Skip to main content Accessibility help
×
Hostname: page-component-cd9895bd7-p9bg8 Total loading time: 0 Render date: 2024-12-25T19:39:30.654Z Has data issue: false hasContentIssue false

Part II - Physiology of Photosynthetic Autotrophs in Present-Day Environments

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

Agostoni, M., Lucker, B. F., Smith, M. A. Y. et al. (2016). Competition-based phenotyping reveals a fitness cost for maintaining phycobilisomes under fluctuating light in the cyanobacterium Fremyella diplosiphon. Algal Research 15: 110119.CrossRefGoogle Scholar
Allorent, G., Tokutsu, R., Roach, T. et al. (2013). A dual strategy to cope with high light in Chlamydomonas reinhardtii. Plant Cell 25: 545557.CrossRefGoogle ScholarPubMed
Asada, K. (2000). The water-water cycle as alternative photon and electron sinks. Philosophical Transactions of the Royal Society B 355: 14191431.CrossRefGoogle ScholarPubMed
Bailleul, B., Rogato, A., de Martino, A. et al. (2010). An atypical member of the light-harvesting complex stress-related protein family modulates diatom responses to light. Proceedings of the National Academy of Sciences USA 107: 1821418219.CrossRefGoogle ScholarPubMed
Bao, H., Melnicki, M. R. & Kerfeld, C. A. (2017). Structure and functions of Orange Carotenoid protein homologs in cyanobacteria. Current Opinion Plant Biology 37: 19.CrossRefGoogle ScholarPubMed
Berner, T., Dubinsky, Z., Wyman, K. et al. (1989). Photoadaptation and the ‘package’ effect in Dunaliella tertiolecta (Chlorophyceae). Journal of Phycology 25: 7078.CrossRefGoogle Scholar
Bibby, T. S., Nield, J. & Barber, J. (2001). Iron deficiency induces the formation of an antenna ring around trimeric photosystem I in cyanobacteria. Nature 412: 743745.CrossRefGoogle ScholarPubMed
Bienfang, J. P. K., Szyper, J. R, Okamoto, M. Y. et al. (1984). Temporal and spatial variability of phytoplankton in a subtropical ecosystem. Limnology and Oceanography 29: 527539.CrossRefGoogle Scholar
Bína, D., Gardian, Z., Herbstová, M. et al. (2014). Novel type of red-shifted chlorophyll a antenna complex from Chromera velia: II. Biochemistry and spectroscopy. Biochimica et Biophysica Acta 1837: 802810.CrossRefGoogle ScholarPubMed
Bonente, G., Ballottari, M., Truong, T. B. et al. (2011). Analysis of LhcSR3, a protein essential for feedback de-excitation in the green alga Chlamydomonas reinhardtii. PLOS Biology 9: e1000577.CrossRefGoogle ScholarPubMed
Brawley, S. H., Blouin, N. A., Ficko-Blean, E. et al. (2017). Insights into the red algae and eukaryotes evolution from the genome of Porphyra umbilicalis (Bangiophyceae, Rhodophyta). Proceedings of the National Academy of Sciences USA 114: E6361E6370.CrossRefGoogle ScholarPubMed
Büchel, C. (2020). Light harvesting complexes in chlorophyll c-containing algae. Biochimica Biophysica Acta – Bioenergetics 1861: 148027.CrossRefGoogle ScholarPubMed
Buck, J. M., Sherman, J., Bártulos, C. R. et al. (2019). Lhcx proteins provide photoprotection via thermal dissipation of absorbed light in the diatom Phaeodactylum tricornutum. Nature Communications 10: 4167. https://doi.org/10.1038/s41467-019-12043-6.CrossRefGoogle ScholarPubMed
Calzadilla, P. I. & Kirilovsky, D. (2020). Revisiting cyanobacterial state transitions. Photochemical and Photobiological Science 19: 585603.CrossRefGoogle ScholarPubMed
Chen, M., Telfer, A., Lin, S. et al. (2005). The nature of the photosystem II reaction centre in the chlorophyll d-containing prokaryote Acaryochloris marina. Photochemical and Photobiological Science 4: 10601064.CrossRefGoogle ScholarPubMed
Chen, M., Li, Y. Q., Birch, D. et al. (2012). A cyanobacterium that contains chlorophyll f – a red-absorbing photopigment. FEBS Letters 586: 32493254.CrossRefGoogle ScholarPubMed
