I and II. Journal of Ecology 29: 280329.CrossRefGoogle Scholar
Mosley, L. M. (2015). Drought impacts on the water quality of freshwater systems; review and integration. Earth-Science Reviews 140: 203214.CrossRefGoogle Scholar
Myrstener, M., Fork, M. L., Bergström, A.-K. et al. (2022). Resolving the drivers of algal nutrient limitation from boreal to arctic lakes and streams. Ecosystems 25: 16821699.CrossRefGoogle Scholar
Njagi, D. M., Routh, J., Odhiambo, M. et al. (2022). A century of human-induced environmental changes and the combined roles of nutrients and land use in Lake Victoria catchment on eutrophication. Science of the Total Environment 835: 155425.CrossRefGoogle Scholar
Norici, A., Gerotto, C., Beardall, J. et al. (2022). Environmental variability and its control of productivity. In: Maberly, S. C. & Gontero, B. (eds.) Blue Planet, Red and Green Photosynthesis: Productivity and Carbon Cycling in Aquatic Ecosystems ISTE-WILEY, London, pp. 225271.CrossRefGoogle Scholar
O’Hare, M. T., Stillman, R. A., McDonnell, J. et al. (2007). Effects of mute swan grazing on a keystone macrophyte. Freshwater Biology 52: 24632475.CrossRefGoogle Scholar
Paerl, H. W. (1997). Coastal eutrophication and harmful algal blooms: Importance of atmospheric deposition and groundwater as ‘new’ nitrogen and other nutrient sources. Limnology and Oceanography 42: 11541165.CrossRefGoogle Scholar
Paerl, H. W., Gardner, W. S., Havens, K. E. et al. (2016a). Mitigating cyanobacterial harmful algal blooms in aquatic ecosystems impacted by climate change and anthropogenic nutrients. Harmful Algae 54: 213222.CrossRefGoogle ScholarPubMed
Paerl, H. W., Scott, J. T., McCarthy, M. J. et al. (2016b). It takes two to tango: When and where dual nutrient (N & P) reductions are needed to protect lakes and downstream ecosystems. Environmental Science and Technology 50: 1080510813.CrossRefGoogle ScholarPubMed
Pearsall, W. H. (1921). The development of vegetation in the English lakes, considered in relation to the general evolution of glacial lakes and rock basins. Proceedings of the Royal Society of London. Series B 92: 259284.Google Scholar
Pedrozo, F., Kelly, L., Diaz, M. et al. (2001). First results on the water chemistry, algae and trophic status of an Andean acidic lake system of volcanic origin in Patagonia (Lake Caviahue). Hydrobiologia 452: 129137.CrossRefGoogle Scholar
Peierls, B. L. & Paerl, H. W. (1997). Bioavailability of atmospheric organic nitrogen deposition to coastal phytoplankton. Limnology and Oceanography 42: 18191823.CrossRefGoogle Scholar
Peñuelas, J., Poulter, B., Sardans, J. et al. (2013). Human-induced nitrogen–phosphorus imbalances alter natural and managed ecosystems across the globe. Nature Communications 4: 2934.CrossRefGoogle ScholarPubMed
Phillips, G. L., Eminson, D. & Moss, B. (1978). A mechanism to account for macrophyte decline in progressively eutrophicated freshwaters. Aquatic Botany 4: 103126.CrossRefGoogle Scholar
Phillips, G., Pietilainen, O. P., Carvalho, L. et al. (2008). Chlorophyll-nutrient relationships of different lake types using a large European dataset. Aquatic Ecology 42: 213226.CrossRefGoogle Scholar
Pikosz, M., Messyasz, B. & Gąbka, M. (2017). Functional structure of algal mat (Cladophora glomerata) in a freshwater in western Poland. Ecological Indicators 74: 19.CrossRefGoogle Scholar
Polyakov, I. V., Tikka, K., Haapala, J. et al. (2022). Depletion of oxygen in the Bothnian Sea since the mid-1950s. Frontiers in Marine Science 9: 917879.CrossRefGoogle Scholar
Pretty, J. (2008). Agricultural sustainability: Concepts, principles and evidence. Philosophical Transactions of the Royal Society B 363: 447465.CrossRefGoogle ScholarPubMed
Qin, L.-Z., Suonan, Z., Kim, S. H. et al. (2021). Growth and reproductive responses of the seagrass Zostera marina to sediment nutrient enrichment. ICES Journal of Marine Science 78: 11601173.CrossRefGoogle Scholar
Qin, B., Zhou, J., Elser, J. J. et al. (2020). Water depth underpins the relative roles and fates of nitrogen and phosphorus in lakes. Environmental Science & Technology 54: 31913198.CrossRefGoogle ScholarPubMed
Quigg, A., Finkel, Z. V., Irwin, A. J. et al. (2003). The evolutionary inheritance of elemental stoichiometry in marine phytoplankton. Nature 425: 291294.CrossRefGoogle ScholarPubMed
Quigg, A., Irwin, A. J. & Finkel, Z. V. (2011). Evolutionary inheritance of elemental stoichiometry in phytoplankton. Proceedings of the Royal Society B: Biological Sciences 278: 526534.CrossRefGoogle ScholarPubMed
Quinlan, R., Filazzola, A., Mahdiyan, O. et al. (2021). Relationships of total phosphorus and chlorophyll in lakes worldwide. Limnology and Oceanography 66: 392404.CrossRefGoogle Scholar
Ratti, S., Knoll, A. H. & Giordano, M. (2011). Did sulfate availability facilitate the evolutionary expansion of chlorophyll a+c phytoplankton in the oceans?: Sulfate and evolution of chlorophyll a+c phytoplankton. Geobiology 9: 301312.CrossRefGoogle ScholarPubMed
Rattray, M. R., Howard-Williams, C. & Brown, J. M. A. (1991). Sediment and water as sources of nitrogen and phosphorus for submerged rooted aquatic macrophytes. Aquatic Botany 40: 225237.CrossRefGoogle Scholar
Raven, J. A. (1983). The transport and function of silicon in plants. Biological Reviews 58: 179207.CrossRefGoogle Scholar
Raven, J. A. (1997). Phagotrophy in phototrophs. Limnology and Oceanography 42: 198205.CrossRefGoogle Scholar
Raven, J. A., Caldeira, K., Elderfield, H. et al. (2005). Ocean Acidification due to Increasing Atmospheric Carbon Dioxide. Royal Society, London.Google Scholar
Redfield, A. C. (1934). On the proportions of organic derivatives in sea water and their relation to the composition of plankton. James Johnstone Memorial Volume, University Press of Liverpool pp. 176192.Google Scholar
Reid, A. J., Carlson, A. K., Creed, I. F. et al. (2019). Emerging threats and persistent conservation challenges for freshwater biodiversity. Biological Reviews 94: 849873.CrossRefGoogle ScholarPubMed
Remick, K. A. & Helmann, J. D. (2023). The elements of life: A biocentric tour of the periodic table. In Advances in Microbial Physiology 82: 1127.CrossRefGoogle ScholarPubMed
Reynolds, C. S. (2000). Hydroecology of river plankton: The role of variability in channel flow. Hydrological Processes 14: 31193132.3.0.CO;2-6>CrossRefGoogle Scholar
Reynolds, C. S. (2006). The Ecology of Phytoplankton, 1st ed. Cambridge University Press, Cambridge. https://doi.org/10.1017/CBO9780511542145.CrossRefGoogle Scholar
RoTAP (2012). Review of transboundary air pollution (RoTAP): Acidification, eutrophication, ground level ozone and heavy metals in the UK, Place of publication not identified: Centre for Ecology & Hydrology on behalf of Defra and the Devolved Administrations.Google Scholar
Rubio, L., and Fernández, J. A. (2019). Seagrasses, the unique adaptation of angiosperms to the marine environment: effect of high carbon and ocean acidification on energetics and ion homeostasis. In Halophytes and climate change: adaptive mechanisms and potential uses. Eds. Hasanuzzaman, M., Shabala, S. and Fujita, M. (Boston, Massachussets: CAB International), pp 81103.Google Scholar
