Skip to main content Accessibility help
×
Hostname: page-component-78c5997874-lj6df Total loading time: 0 Render date: 2024-10-28T06:34:30.406Z Has data issue: false hasContentIssue false

15 - Aquatic Phototrophs and the Greenhouse Effect

from 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

Summary

Increases in atmospheric CO2 expected over the next century will cause further global warming and further increases in the CO2 concentration in water bodies and, by equilibration of CO2 with HCO3 - CO32− - H+, increased HCO3 and H+ and decreased CO32−. Warming increases stratification and shoaling of the thermocline; this decreases the supply of nutrients regenerated in deep waters to the upper mixed layer across the thermocline, and increases mean photosynthetically active and UV radiation in the upper mixed layer. Taken separately, these changes can have profound changes on the performance of algae and, because of differences among taxa, in the species composition of primary producer populations. However, it is becoming increasingly clear that the effects of individual components of global change cannot be used as useful predictors of what will happen to aquatic ecosystems into the future and that studies need to take more cognisance of the interactive effects between such factors. There is evidence for genetic adaptation, as well as phenotypic acclimation, in algae exposed to either elevated CO2 or increased temperature. Our understanding of the effects on global change requires further studies into the genetic and acclimation responses of algae exposed to combinations of changed environmental factors.

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

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

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.

Available formats
×

Save book to Dropbox

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

Available formats
×

Save book to Google Drive

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

Available formats
×