Christie, K. M., Salomon, M., Nozaki, K. et al. (1999). LOV (light, oxygen or voltage) domain of the blue light photoreceptor phototropin (nph1): Binding sites for the chromophore flavin mononucleotide. Proceedings of the National Academy of Sciences USA 96: 87799783.CrossRefGoogle ScholarPubMed
Clegg, M. R., Maberly, S. C. & Jones, R. I. (2004). Dominance and compromise in freshwater phytoplanktonic flagellates: The interaction of behavioural preferences for conflicting environmental gradients. Functional Ecology 18: 371380.CrossRefGoogle Scholar
Clegg, M. R., Maberly, S. C. & Jones, R. I. (2007). Behavioural response as a predictor of seasonal depth distribution and vertical niche separation in freshwater phytoplanktonic flagellates. Limnology and Oceanography 52: 441455.CrossRefGoogle Scholar
Coesel, S. N., Durham, B. P., Groussman, R. B. et al. (2021). Diel transcriptional oscillations of light-sensitive regulating elements in open-ocean eukaryotic plankton communities. Proceedings of the National Academy of Sciences USA 118: e8201038118.CrossRefGoogle Scholar
Costa, B., Sachse, M., Jugandras, A. et al. (2013). Aureochome 1a is involved in photoacclimation in the diatom Phaeodactylum tricornutom. PLOS ONE 8: e74451.Google Scholar
Depauw, F. A., Rogato, A., D’Alcalà, M. R. et al. (2012). Exploring the molecular basis of responses to light in marine diatoms. Journal of Experimental Botany 63: 15751591.CrossRefGoogle ScholarPubMed
Depège, N., Bellafiore, S. & Rochaix, J.-D. (2003). Role of chloroplast protein kinase Stt7 in LHCII phosphorylation and state transition in Chlamydomonas. Science 299: 15721575.CrossRefGoogle ScholarPubMed
Dominguez-Martin, M. A., Sauer, P. V., Kirst, H. et al. (2022). Structures of a phycobilisome in light-harvesting and photoprotected states. Nature 609: 835845.CrossRefGoogle ScholarPubMed
Dring, M. J. (1967). Phytochrome in red alga, Porphyra tenera. Nature 215: 14111412.CrossRefGoogle Scholar
Dring, M. J. (1988). Photocontrol of development in algae. Annual Review of Plant Biology and Plant Molecular Biology 39: 57174.Google Scholar
Emido Fortunato, A., Jaubert, M., Enomoto, G. et al. (2016). Diatom phytochromes reveal the existence of far-red light-based sensing in the ocean. The Plant Cell 28: 616628.CrossRefGoogle ScholarPubMed
Engelken, J., Brinkmann, H. & Adamska, I. (2010). Taxonomic distribution and origins of the extended LHC (light-harvesting complex) antenna protein superfamily. BMC Evolutionary Biology 10: 233.CrossRefGoogle ScholarPubMed
Erickson, E., Wakao, S. & Niyogi, K. K. (2015). Light stress and photoprotection in Chlamydomonas reinhardtii. Plant Journal 82: 449465.CrossRefGoogle ScholarPubMed
Falkowski, P. G. & Owens, T. G. (1980). Light-shade adaptation. Two strategies in marine phytoplankton. Plant Physiology 66: 592595.CrossRefGoogle ScholarPubMed
Feulner, G. (2012). The faint young Sun problem. Reviews of Geophysics 50: 2011RG000375.CrossRefGoogle Scholar
Finkel, Z. V. & Irwin, A. J. (2000). Modeling size-dependent photosynthesis: Light absorption and the allometric rule. Journal of Theoretical Biology 204: 361369.CrossRefGoogle ScholarPubMed
Finkel, Z. V. (2001). Light absorption and size scaling of light-limited metabolism in marine diatoms. Limnology and Oceanography 46: 8694.CrossRefGoogle Scholar
Fisher, T., Shurtz-Swirski, R., Gepstein, S. et al. (1989). Changes in the levels of ribulose-l,5-bisphosphate carboxylase/oxygenase (Rubisco) in Tetraedron minimum (Chlorophyta) during light and shade adaptation. Plant and Cell Physiology 30: 221228.CrossRefGoogle Scholar