Sakamoto, M. (1966). Primary production by phytoplankton community in some Japanese lakes and its dependence on lake depth. Archiv Fur Hydrobiologie 62: 128.Google Scholar
Salmaso, N. (2010). Long-term phytoplankton community changes in a deep subalpine lake: Responses to nutrient availability and climatic fluctuations. Freshwater Biology 55: 825846.CrossRefGoogle Scholar
Sand-Jensen, K. (1977). Effect of epiphytes on eelgrass photosynthesis. Aquatic Botany 3: 5563.CrossRefGoogle Scholar
Sand-Jensen, K. & Søndergaard, M. (1981). Phytoplankton and epiphyte development and their shading effect on submerged macrophytes in lakes of different nutrient status. Internationale Revue Der Gesamten Hydrobiologie 66: 529552.CrossRefGoogle Scholar
Scheffer, M., Hosper, S. H., Meijer, M. L. et al. (1993). Alternative equilibria in shallow lakes. Trends in Ecology and Evolution 8: 275279.CrossRefGoogle ScholarPubMed
Schindler, D. W. (1977). Evolution of phosphorus limitation in lakes. Science 195: 260262.CrossRefGoogle ScholarPubMed
Schindler, D. W., Fee, E. J. & Ruszczynski, T. (1978). Phosphorus input and its consequences for phytoplankton standing crop and production in Experimental Lakes Area and in similar lakes. Journal of the Fisheries Research Board of Canada 35: 190196.CrossRefGoogle Scholar
Schindler, D. W., Hecky, R. E., Findlay, D. L. et al. (2008.) Eutrophication of lakes cannot be controlled by reducing nitrogen input: Results of a 37-year whole-ecosystem experiment. Proceedings of the National Academy of Sciences USA 105: 1125411258.CrossRefGoogle ScholarPubMed
Schoelynck, J., Bal, K., Backx, H. et al. (2010). Silica uptake in aquatic and wetland macrophytes: A strategic choice between silica, lignin and cellulose? New Phytologist 186: 385391.CrossRefGoogle ScholarPubMed
Schvarcz, C. R., Wilson, S. T., Caffin, M. et al. (2022). Overlooked and widespread pennate diatom-diazotroph symbioses in the sea. Nature Communications 13: 799.CrossRefGoogle ScholarPubMed
Seitzinger, S. P., Mayorga, E., Bouwman, A. F. et al. (2010). Global river nutrient export: A scenario analysis of past and future trends. Global Biogeochemical Cycles 24: GB0A08.CrossRefGoogle Scholar
Sharoni, S. & Halevy, I. (2022). Geologic controls on phytoplankton elemental composition. Proceedings of the National Academy of Sciences USA 119: e2113263118.CrossRefGoogle ScholarPubMed
Sharpley, A., Foy, B. & Withers, P. (2000). Practical and innovative measures for the control of agricultural phosphorus losses to water: An overview. Journal of Environmental Quality 29: 19.CrossRefGoogle Scholar
Shepon, A., Eshel, G., Noor, E. et al. (2018). The opportunity cost of animal based diets exceeds all food losses. Proceedings of the National Academy of Sciences 115: 38043809.CrossRefGoogle ScholarPubMed
Sicko-Goad, L. M., Schelske, C. L. & Stoermer, E. F. (1984). Estimation of intracellular carbon and silica content of diatoms from natural assemblages using morphometric techniques. Limnology and Oceanography 29: 11701178.CrossRefGoogle Scholar
Skidmore, R. E., Maberly, S. C. & Whitton, B. A. (1998). Patterns of spatial and temporal variation in phytoplankton chlorophyll a in the River Trent and its tributaries. Science of the Total Environment 210: 357365.CrossRefGoogle Scholar
Smetacek, V. & Zingone, A. (2013). Green and golden seaweed tides on the rise. Nature 504: 8488.CrossRefGoogle ScholarPubMed
Smith, L. V., McMinn, A., Martin, A. et al. (2013). Preliminary investigation into the stimulation of phytoplankton photophysiology and growth by whale faeces. Journal of Experimental Marine Biology and Ecology 446: 19.CrossRefGoogle Scholar
Smith, S. M., Fox, S. E., Plaisted, H. K. et al. (2018). Changes in the thermal structure of freshwater lakes within Cape Cod National Seashore (Massachusetts, USA) from 1996 to 2014. Inland Waters 8: 3649.CrossRefGoogle Scholar
Smith, V. H., Joye, S. B. & Howarth, R. W. (2006). Eutrophication of freshwater and marine ecosystems. Limnology and Oceanography 51: 351355.CrossRefGoogle Scholar
Sommer, U., Adrian, R., De Senerpont Domis, L., Elser, J. J., Gaedke, U., Ibelings, B., Jeppesen, E., Lürling, M., Molinero, J. C., Mooij, W. M., van Donk, E., Winder, M., (2012). Beyond the Plankton Ecology Group (PEG) model: Mechanisms driving plankton succession. Annual Review of Ecology Evolution and Systematics 43: 429448.CrossRefGoogle Scholar
Søndergaard, M., Jeppesen, E., Lauridsen, T. L. et al. (2007). Lake restoration: Successes, failures and long-term effects. Journal of Applied Ecology 44: 10951105.CrossRefGoogle Scholar
Soria-Dengg, S., Reissbrodt, R. & Horstmann, U. (2001). Siderophores in marine coastal waters and their relevance for iron uptake by phytoplankton: Experiments with the diatom Phaeodactylum tricornutum. Marine Ecology Progress Series 220: 7382.CrossRefGoogle Scholar
Spears, B. M., Meis, S., Anderson, A. et al. (2013). Comparison of phosphorus (P) removal properties of materials proposed for the control of sediment p release in UK lakes. Science of the Total Environment 442: 103110.CrossRefGoogle ScholarPubMed
Steffen, W., Broadgate, W., Deutsch, L. et al. (2015). The trajectory of the Anthropocene: The Great Acceleration. The Anthropocene Review 2: 8198.CrossRefGoogle Scholar
Stockdale, A., Tipping, E., Fjellheim, A. et al. (2014). Recovery of macroinvertebrate species richness in acidified upland waters assessed with a field toxicity model. Ecological Indicators 37: 341350.CrossRefGoogle Scholar
Stoddard, J. L., Jeffries, D. S., Lukewille, A. et al. (1999). Regional trends in aquatic recovery from acidification in North America and Europe. Nature 401: 575578.CrossRefGoogle Scholar
Strojsová, A., Vrba, J., Nedoma, J. et al. (2003). Seasonal study of extracellular phosphatase expression in the phytoplankton of a eutrophic reservoir. European Journal of Phycology 38: 295306.CrossRefGoogle Scholar
Sutton, M. A. (2013). Our nutrient world: The challenge to produce more food and energy with less pollution. United Nations Environment Programme, Global Partnership on Nutrient Management, & International Nitrogen Initiative https://wedocs.unep.org/20.500.11822/10747.Google Scholar
Talling, J. F. (1971). The underwater light climate as a controlling factor in the production ecology of freshwater phytoplankton. Mitt. Int. Ver. Limnol 19: 214243.Google Scholar
Talling, J. F. (2010). Potassium – a non-limiting nutrient in fresh waters? Freshwater Reviews 3: 97104.CrossRefGoogle Scholar
Terrado, R., Monier, A., Edgar, R. et al. (2015). Diversity of nitrogen assimilation pathways among microbial photosynthetic eukaryotes. Journal of Phycology 51: 490506.CrossRefGoogle ScholarPubMed
Thronson, A. & Quigg, A. (2008). Fifty-five years of fish kills in coastal Texas. Estuaries and Coasts 31: 802813.CrossRefGoogle Scholar
Tickner, D., Opperman, J. J., Abell, R. et al. (2020). Bending the curve of global freshwater biodiversity loss: An emergency recovery plan. BioScience 70: 330342.CrossRefGoogle ScholarPubMed