Fu, G., Nagasato, C., Yamagichi, T. et al. (2016). Ubiquitous distribution of helmchrome in phototactic swarmers of the stramenopiles. Protoplasma 253: 929941.CrossRefGoogle ScholarPubMed
Fujita, Y. & Ohki, K. (2004). On the 710 nm fluorescence emitted by the diatom Phaeodactylum tricornutum at room temperature. Plant and Cell Physiology 45: 392397.CrossRefGoogle ScholarPubMed
Galvao, V. C. & Fankhauser, C. (2015). Sensing the light environment in plants: Photoreceptors and early signaling steps. Current Opinion in Neurobiology 34: 4653.CrossRefGoogle ScholarPubMed
Gantt, E. and CunninghamJr, F. X. (2001). Algal pigments. eLS. https://doi.org/10.1038/npg.els.0000323.Google Scholar
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
Geider, R. & Osborne, B. (1987). Light absorption by a marine diatom: Experimental observations and theoretical calculations of the package effect in a small Thalassiosira species. Marine Biology 96: 299308.CrossRefGoogle Scholar
Geider, R. J., Platt, T. & Raven, J. A. (1986) Size dependence of growth and photosynthesis in diatoms – a synthesis. Marine Ecology Progress Series 30: 93104.CrossRefGoogle Scholar
Girolomoni, L., Cazzaniga, S., Pinnola, A. et al. (2019). LHCSR3 is a nonphotochemical quencher of both photosystems in Chlamydomonas reinhardtii. Proceedings of the National Academy of Sciences USA 116: 42124217.CrossRefGoogle ScholarPubMed
Gómez-Consarnau, L., Raven, J. A., Levine, N. M. et al. (2019). Microbial rhodopsins are major contributors to the solar energy captured in the sea. Science Advances 5: eaaw8855.CrossRefGoogle Scholar
Goss, R. & Lepetit, B. (2015). Biodiversity of NPQ. Journal of Plant Physiology 172: 1332.CrossRefGoogle ScholarPubMed
Grebert, T., Dore, H., Partensky, F. et al. (2018). Light color acclimation is a key process in the global ocean distribution of Synechococcus cyanobacteria. Proceedings of the National Academy of Sciences USA 115: E2010E2019.CrossRefGoogle ScholarPubMed
Häder, D.-P. & Iseki, M. (2017). Photomovement in Euglena. In: Schwartzbach, C. & Shigeoka, S. (eds.) Euglena: Biochemistry, Cell and Molecular Biology. Springer, Cham, pp. 207335.CrossRefGoogle Scholar
Herbstová, M., Bína, D., Koník, P. et al. (2015). Molecular basis of chromatic adaptation in pennate diatom Phaeodactylum tricornutum. Biochimica et Biophysica Acta – Bioenergetics 1847: 534543.CrossRefGoogle ScholarPubMed
Herbstová, M., Bina, D., Kana, R. et al. (2017). Red-light phenotype in a marine diatom involves a specialized oligomeric red-shifted antenna and altered cell morphology. Scientific Reports 7: 11976.CrossRefGoogle Scholar
Hieronymi, M. & Macke, A. (2010). Spatiotemporal underwater light field fluctuations in the open ocean. Journal of the European Optical Society: Rapid Publications 5: 19902573.CrossRefGoogle Scholar
Hu, Q., Mayashita., H., Iwakai, I. et al. (1998). A photosystem I reaction center driven by chlorophyll d in oxygenic photosynthesis. Proceedings of the National Academy of Sciences USA 95: 1331913323.CrossRefGoogle ScholarPubMed
Huysman, M. J. J., Fortunato, A. E., Matthijs, M. et al. (2013). Aurochrome 1s-mediated induction in the diatom-specific cyclin dscyc2 controls the onset of cell division in diatom Phaeodactylum tricornutum. The Plant Cell 25: 215228.CrossRefGoogle Scholar
Ito, H. & Tanaka, A. (2011). Evolution of a divinyl chlorophyll-based photosystem in Prochlorococcus. Proceedings of the National Academy of Sciences USA 108: 1801418019.CrossRefGoogle ScholarPubMed
Jékely, G. (2009). Evolution of phototaxis. Philosophical Transactions of the Royal Society B 364: 27952808.CrossRefGoogle ScholarPubMed