Tilman, D., Isbell, F. & Cowles, J. M. (2014). Biodiversity and ecosystem functioning. Annual Review of Ecology, Evolution, and Systematics 45: 471493.CrossRefGoogle Scholar
Tong, Y., Zhang, W., Wang, X. et al. (2017). Decline in Chinese lake phosphorus concentration accompanied by shift in sources since 2006. Nature Geoscience 10: 507511.CrossRefGoogle Scholar
Twiss, M. R., Gouvêa, S. P., Bourbonniere, R. A. et al. (2005). Field investigations of trace metal effects on Lake Erie phytoplankton productivity. Journal of Great Lakes Research 31: 168179.CrossRefGoogle Scholar
UN Environment (2019). Global Environment Outlook – GEO-6: Summary for Policymakers. Cambridge University Press, Cambridge. https://doi.org/10.1017/9781108639217.Google Scholar
Urabe, J. (1993). N and P cycling coupled by grazers’ activities: food quality and nutrient release by zooplankton. Ecology 74: 23372350.CrossRefGoogle Scholar
Valiela, I., McClelland, J., Hauxwell, J. et al. (1997). Macroalgal blooms in shallow estuaries: Controls and ecophysiological and ecosystem consequences. Limnology and Oceanography 42: 11051118.CrossRefGoogle Scholar
Vallentyne, J. R. (1974). The Algal Bowl: Lakes and Man. Department of the Environment, Fisheries and Marine Service, Ottawa.Google Scholar
Van Mooy, B. A. S., Fredricks, H. F., Pedler, B. E. et al. (2009). Phytoplankton in the ocean use non-phosphorus lipids in response to phosphorus scarcity. Nature 458: 6972.CrossRefGoogle ScholarPubMed
van Puijenbroek, P. J. T. M., Beusen, A. H. W. & Bouwman, A. F. (2019). Global nitrogen and phosphorus in urban waste water based on the shared socio-economic pathways. Journal of Environmental Management 231: 446456.CrossRefGoogle ScholarPubMed
Vilmin, L., Mogollón, J. M., Beusen, A. H. W. et al. (2018). Forms and subannual variability of nitrogen and phosphorus loading to global river networks over the 20th century. Global and Planetary Change 163: 6785.CrossRefGoogle Scholar
Vollenweider, R. A. & Kerekes, J. (1980). The loading concept as basis for controlling eutrophication philosophy and preliminary-results of the OECD program on eutrophication. Progress in Water Technology 12: 538.Google Scholar
Volponi, S. N., Wander, H. L., Richardson, D. C. et al. (2023). Nutrient function over form: Organic and inorganic nitrogen additions have similar effects on lake phytoplankton nutrient limitation. Limnology & Oceanography 68: 307321.CrossRefGoogle Scholar
Vymazal, J. (2007). Removal of nutrients in various types of constructed wetlands. Science of the Total Environment 380: 4865.CrossRefGoogle ScholarPubMed
Wang, L., Robertson, D. M. & Garrison, P. J. (2007). Linkages between nutrients and assemblages of macroinvertebrates and fish in wadeable streams: Implication to nutrient criteria development. Environmental Management 39: 194212.CrossRefGoogle ScholarPubMed
Watson, S. B., Miller, C., Arhonditsis, G. et al. (2016). The re-eutrophication of Lake Erie: Harmful algal blooms and hypoxia. Harmful Algae 56: 4466.CrossRefGoogle ScholarPubMed
Wells, M. L., Trainer, V. L., Smayda, T. J. et al. (2015). Harmful algal blooms and climate change: Learning from the past and present to forecast the future. Harmful Algae 49: 6893.CrossRefGoogle ScholarPubMed
Williams, S. L. & Dethier, M. N. (2005). High and dry: Variation in net photosynthesis of the intertidal seaweed Fucus gardneri. Ecology 86: 23732379.CrossRefGoogle Scholar
Woolway, R. I., Jennings, E., Shatwell, T. et al. (2021). Lake heatwaves under climate change. Nature 589: 402407.CrossRefGoogle ScholarPubMed
Woolway, R. I., Jones, I. D., Maberly, S. C. et al. (2016). Diel surface temperature range scales with lake size. PLOS ONE 11: e0152466.CrossRefGoogle ScholarPubMed
Wright, R. F., Norton, S. A., Brakke, D. F. et al. (1988). Experimental verification of episodic acidification of freshwaters by sea salts. Nature 334: 422424.CrossRefGoogle Scholar
Wu, M., McCain, J. S. P., Rowland, E. et al. (2019). Manganese and iron deficiency in Southern Ocean Phaeocystis antarctica populations revealed through taxon-specific protein indicators. Nature Communications 10: 3582. https://doi.org/10.1038/s41467–019–11426-z.CrossRefGoogle ScholarPubMed
Wu, Z., Li, J., Sun, Y., … Liu, Y. (2022). Imbalance of global nutrient cycles exacerbated by the greater retention of phosphorus over nitrogen in lakes. Nature Geoscience 15: 464468.CrossRefGoogle Scholar
Wurtsbaugh, W. A., Paerl, H. W. & Dodds, W. K. (2019). Nutrients, eutrophication and harmful algal blooms along the freshwater to marine continuum. Wiley Interdisciplinary Reviews-Water 6: e1373. https://doi.org/10.1002/wat2.1373.CrossRefGoogle Scholar
Wymore, A. S., Johnes, P. J., Bernal, S. et al. (2021). Gradients of anthropogenic nutrient enrichment alter N composition and DOM stoichiometry in freshwater ecosystems. Global Biogeochemical Cycles 35: e2021GB006953. https://doi.org/10.1029/2021GB006953.CrossRefGoogle Scholar
Yates, C. A., Johnes, P. J., Owen, A. T. et al. (2019). Variation in dissolved organic matter (DOM) stoichiometry in U.K. freshwaters: Assessing the influence of land cover and soil C:N ratio on DOM composition. Limnology and Oceanography 64: 23282340.CrossRefGoogle Scholar
Yu, C., Huang, X., Chen, H. et al. (2019). Managing nitrogen to restore water quality in China. Nature 567: 516520.CrossRefGoogle ScholarPubMed
Zak, D., Hupfer, M., Cabezas, A. et al. (2021). Sulphate in freshwater ecosystems: A review of sources, biogeochemical cycles, ecotoxicological effects and bioremediation. Earth-Science Reviews 212: 103446.CrossRefGoogle Scholar
Zehr, J. P., Jenkins, B. D., Short, S. M. et al. (2003). Nitrogenase gene diversity and microbial community structure: A cross-system comparison. Environmental Microbiology 5: 539554.CrossRefGoogle ScholarPubMed
Zehr, J. P., Waterbury, J. B., Turner, P. J. et al. (2001). Unicellular cyanobacteria fix N2 in the subtropical North Pacific Ocean. Nature 412: 635638.CrossRefGoogle ScholarPubMed
Zhang, Y., Jeppesen, E., Liu, X. et al. (2017). Global loss of aquatic vegetation in lakes. Earth-Science Reviews 173: 259265.CrossRefGoogle Scholar

References

Abouraïcha, E. F., El Alaoui-Talibi, Z., Tadlaoui-Ouafi, A. et al. (2017). Glucuronan and oligoglucuronans isolated from green algae activate natural defense responses in apple fruit and reduce postharvest blue and gray mold decay. Journal of Applied Phycology 29: 471480.CrossRefGoogle Scholar
Achitouv, E., Metzger, P., Rager, M. et al. (2004). C31–C34 methylated squalenes from a Bolivian strain of Botryococcus braunii. Phytochemistry 65: 31593165.CrossRefGoogle ScholarPubMed
Adey, W. H., Kangas, P. C. & Mulbry, W. (2011). Algal turf scrubbing: Cleaning surface waters with solar energy while producing a biofuel. BioScience 61: 434441.CrossRefGoogle Scholar
Ahmed, F., Zhou, W. & Schenk, P. M. (2015). Pavlova lutheri is a high-level producer of phytosterols. Algal Research 10: 210217.CrossRefGoogle Scholar
Angioni, S., Millia, L., Mustarelli, P. et al. (2018). Photosynthetic microbial fuel cell with polybenzimidazole membrane: Synergy between bacteria and algae for wastewater removal and biorefinery. Helion 4: e00560.Google ScholarPubMed
Araújo, M., Rema, P., Sousa-Pinto, I. et al. (2016). Dietary inclusion of IMTA-cultivated Gracilaria vermiculophylla in rainbow trout (Oncorhynchus mykiss) diets: Effects on growth, intestinal morphology, tissue pigmentation, and immunological response. Journal of Applied Phycology 28: 679689.CrossRefGoogle Scholar
Arioli, T., Mattner, S. W. & Winberg, P. C. (2015). Applications of seaweed extracts in Australian agriculture: Past, present and future. Journal of Applied Phycology 27: 20072015.CrossRefGoogle ScholarPubMed
Baxter, L., Brain, R. A., Lissemore, L. et al. (2016). Influence of light, nutrients, and temperature on the toxicity of atrazine to the algal species Raphidocelis subcapitata: Implications for the risk assessment of herbicides. Ecotoxicology and Environmental Safety 132: 250259.CrossRefGoogle Scholar
Belay, A. (2013). Biology and industrial production of Arthrospira (Spirulina). In: Richmond, A. & Hu, Q. (eds.) Handbook of Microalgal Culture: Applied Phycology and Biotechnology. Blackwell, Oxford, pp. 339358.CrossRefGoogle Scholar
Ben Ouada, S., Ben Ali, R., Leboulanger, C. et al. (2018). Effect and removal of bisphenol A by two extremophilic microalgal strains (Chlorophyta). Journal of Applied Phycology 30: 17651776.CrossRefGoogle Scholar
Bhati, R. & Mallick, N. (2015). Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) copolymer production by the diazotrophic cyanobacterium Nostoc muscorum Agardh: Process optimization and polymer characterization. Algal Research 7: 7885.CrossRefGoogle Scholar
Blanco, A. M., Moreno, J., Del Campo, J. A. et al. (2007). Outdoor cultivation of lutein-rich cells of Muriellopsis sp. in open ponds. Applied Microbiology and Biotechnology 73: 12591266.CrossRefGoogle ScholarPubMed
Bolton, J. J., Robertson-Andersson, D. V., Shuuluka, D. et al. (2009). Growing Ulva (Chlorophyta) in integrated systems as a commercial crop for abalone feed in South Africa: A SWOT analysis. Journal of Applied Phycology 21: 575583.CrossRefGoogle Scholar
Borowitzka, M. A. (1997). Algae for aquaculture: Opportunities and constraints. Journal of Applied Phycology 9: 393401.CrossRefGoogle Scholar
Borowitzka, M. A. (2013a). Dunaliella: Biology, production, and markets. In: Richmond, A. & Hu, Q. (eds.) Handbook of Microalgal Culture. John Wiley and Sons Ltd., Chichester, UK, pp. 359368.CrossRefGoogle Scholar
Borowitzka, M. A. (2013b). High-value products from microalgae – Their development and commercialisation. Journal of Applied Phycology 25: 743756.CrossRefGoogle Scholar
Borowitzka, M. A. (2018). Microalgae in medicine and human health: A historical perspective. In: Levine, I. A. & Fleurence, J. (eds.) Microalgae in Health and Disease Prevention. Academic Press, London, pp. 195210.CrossRefGoogle Scholar
Borowitzka, M. A. & Moheimani, N. R. (2013). Sustainable biofuels from algae. Mitigation and Adaptation Strategies for Global Change 18: 1325.CrossRefGoogle Scholar
Buck, B. H., Nevejan, N., Wille, M. et al. (2017). Offshore and multi-use aquaculture with extractive species: Seaweeds and bivalves. In: Buck, B. H. & Langan, R. (eds.) Aquaculture Perspective of Multi-use Sites in the Open Ocean. Springer, Dordrecht, pp 2369.CrossRefGoogle Scholar
Buck, B. H., Troell, M. F., Krause, G. et al. (2018). State of the art and challenges for offshore Integrated Multi-Trophic Aquaculture (IMTA). Frontiers in Marine Science 5: 165.CrossRefGoogle Scholar
Cabrita, A. R. J., Correia, A., Rodrigues, A. R. et al. (2017). Assessing in vivo digestibility and effects on immune system of sheep fed alfalfa hay supplemented with a fixed amount of Ulva rigida and Gracilaria vermiculophylla. Journal of Applied Phycology 29: 10571067.CrossRefGoogle Scholar
Cabrita, A. R. J., Maia, M. R. G., Oliveira, H. M. et al. (2016). Tracing seaweeds as mineral sources for farm-animals. Journal of Applied Phycology 28: 31353150.CrossRefGoogle Scholar
Casal, C., Cuaresma, M., Vega, J. M. et al. (2011). Enhanced productivity of a lutein-enriched novel acidophile microalga grown on urea. Marine Drugs 9: 2942.CrossRefGoogle Scholar
Chalifour, A. & Tam, N. F.-Y. (2016). Tolerance of cyanobacteria to the toxicity of BDE-47 and their removal ability. Chemosphere 164: 451461.CrossRefGoogle Scholar
Chatzikonstantinou, M., Vlachakis, D., Chronopoulou, E. et al. (2017). The glutathione transferase family of Chlamydomonas reinhardtii: Identification and characterization of novel Sigma class-like enzymes. Algal Research 24: 237250.CrossRefGoogle Scholar
Chen, C.-Y. & Liu, C.-C. (2018). Optimization of lutein production with a two-stage mixotrophic cultivation system with Chlorella sorokiniana MB-1. Bioresource Technology 262: 7479.CrossRefGoogle ScholarPubMed
Chen, C.-Y., Lu, I. C., Nagarajan, D. et al. (2018). A highly efficient two-stage cultivation strategy for lutein production using heterotrophic culture of Chlorella sorokiniana MB-1-M12. Bioresource Technology 253: 141147.CrossRefGoogle ScholarPubMed
Chisti, Y. (2013). Constraints to commercialization of algal fuels. Journal of Biotechnology 167: 201214.CrossRefGoogle ScholarPubMed
Cho, C.-W., Pham, T. P. T., Kim, S. et al. (2009). Toxicity assessment of common organic solvents using a biosensor based on algal photosynthetic activity measurement. Journal of Applied Phycology 21: 683689.CrossRefGoogle Scholar
Chrismadha, T. & Borowitzka, M. A. (1994). Effect of cell density and irradiance on growth, proximate composition and eicosapentaenoic acid production of Phaeodactylum tricornutum grown in a tubular photobioreactor. Journal of Applied Phycology 6: 6774.CrossRefGoogle Scholar
Chu, W.-L., See, Y.-C. & Phang, S.-M. (2008). Use of immobilised Chlorella vulgaris for the removal of colour from textile dyes. Journal of Applied Phycology 21: 641.CrossRefGoogle Scholar
Coppens, J., Grunert, O., Van Den Hende, S. et al. (2016). The use of microalgae as a high-value organic slow-release fertilizer results in tomatoes with increased carotenoid and sugar levels. Journal of Applied Phycology 28: 23672377.CrossRefGoogle Scholar
Cordero, B. F., Obraztsova, I., Couso, I. et al. (2011). Enhancement of lutein production in Chlorella sorokiniana (Chlorophyta) by improvement of culture conditions and random mutagenesis. Marine Drugs 9: 16071624.CrossRefGoogle ScholarPubMed
Craggs, R., Park, J., Heubeck, S. & Sutherland, D. (2014). High rate algal pond systems for low-energy wastewater treatment, nutrient recovery and energy production. New Zealand Journal of Botany 52: 6073.CrossRefGoogle Scholar
Craggs, R., Park, J., Sutherland, D. & Heubeck, S. (2015). Economic construction and operation of hectare-scale wastewater treatment enhanced pond systems. Journal of Applied Phycology 27: 19131922.CrossRefGoogle Scholar
Craigie, J. S. (2011). Seaweed extract stimuli in plant science and agriculture. Journal of Applied Phycology 23: 371393.CrossRefGoogle Scholar