Johnsen, G., Nelson, N. B., Jovine, R. V. M. et al. (1994). Chromoprotein and pigment-dependent modelling of spectral light absorption in two dinoflagellates, Prorocentrum minimum and Hetercapsa pygmaea. Marine Ecology Progress Series 114: 245258.CrossRefGoogle Scholar
Kain, J. A. (1989). The seasons in the subtidal. British Phycological Journal 24: 203215.CrossRefGoogle Scholar
Kanazawa, A., Neofotis, P., Davis, G. A. et al. (2020). Diversity in photoprotection and energy balancing in terrestrial and aquatic phototrophs. In: Larkum, A. W. D., Grossman, A. & Raven, J. A. (eds.) Photosynthesis in Algae: Biochemical and Physiological Mechanisms. Advances in Photosynthesis and Respiration 45. Springer, Cham, pp. 299327.CrossRefGoogle Scholar
Kehoe, D. M. & Gatu, A. (2006). Responding to color: The regulation of complementary chromatic adaptation. Annual Review of Plant Biology 57: 127150.CrossRefGoogle ScholarPubMed
Kianiannoment, A. & Hallman, A. (2014). Algal photoreceptors: In vivo functions and potential applications. Planta 239: 126.CrossRefGoogle Scholar
Kirk, J. T. (1994). Light and Photosynthesis in Aquatic Ecosystems. Cambridge: Cambridge University PressCrossRefGoogle Scholar
Koehne, B., Elli, G., Jennings, R. C. et al. (1999). Spectroscopic and molecular characterization of a long wavelength absorbing antenna of Ostreobium sp. Biochimica et Biophysica Acta-Bioenergetics 1412: 94107.CrossRefGoogle ScholarPubMed
Kottke, T., Oldemeyer, S., Wenzel, S. et al. (2017). Cryptochrome photoreceptors in green algae: Unexpected versatility of mechanisms and functions. Journal of Plant Physiology 217: 412.CrossRefGoogle ScholarPubMed
Koziol, A. G., Borza, T., Ishida, K.-I. et al. (2007). Tracing the evolution of the light-harvesting antennae in chlorophyll a/b-containing organisms. Plant Physiology 143: 18021816.CrossRefGoogle ScholarPubMed
Larkum, A. W. D. (2020). Light-harvesting in cyanobacteria and eukaryotic algae: An Overview. In: Larkum, A. W. D., Grossman, A. and Raven, J. A. (eds.) Photosynthesis in Algae: Biochemical and Physiological Mechanisms, Advances in Photosynthesis and Respiration 45. Springer, Cham, pp. 207260. https://doi.org/10.1007/978–3-030–33397–3_10.CrossRefGoogle Scholar
Lepetit, B., Gelin, G., Lepetit, M. et al. (2017). The diatom Phaeodactylum tricornutum adjusts nonphotochemical fluorescence quenching capacity in response to dynamic light via fine-tuned Lhcx and xanthophyll cycle pigment synthesis. New Phytologist 214: 205218.CrossRefGoogle ScholarPubMed
Lepetit, B., Sturm, S., Rogato, A. et al. (2013). High light acclimation in the secondary plastids containing diatom Phaeodactylum tricornutum is triggered by the redox state of the plastoquinone pool. Plant Physiology 161: 853865.CrossRefGoogle ScholarPubMed
Li, F. W., Rothfels, C. J., Melkonian, M. et al. (2015). The origin and evolution of phototropins. Frontiers in Plant Science 56: article 697.Google Scholar
Litchman, E. (2000). Growth rates of phytoplankton under fluctuating light. Freshwater Biology 44: 223235.CrossRefGoogle Scholar
Litvín, R., Bína, D., Herbstová, M. et al. (2019). Red-shifted light-harvesting system of freshwater eukaryotic alga Trachydiscus minutus (Eustigmatophyta, Stramenopila). Photosynthesis Research 142: 137151.CrossRefGoogle ScholarPubMed
Malerba, M. E., Palacios, M. M., Palacios Delgado, Y. M. et al. (2018). Cell size, photosynthesis and the package effect: An artificial selection approach. New Phytologist 219: 449461.CrossRefGoogle ScholarPubMed
Mann, M., Serif, M., Wrobel, T. et al. (2020). The aureochrome photoreceptor PtAUREO1a is a highly effective blue light switch in diatoms. iScience 23: 101730.CrossRefGoogle ScholarPubMed
Mao, Z., Stuart, V., Pan, D. et al. (2010). Effects of phytoplankton species competition on absorption spectra and modelled hyperspectral reflectance. Ecological Informatics 5: 359366.CrossRefGoogle Scholar