Cui, Y., Rashid, N., Hu, N. et al. (2014). Electricity generation and microalgae cultivation in microbial fuel cell using microalgae-enriched anode and bio-cathode. Energy Conversion and Management 79: 674680.CrossRefGoogle Scholar
Czarny, K., Szczukocki, D., Krawczyk, B. et al. (2018). Inhibition of growth of Anabaena variabilis population by single and mixed steroid hormones. Journal of Applied Phycology 31: 389398.CrossRefGoogle Scholar
De Wilt, A., Butkovskyi, A., Tuantet, K. et al. (2016). Micropollutant removal in an algal treatment system fed with source separated wastewater streams. Journal of Hazardous Materials 304: 8492.CrossRefGoogle Scholar
Del Campo, J. A., Moreno, J., Rodriguez, H. et al. (2000). Carotenoid content of chlorophycean microalgae: Factors determining lutein accumulation in Muriellopsis sp. (Chlorophyta). Journal of Biotechnology 76: 5159.CrossRefGoogle ScholarPubMed
Dunn, K., Maart, B. & Rose, P. (2013). Arthrospira (Spirulina) in tannery wastewaters. Part 2: Evaluation of tannery wastewater as production media for the mass culture of Arthrospira biomass. Water SA 39: 279284.Google Scholar
Egeland, E. S. (2016). Carotenoids. In: Borowitzka, M. A., Beardall, J. & Raven, J. A. (eds.) The Physiology of Microalgae. Springer, Dordrecht, pp. 507563.CrossRefGoogle Scholar
Eilers, U., Bikoulis, A., Breitenbach, J. et al. (2016). Limitations in the biosynthesis of fucoxanthin as targets for genetic engineering in Phaeodactylum tricornutum. Journal of Applied Phycology 28: 123129.CrossRefGoogle Scholar
Elizondo-Reyna, E., Medina-González, R., Nieto-López, M. G. et al. (2016). Consumption of Ulva clathrata as a dietary supplement stimulates immune and lipid metabolism genes in Pacific white shrimp Litopenaeus vannamei. Journal of Applied Phycology 28: 36673677.CrossRefGoogle Scholar
Esserti, S., Smaili, A., Rifai, L. A. et al. (2017). Protective effect of three brown seaweed extracts against fungal and bacterial diseases of tomato. Journal of Applied Phycology 29: 10811093.CrossRefGoogle Scholar
Evans, F. D. & Critchley, A. T. (2014). Seaweeds for animal production use. Journal of Applied Phycology 26: 891899.CrossRefGoogle Scholar
Fallowfield, H. J., Young, P., Taylor, M. J. et al. (2018). Independent validation and regulatory agency approval for high rate algal ponds to treat wastewater from rural communities. Environmental Science: Water Research and Technology 4: 195205.Google Scholar
Fan, K., Aki, T., Chen, F. et al. (2010). Enhanced production of squalene in the thraustochytrid Aurantiochytrium mangrovei by medium optimization and treatment with terbinafine. World Journal of Microbiology and Biotechnology 26: 13031309.CrossRefGoogle ScholarPubMed
Fischer, F. (2018). Photoelectrode, photovoltaic and photosynthetic microbial fuel cells. Renewable and Sustainable Energy Reviews 90: 1627.CrossRefGoogle Scholar
Forootanfar, H., Shakibaie, M., Bagherzadeh, Z. et al. (2013). The removal of ρ-chlorophenol in aqueous cultures with free and alginate-immobilized cells of the microalga Tetraselmis suecica. Journal of Applied Phycology 25: 5157.CrossRefGoogle Scholar
Francis, T. L., Maneveldt, G. W. & Venter, J. (2008). Determining the most appropriate feeding regime for the South African abalone Haliotis midae Linnaeus grown on kelp. Journal of Applied Phycology 20: 597602.CrossRefGoogle Scholar
Garcia-Gonzalez, J. & Sommerfeld, M. (2016). Biofertilizer and biostimulant properties of the microalga Acutodesmus dimorphus. Journal of Applied Phycology 28: 10511061.CrossRefGoogle ScholarPubMed
García De Llasera, M. P., Olmos-Espejel, J. D. J., Díaz-Flores, G. et al. (2016). Biodegradation of benzo(a)pyrene by two freshwater microalgae Selenastrum capricornutum and Scenedesmus acutus: A comparative study useful for bioremediation. Environmental Science and Pollution Research 23: 33653375.CrossRefGoogle ScholarPubMed
Gentili, F. G. & Fick, J. (2017). Algal cultivation in urban wastewater: An efficient way to reduce pharmaceutical pollutants. Journal of Applied Phycology 29: 255262.CrossRefGoogle ScholarPubMed
Gómez-Loredo, A., Benavides, J. & Rito-Palomares, M. (2016). Growth kinetics and fucoxanthin production of Phaeodactylum tricornutum and Isochrysis galbana cultures at different light and agitation conditions. Journal of Applied Phycology 28: 849860.CrossRefGoogle Scholar
Granado-Lorencio, F., Herrero-Barbudo, C., Acién Fernández, F. G. et al. (2009). In vitro bioaccessibility of lutein and zeaxanthin in the microalgae Scenedesmus almeriensis. Food Chemistry 114: 747752.CrossRefGoogle Scholar
Green, F. B., Bernstone, L. S., Lundquist, T. J. et al. (1996). Advanced integrated wastewater pond systems for nitrogen removal. Water Science and Technology 33: 207217.CrossRefGoogle Scholar
Guiry, M. D. (2012). How many species of algae are there? Journal of Phycology 48: 10571063.CrossRefGoogle ScholarPubMed
Guo, B., Liu, B., Yang, B. et al. (2016). Screening of diatom strains and characterization of Cyclotella cryptica as a potential fucoxanthin producer. Marine Drugs 14: 125. https://doi.org/10.3390/md14070125.CrossRefGoogle ScholarPubMed
Haritash, A. K. & Kaushik, C. P. (2009). Biodegradation aspects of Polycyclic Aromatic Hydrocarbons (PAHs): A review. Journal of Hazardous Materials 169: 115.CrossRefGoogle ScholarPubMed
Heo, J., Shin, D.-S., Cho, K. et al. (2018). Indigenous microalga Parachlorella sp. JD-076 as a potential source for lutein production: Optimization of lutein productivity via regulation of light intensity and carbon source. Algal Research 33: 17.CrossRefGoogle Scholar
Hernández-Herrera, R. M., Santacruz-Ruvalcaba, F., Zañudo-Hernández, J. et al. (2016). Activity of seaweed extracts and polysaccharide-enriched extracts from Ulva lactuca and Padina gymnospora as growth promoters of tomato and mung bean plants. Journal of Applied Phycology 28: 25492560.CrossRefGoogle Scholar
Huang, W., Lin, Y., He, M. et al. (2018). Induced high-yield production of zeaxanthin, lutein, and β-carotene by a mutant of Chlorella zofingiensis. Journal of Agricultural and Food Chemistry 66: 891897.CrossRefGoogle ScholarPubMed
Hurtado, A. Q. & Critchley, A. T. (2018). A review of multiple biostimulant and bioeffector benefits of AMPEP, an extract of the brown alga Ascophyllum nodosum, as applied to the enhanced cultivation and micropropagation of the commercially important red algal carrageenophyte Kappaphycus alvarezii and its selected cultivars. Journal of Applied Phycology 30: 28592873.CrossRefGoogle Scholar
Ishika, T., Moheimani, N. R., Bahri, P. A. et al. (2017). Halo-adapted microalgae for fucoxanthin production: Effect of incremental increase in salinity. Algal Research 28: 6673.CrossRefGoogle Scholar
Jin, E., Feth, B. & Melis, A. (2003). A mutant of the green alga Dunaliella salina constitutively accumulates zeaxanthin under all growth conditions. Biotechnology and Bioengineering 81: 115124.CrossRefGoogle ScholarPubMed
Kenny, P. & Flynn, K. J. (2017). Physiology limits commercially viable photoautotrophic production of microalgal biofuels. Journal of Applied Phycology 29: 27132727.CrossRefGoogle ScholarPubMed
Kim, M., Ahn, J., Jeon, H. et al. (2017). Development of a Dunaliella tertiolecta strain with increased zeaxanthin content using random mutagenesis. Marine Drugs 15: 189.CrossRefGoogle ScholarPubMed
Kim, S. M., Kang, S.-W., Kwon, O.-N. et al. (2012). Fucoxanthin as a major carotenoid in Isochrysis aff. galbana: Characterization of extraction for commercial application. Journal of the Korean Society for Applied Biological Chemistry 55: 477483.CrossRefGoogle Scholar
Koo, S., Cha, K., Song, D.-G. et al. (2012). Optimization of pressurized liquid extraction of zeaxanthin from Chlorella ellipsoidea. Journal of Applied Phycology 24: 725730.CrossRefGoogle Scholar
Kovalenko, I., Zdyrko, B., Magasinski, A. et al. (2011). A major constituent of brown algae for use in high-capacity Li-Ion batteries. Science 334: 7579.CrossRefGoogle Scholar
Leonardo, S., Prieto-Simón, B. & Campàs, M. (2016). Past, present and future of diatoms in biosensing. TrAC Trends in Analytical Chemistry 79: 276285.CrossRefGoogle Scholar
Leusch, F. D. L., Neale, P. A., Arnal, C. et al. (2018). Analysis of endocrine activity in drinking water, surface water and treated wastewater from six countries. Water Research 139: 1018.CrossRefGoogle ScholarPubMed
Li, X., Norman, H. C., Kinley, R. D. et al. (2018). Asparagopsis taxiformis decreases enteric methane production from sheep. Animal Production Science 58: 681688.CrossRefGoogle Scholar
Lin, J.-H., Lee, D.-J. & Chang, J.-S. (2015). Lutein production from biomass: Marigold flowers versus microalgae. Bioresource Technology 184: 421428.CrossRefGoogle ScholarPubMed
Lin, J., Huang, L., Yu, J. et al. (2016). Fucoxanthin, a marine carotenoid, reverses scopolamine-induced cognitive impairments in mice and inhibits acetylcholinesterase in vitro. Marine Drugs 14: 67.CrossRefGoogle ScholarPubMed
Liu, B., Zhang, S.-G. & Chang, C.-C. (2018). Emerging Pollutants – Part II: Treatment. Water Environment Research 90: 17921820.CrossRefGoogle ScholarPubMed
Longo, S., D’antoni, B. M., Bongards, M. et al. (2016). Monitoring and diagnosis of energy consumption in wastewater treatment plants. A state of the art and proposals for improvement. Applied Energy 179: 12511268.CrossRefGoogle Scholar
Luo, X., Su, P. & Zhang, W. (2015). Advances in microalgae-derived phytosterols for functional food and pharmaceutical applications. Marine Drugs 13: 42314254.CrossRefGoogle ScholarPubMed
Machado, L., Kinley, R. D., Magnusson, M. et al. (2015). The potential of macroalgae for beef production systems in Northern Australia. Journal of Applied Phycology 27: 20012005.CrossRefGoogle Scholar
Machado, L., Magnusson, M., Paul, N. A. et al. (2016a). Dose-response effects of Asparagopsis taxiformis and Oedogonium sp. on in vitro fermentation and methane production. Journal of Applied Phycology 28: 14431452.CrossRefGoogle Scholar
Machado, L., Magnusson, M., Paul, N. A. et al. (2016b). Identification of bioactives from the red seaweed Asparagopsis taxiformis that promote antimethanogenic activity in vitro. Journal of Applied Phycology 28: 31173126.CrossRefGoogle Scholar
Machado, L., Tomkins, N., Magnusson, M. et al. (2018). In vitro response of rumen microbiota to the antimethanogenic red macroalga Asparagopsis taxiformis. Microbial Ecology 75: 811818.CrossRefGoogle Scholar
Machado, M. D. & Soares, E. V. (2018). Sensitivity of freshwater and marine green algae to three compounds of emerging concern. Journal of Applied Phycology 31: 399408.CrossRefGoogle Scholar
Maeda, H., Matsumoto, M., Maeda, Y. et al. (2017). Utilization of diatom frustules for thermal management applications. Journal of Applied Phycology 29: 19071911.CrossRefGoogle Scholar
Maes, H. M., Maletz, S. X., Ratte, H. T. et al. (2014). Uptake, elimination, and biotransformation of 17α-ethinylestradiol by the freshwater alga Desmodesmus subspicatus. Environmental Science and Technology 48: 1235412361.CrossRefGoogle ScholarPubMed
Maher, S., Alsawat, M., Kumeria, T. et al. (2015). Luminescent silicon diatom replicas: Self-reporting and degradable drug carriers with biologically derived shape for sustained delivery of therapeutics. Advanced Functional Materials 25: 51075116.CrossRefGoogle Scholar
Maher, S., Kumeria, T., Aw, M. S. et al. (2018). Diatom silica for biomedical applications: Recent progress and advances. Advanced Healthcare Materials 7: 1800552.CrossRefGoogle ScholarPubMed
Mani, S. D. & Nagarathnam, R. (2018). Sulfated polysaccharide from Kappaphycus alvarezii (Doty) Doty ex P.C. Silva primes defense responses against anthracnose disease of Capsicum annuum Linn. Algal Research 32: 121130.CrossRefGoogle Scholar
Martin, L. (2015). Fucoxanthin and its metabolite fucoxanthinol in cancer prevention and treatment. Marine Drugs 13: 4784.CrossRefGoogle ScholarPubMed
Mata, L., Schuenhoff, A. & Santos, R. (2010). A direct comparison of the performance of the seaweed biofilters, Asparagopsis armata and Ulva rigida. Journal of Applied Phycology 22: 639644.CrossRefGoogle Scholar
Matthews, C. B., Wright, C., Kuo, A. et al. (2017). Reexamining opportunities for therapeutic protein production in eukaryotic microorganisms. Biotechnology and Bioengineering 114: 24322444.CrossRefGoogle ScholarPubMed
Mattner, S. W., Milinkovic, M. & Arioli, T. (2018). Increased growth response of strawberry roots to a commercial extract from Durvillaea potatorum and Ascophyllum nodosum. Journal of Applied Phycology 30: 29432951.CrossRefGoogle ScholarPubMed
McClure, D. D., Luiz, A., Gerber, B. et al. (2018). An investigation into the effect of culture conditions on fucoxanthin production using the marine microalgae Phaeodactylum tricornutum. Algal Research 29: 4148.CrossRefGoogle Scholar
McCormick, A. J., Bombelli, P., Bradley, R. W. et al. (2015). Biophotovoltaics: Oxygenic photosynthetic organisms in the world of bioelectrochemical systems. Energy and Environmental Science 8: 10921109.CrossRefGoogle Scholar
Miyashita, K. & Hosokawa, M. (2017). Fucoxanthin in the management of obesity and its related disorders. Journal of Functional Foods 36: 195202.CrossRefGoogle Scholar
Mohibbullah, M., Haque, M. N., Khan, M. N. A. et al. (2018). Neuroprotective effects of fucoxanthin and its derivative fucoxanthinol from the phaeophyte Undaria pinnatifida attenuate oxidative stress in hippocampal neurons. Journal of Applied Phycology 30: 32433252.CrossRefGoogle Scholar
Molino, J. V. D., De Carvalho, J. C. M. & Mayfield, S. P. (2018). Comparison of secretory signal peptides for heterologous protein expression in microalgae: Expanding the secretion portfolio for Chlamydomonas reinhardtii. PLOS ONE 13: e0192433.CrossRefGoogle ScholarPubMed
Muñoz, R. & Guieysse, B. (2006). Algal–bacterial processes for the treatment of hazardous contaminants: A review. Water Research 40: 27992815.CrossRefGoogle ScholarPubMed
Muradian, K., Vaiserman, A., Min, K. J. et al. (2015). Fucoxanthin and lipid metabolism: A minireview. Nutrition, Metabolism and Cardiovascular Diseases 25: 891897.CrossRefGoogle ScholarPubMed