Mengelt, C. & Brezelin, 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
Miyashita, H., Ikemoto, H., Kurano, N. et al. (1996). Chlorophyll d as a major pigment. Nature 383: 402402.CrossRefGoogle Scholar
Moejes, F. W., Matuszynska, A., Adhikari, K. et al. (2017). A systems-wide understanding of photosynthetic acclimation in algae and higher plants. Journal of Experimental Botany. 68: 26672681.CrossRefGoogle ScholarPubMed
Moisan, T. A. & Mitchell, B. G. (1999). Photophysical acclimation of Phaeocystis antarctica. Limnology and Oceanography 44: 247258.CrossRefGoogle Scholar
Morel, A. & Bricaud, A. (1981). Theoretical results concerning light absorption in a discrete medium, and application to specific absorption of phytoplankton. Deep Sea Research Part A, Oceanographic Research Papers 28: 13751393.CrossRefGoogle Scholar
Morel, A., Gentili, B., Claustre, H. et al. (2007). Optical properties of the ‘clearest’ natural waters. Limnology and Oceanography 52: 217229.CrossRefGoogle Scholar
Muzzopappa, F. & Kirilovsky, D. (2020). Changing color for photoprotection: The orange carotenoid protein. Trends in Plant Science 25: 92104.CrossRefGoogle ScholarPubMed
Nagy, G., Ünnep, R., Zsiros, O. et al. (2014). Chloroplast remodeling during state transitions in Chlamydomonas reinhardtii as revealed by noninvasive techniques in vivo. Proceedings of the National Academy of Sciences USA 111: 50425047.CrossRefGoogle ScholarPubMed
Neilson, J. A. D. & Durnford, D. G. (2010). Structural and functional diversification of the light-harvesting complexes in photosynthetic eukaryotes. Photosynthesis Research 106: 5771.CrossRefGoogle ScholarPubMed
Nelson, N. B., Prézelin, B. B. & Bidigare, R. R. (1993). Phytoplankton light absorption and the package effect in California coastal waters. Marine Ecology Progress Series 94: 217227.CrossRefGoogle Scholar
Nikkanen, L., Solymosi, D., Jokel, M. et al. (2021). Regulatory electron transport pathways of photosynthesis in cyanobacteria and microalgae: Recent advances and biotechnological prospects. Physiologia Plantarum 173: 514525.CrossRefGoogle ScholarPubMed
Norici, A., Gerotto, C., Beardall, J. & Raven, J. A. (2022). Environmental variability and its control of productivity. In: Maberly, S. C. & Gontero, B. (eds.) Blue Planet, Red and Green Photosynthesis. ISTE-Wiley, London, UK, pp. 225272.CrossRefGoogle Scholar
Peers, G., Truong, T. B., Ostendorf, E. et al. (2009). An ancient light-harvesting protein is critical for the regulation of algal photosynthesis. Nature 462: 518521.CrossRefGoogle ScholarPubMed
Pfannschmidt, T. (2003). Chloroplast redox signals: How photosynthesis controls its own genes. Trends in Plant Science 8: 3341.CrossRefGoogle ScholarPubMed
Quigg, A., Kevekordes, K., Raven, J. A. et al. (2006). Limitations on microalgal growth at very low photon fluence rates: The role of energy slippage. Photosynthesis Research 88: 299310.CrossRefGoogle ScholarPubMed
Ragni, M. & Ribera d’Alcalà, M. (2004). Light as an information carrier underwater. Journal of Plankton Research 26: 433443.CrossRefGoogle Scholar
Raven, J. A. & Geider, R. J. (2003). Adaptation, acclimation and regulation in algal photosynthesis. In: Larkum, A. W. D., Douglas, S. E. and Raven, J. A. (eds.) Photosynthesis in Algae. Kluwer Academic Publishers, Dordrecht, The Netherlands, pp. 385412.CrossRefGoogle ScholarPubMed
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. & Quigg, A. (2020). Light-driven oxygen consumption in the water-water cycles and photorespiration, and light stimulated mitochondrial respiration. In: Larkum, A. W. D., Grossman, A. R. & Raven, J. A. (eds.) Photosynthesis in Algae: Biochemical and Physiological Mechanisms, Advances in Photosynthesis and Respiration, Vol. 45. Springer, Cham, pp. 161168.CrossRefGoogle Scholar