Nakazawa, A., Kokubun, Y., Matsuura, H. et al. (2014). TLC screening of thraustochytrid strains for squalene production. Journal of Applied Phycology 26: 2941.CrossRefGoogle Scholar
Nazos, T. T., Kokarakis, E. J. & Ghanotakis, D. F. (2017). Metabolism of xenobiotics by Chlamydomonas reinhardtii: Phenol degradation under conditions affecting photosynthesis. Photosynthesis Research 131: 3140.CrossRefGoogle ScholarPubMed
Nelson, D. & Werck-Reichhart, D. (2011). A P450-centric view of plant evolution. The Plant Journal 66: 194211.CrossRefGoogle ScholarPubMed
Ng, F.-L., Phang, S.-M., Periasamy, V. et al. (2018). Algal biophotovoltaic (BPV) device for generation of bioelectricity using Synechococcus elongatus (Cyanophyta). Journal of Applied Phycology 30: 29812988.CrossRefGoogle Scholar
Norvill, Z. N., Toledo-Cervantes, A., Blanco, S. et al. (2017). Photodegradation and sorption govern tetracycline removal during wastewater treatment in algal ponds. Bioresource Technology 232: 3543.CrossRefGoogle ScholarPubMed
Oswald, W. J. (2003). My sixty years in applied algology. Journal of Applied Phycology 15: 99106.CrossRefGoogle Scholar
Otto, B., Beuchel, C., Liers, C. et al. (2015). Laccase-like enzyme activities from chlorophycean green algae with potential for bioconversion of phenolic pollutants. FEMS Microbiology Letters 362: fnv072.CrossRefGoogle ScholarPubMed
Papazi, A. & Kotzabasis, K. (2013). ‘Rational’ management of dichlorophenols biodegradation by the microalga Scenedesmus obliquus. PLOS ONE 8: e61682.CrossRefGoogle ScholarPubMed
Peng, F.-Q., Ying, G.-G., Yang, B. et al. (2014). Biotransformation of progesterone and norgestrel by two freshwater microalgae (Scenedesmus obliquus and Chlorella pyrenoidosa): Transformation kinetics and products identification. Chemosphere 95: 581588.CrossRefGoogle ScholarPubMed
Petersen, K., Heiaas, H. H. & Tollefsen, K. E. (2014). Combined effects of pharmaceuticals, personal care products, biocides and organic contaminants on the growth of Skeletonema pseudocostatum. Aquatic Toxicology 150: 4554.CrossRefGoogle ScholarPubMed
Petroutsos, D., Katapodis, P., Christakopoulos, P. et al. (2007). Removal of p-chlorophenol by the marine microalga Tetraselmis marina. Journal of Applied Phycology 19: 485490.CrossRefGoogle Scholar
Petrushkina, M., Gusev, E., Sorokin, B. et al. (2017). Fucoxanthin production by heterokont microalgae. Algal Research 24: 387393.CrossRefGoogle Scholar
Pflugmacher, S., Schröder, P. & Sandermann, H. (2000). Taxonomic distribution of plant glutathione S-transferases acting on xenobiotics. Phytochemistry 54: 267273.CrossRefGoogle ScholarPubMed
Porse, H. & Rudolph, B. (2017). The seaweed hydrocolloid industry: 2016 updates, requirements, and outlook. Journal of Applied Phycology 29: 21872200.CrossRefGoogle Scholar
Pramanick, B., Brahmachari, K., Mahapatra, B. S. et al. (2017). Growth, yield and quality improvement of potato tubers through the application of seaweed sap derived from the marine alga Kappaphycus alvarezii. Journal of Applied Phycology 29: 32533260.CrossRefGoogle Scholar
Ramkissoon, A., Ramsubhag, A. & Jayaraman, J. (2017). Phytoelicitor activity of three Caribbean seaweed species on suppression of pathogenic infections in tomato plants. Journal of Applied Phycology 29: 32353244.CrossRefGoogle Scholar
Ramos-Martinez, E. M., Fimognari, L. & Sakuragi, Y. (2017). High-yield secretion of recombinant proteins from the microalga Chlamydomonas reinhardtii. Plant Biotechnology Journal 15: 12141224.CrossRefGoogle ScholarPubMed
Rasala, B. A. & Mayfield, S. P. (2015). Photosynthetic biomanufacturing in green algae; production of recombinant proteins for industrial, nutritional, and medical uses. Photosynthesis Research 123: 227239.CrossRefGoogle ScholarPubMed
Ratledge, C. (2010). Single cell oils for the 21st century. In: Cohen, Z. & Ratledge, C. (eds.) Single Cell Oils. Microbial and Algal Oils. AOCS Press, Urbana, pp. 326.CrossRefGoogle Scholar
Righini, H., Roberti, R. & Baraldi, E. (2018). Use of algae in strawberry management. Journal of Applied Phycology 30: 35513564.CrossRefGoogle Scholar
Robert, F. O., Pandhal, J. & Wright, P. C. (2010). Exploiting cyanobacterial P450 pathways. Current Opinion in Microbiology 13: 301306.CrossRefGoogle ScholarPubMed
Robinson, N., Winberg, P. & Kirkendale, L. (2013). Genetic improvement of macroalgae: Status to date and needs for the future. Journal of Applied Phycology 25: 703716.CrossRefGoogle Scholar
Roy, J. J., Sun, L. & Ji, L. (2014). Microalgal proteins: A new source of raw material for production of plywood adhesive. Journal of Applied Phycology 26: 14151422.CrossRefGoogle Scholar
Sánchez, J., Fernández-Sevilla, J., Acién, F. et al. (2008). Biomass and lutein productivity of Scenedesmus almeriensis: Influence of irradiance, dilution rate and temperature. Applied Microbiology and Biotechnology 79: 719729.CrossRefGoogle ScholarPubMed
Saravana, P. S., Getachew, A. T., Cho, Y.-J. et al. (2017). Influence of co-solvents on fucoxanthin and phlorotannin recovery from brown seaweed using supercritical CO2. The Journal of Supercritical Fluids 120: 295303.CrossRefGoogle Scholar
Schreiber, C., Schiedung, H., Harrison, L. et al. (2018). Evaluating potential of green alga Chlorella vulgaris to accumulate phosphorus and to fertilize nutrient-poor soil substrates for crop plants. Journal of Applied Phycology 30: 28272836.CrossRefGoogle Scholar
Scodelaro Bilbao, P. G., Damiani, C., Salvador, G. A. et al. (2016). Haematococcus pluvialis as a source of fatty acids and phytosterols: Potential nutritional and biological implications. Journal of Applied Phycology 28: 32833294.CrossRefGoogle Scholar
Selvaraj, V., Muthukumar, A., Nagamony, P. et al. (2018). Detection of typhoid fever by diatom-based optical biosensor. Environmental Science and Pollution Research 25: 2038520390.CrossRefGoogle ScholarPubMed
Semple, K. T. & Cain, R. B. (1996). Biodegradation of phenols by the alga Ochromonas danica. Applied and Environmental Microbiology 62: 12651273.CrossRefGoogle ScholarPubMed
Semple, K. T., Cain, R. B. & Schmidt, S. (1999). Biodegradation of aromatic compounds by microalgae. FEMS Microbiology Letters 170: 291300.CrossRefGoogle Scholar
Shah, M. M. R., Liang, Y., Cheng, J. J. et al. (2016). Astaxanthin-producing green microalga Haematococcus pluvialis: From single cell to high value commercial products. Frontiers in Plant Science 7: 531.CrossRefGoogle ScholarPubMed
Shan, T. F., Pang, S. J., Li, J. et al. (2016). Breeding of an elite cultivar Haibao No. 1 of Undaria pinnatifida (Phaeophyceae) through gametophyte clone crossing and consecutive selection. Journal of Applied Phycology 28: 24192426.CrossRefGoogle Scholar
Shannon, E. & Abu-Ghannam, N. (2017). Optimisation of fucoxanthin extraction from Irish seaweeds by response surface methodology. Journal of Applied Phycology 29: 10271036.CrossRefGoogle Scholar
Silva Benavides, A. M., Torzillo, G., Kopecký, J. et al. (2013). Productivity and biochemical composition of Phaeodactylum tricornutum (Bacillariophyceae) cultures grown outdoors in tubular photobioreactors and open ponds. Biomass and Bioenergy 54: 115122.CrossRefGoogle Scholar
Singh, A. K. & Mallick, N. (2017). Advances in cyanobacterial polyhydroxyalkanoates production. FEMS Microbiology Letters 364: fnx189–fnx189.CrossRefGoogle ScholarPubMed
Singh, A. K., Sharma, L., Mallick, N. et al. (2017). Progress and challenges in producing polyhydroxyalkanoate biopolymers from cyanobacteria. Journal of Applied Phycology 29: 12131232.CrossRefGoogle Scholar
Singh, S., Singh, M. K., Pal, S. K. et al. (2016). Sustainable enhancement in yield and quality of rain-fed maize through Gracilaria edulis and Kappaphycus alvarezii seaweed sap. Journal of Applied Phycology 28: 20992112.CrossRefGoogle Scholar
Skulberg, O. M. (2000). Microalgae as a source of bioactive molecules – experience from cyanophyte research. Journal of Applied Phycology 12: 341348.CrossRefGoogle Scholar
Soeder, C. J., Liersch, R. & Trültzsch, U. (1969). Unterschiedliche Wirkung von Captan auf das Wachstum einiger Stämme von Chlorella und Scenedesmus. Archiv für Mikrobiologie 67: 166172.CrossRefGoogle ScholarPubMed
Stirk, W. A., Tarkowská, D., Turečová, V. et al. (2014). Abscisic acid, gibberellins and brassinosteroids in Kelpak®, a commercial seaweed extract made from Ecklonia maxima. Journal of Applied Phycology 26: 561567.CrossRefGoogle Scholar
Stravs, M. A., Pomati, F. & Hollender, J. (2017). Exploring micropollutant biotransformation in three freshwater phytoplankton species. Environmental Science: Processes and Impacts 19: 822832.Google ScholarPubMed
Suzuki, K. (2017). Large-scale cultivation of Euglena. In: Schwartzbach, S. D. & Shigeoka, S. (eds.) Euglena: Biochemistry, Cell and Molecular Biology. Springer, Cham, pp. 285293.CrossRefGoogle Scholar
Troschl, C., Meixner, K. & Drosg, B. (2017). Cyanobacterial PHA Production—Review of recent advances and a summary of three years’ working experience running a pilot plant. Bioengineering 4: 26.CrossRefGoogle Scholar
Velasquez-Orta, S. B., Curtis, T. P. & Logan, B. E. (2009). Energy from algae using microbial fuel cells. Biotechnology and Bioengineering 103: 10681076.CrossRefGoogle ScholarPubMed
Volkman, J. K. (2016). Sterols in microalgae. In: Borowitzka, M. A, Beardall, J. & Raven, J. A. (eds.) The Physiology of Microalgae. Springer, Dordrecht, pp. 485505.CrossRefGoogle Scholar
Wang, H., Zhang, Y., Chen, L. et al. (2018a). Combined production of fucoxanthin and EPA from two diatom strains Phaeodactylum tricornutum and Cylindrotheca fusiformis cultures. Bioprocess and Biosystems Engineering 41: 10611071.CrossRefGoogle ScholarPubMed
Wang, K., Mandal, A., Ayton, E. et al. (2016). Modification of protein rich algal-biomass to form bioplastics and odor removal. In: Singh Dhillon, G. (ed.) Protein Byproducts. Academic Press, Boston, MA. pp 107117.CrossRefGoogle Scholar
Wang, P., Wong, Y.-S. & Tam, N. F.-Y. (2017). Green microalgae in removal and biotransformation of estradiol and ethinylestradiol. Journal of Applied Phycology 29: 263273.CrossRefGoogle Scholar
Wang, Y., Meng, F., Li, H. et al. (2018b). Biodegradation of phenol by Isochrysis galbana screened from eight species of marine microalgae: Growth kinetic models, enzyme analysis and biodegradation pathway. Journal of Applied Phycology 31: 445455.CrossRefGoogle Scholar
Wijffels, R. H. & Barbosa, M. J. (2010). An outlook on microalgal biofuels. Science 329: 796799.CrossRefGoogle ScholarPubMed
Wite, D., Mattner, S. W., Porter, I. J. et al. (2015). The suppressive effect of a commercial extract from Durvillaea potatorum and Ascophyllum nodosum on infection of broccoli by Plasmodiophora brassicae. Journal of Applied Phycology 27: 21572161.CrossRefGoogle ScholarPubMed
Wynn, J., Behrens, P., Sundararajan, A. et al. (2010). Production of single cell oils from dinoflagellates. In: Cohen, Z. & Ratledge, C. (eds.) Single Cell Oils. Microbial and Algal Oils. AOCS Press, Urbana, pp. 115129.CrossRefGoogle Scholar
Xie, Y., Zhao, X., Chen, J. et al. (2017). Enhancing cell growth and lutein productivity of Desmodesmus sp. F51 by optimal utilization of inorganic carbon sources and ammonium salt. Bioresource Technology 244: 664671.CrossRefGoogle ScholarPubMed
Xiong, J.-Q., Kurade, M. B., Abou-Shanab, R. A. I. et al. (2016). Biodegradation of carbamazepine using freshwater microalgae Chlamydomonas mexicana and Scenedesmus obliquus and the determination of its metabolic fate. Bioresource Technology 205: 183190.CrossRefGoogle ScholarPubMed
Xiong, J.-Q., Kurade, M. B. & Jeon, B.-H. (2018). Can microalgae remove pharmaceutical contaminants from water? Trends in Biotechnology 36: 3044.CrossRefGoogle ScholarPubMed
Xiong, J.-Q., Kurade, M. B., Kim, J. R. et al. (2017). Ciprofloxacin toxicity and its co-metabolic removal by a freshwater microalga Chlamydomonas mexicana. Journal of Hazardous Materials 323: 212219.CrossRefGoogle ScholarPubMed
Yangthong, M., Hutadilok-Towatana, N., Thawonsuwan, J. et al. (2016). An aqueous extract from Sargassum sp. enhances the immune response and resistance against Streptococcus iniae in the Asian sea bass (Lates calcarifer Bloch). Journal of Applied Phycology 28: 35873598.CrossRefGoogle Scholar
Zarekarizi, A., Hoffmann, L. & Burritt, D. (2018). Approaches for the sustainable production of fucoxanthin, a xanthophyll with potential health benefits. Journal of Applied Phycology 31: 281299.CrossRefGoogle Scholar
Zhang, B.-L., Yan, X.-H. & Huang, L.-B. (2011). Evaluation of an improved strain of Porphyra yezoensis Ueda (Bangiales, Rhodophyta) with high-temperature tolerance. Journal of Applied Phycology 23: 841847.CrossRefGoogle Scholar

Save book to Kindle

To save this book to your Kindle, first ensure [email protected] 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.

  • The Future
  • Edited by Mario Giordano, Università degli Studi di Ancona, Italy, John Beardall, Monash University, Victoria, John A. Raven, University of Dundee, Stephen C. Maberly, UK Centre for Ecology & Hydrology, Lancaster
  • Book: Evolutionary Physiology of Algae and Aquatic Plants
  • Online publication: 24 October 2024
  • Chapter DOI: https://doi.org/10.1017/9781139049979.018
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.

  • The Future
  • Edited by Mario Giordano, Università degli Studi di Ancona, Italy, John Beardall, Monash University, Victoria, John A. Raven, University of Dundee, Stephen C. Maberly, UK Centre for Ecology & Hydrology, Lancaster
  • Book: Evolutionary Physiology of Algae and Aquatic Plants
  • Online publication: 24 October 2024
  • Chapter DOI: https://doi.org/10.1017/9781139049979.018
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.

  • The Future
  • Edited by Mario Giordano, Università degli Studi di Ancona, Italy, John Beardall, Monash University, Victoria, John A. Raven, University of Dundee, Stephen C. Maberly, UK Centre for Ecology & Hydrology, Lancaster
  • Book: Evolutionary Physiology of Algae and Aquatic Plants
  • Online publication: 24 October 2024
  • Chapter DOI: https://doi.org/10.1017/9781139049979.018
Available formats
×