Raven, J. A., Evans, M. C. W. & Korb, R. E. (1999). The role of trace metals in photosynthetic electron transport in O2-evolving organisms. Photosynthesis Research 60: 11149.CrossRefGoogle Scholar
Raven, J. A. (2013). Iron acquisition and allocation in stramenopile algae. Journal of Experimental Botany 64: 21192127.CrossRefGoogle ScholarPubMed
Raven, J. A. (2020). Chloride involvement in the synthesis, functioning and repair of the photosynthetic apparatus in vivo. New Phytologist 227: 334342.CrossRefGoogle ScholarPubMed
Raven, J. A., Kubler, J. E., & Beardall, J. (2000). Put out the light, and then put out the light. Journal of the Marine Biological Association of the United Kingdom 80: 125.CrossRefGoogle Scholar
Richardson, K., Beardall, J. & Raven, J. (1983). Adaptation of unicellular algae to irradiance: An analysis of strategies. New Phytologist 93: 157191.CrossRefGoogle Scholar
Roberts, E. M., Bowers, D. G. & Davies, A. J. (2017). Tidal modulation of seabed light and its implications for benthic algae. Limnology and Oceanography 63: 91106.CrossRefGoogle Scholar
Rockwell, N. C. & Lagarios, J. C. (2019). Phytochrome evolution in 3D: Deletion, duplication, and diversification. New Phytologist 228: 22832300.Google Scholar
Rockwell, N. C., Lagarios, J. E. & Bhattacharya, D. (2014a). Primary endosymbiosis and the evolution of light and oxygen sensing in photosynthetic eukaryotes. Frontiers in Ecology and Evolution 2: article 66.CrossRefGoogle ScholarPubMed
Rockwell, N. C., Duanmu, D., Martin, S. S. et al. (2014b). Eukaryotic algal phytochromes span the visible spectrum. Proceedings of the National Academy of Sciences USA 111: 38713876.CrossRefGoogle ScholarPubMed
Ruban, A. V. (2015). Evolution under the sun: Optimizing light harvesting in photosynthesis. Journal of Experimental Botany 66: 723.CrossRefGoogle ScholarPubMed
Schiphorst, C. & Bassi, R. (2020). Chlorophyll-xanthophyll antenna complexes: In between light harvesting and energy dissipation. In: Larkum, A. W. D., Grossman., A. R. & Raven, J. A. (eds.) Photosynthesis in Algae: Biochemical and Physiological Mechanisms. Springer International Publishing, Cham, pp 2755.CrossRefGoogle Scholar
Schubert, H., Sagert, S. & Forster, R. M. (2001). Evaluation of the different levels of variability in the underwater light field of a shallow estuary. Helgoland Marine Research 55: 1222.CrossRefGoogle Scholar
Simionato, D., Sforza, E., Corteggiani Carpinelli, E. et al. (2011). Acclimation of Nannochloropsis gaditana to different illumination regimes: Effects on lipids accumulation. Bioresource Technology 102: 60266032.CrossRefGoogle ScholarPubMed
Slater, B., Kosmützky, D., Nisbet, R. E. R. et al. (2021). The evolution of the cytochrome c6 family of photosynthetic electron transfer proteins. Genome Biology and Evolution 13: evab146. https://doi.org/10.1093/gbe/evab146.CrossRefGoogle ScholarPubMed
Stomp, M., Huisman, J., De Jongh, F. et al. (2004). Adaptive divergence in pigment composition promotes phytoplankton biodiversity. Nature 432: 104107.CrossRefGoogle ScholarPubMed
Stomp, M., Huisman, J., Stal, L. J. et al. (2007). Colorful niches of phototrophic microorganisms shaped by vibrations of the water molecule. ISME Journal 1: 271282.CrossRefGoogle ScholarPubMed
Sukenik, A., Livne, A., Apt, K. E. et al. (2000). Characterization of a gene encoding the light-harvesting violaxanthin-chlorophyll protein of Nannochloropsis sp. (Eustigmatophyceae). Journal of Phycology 36: 563570.CrossRefGoogle ScholarPubMed
Takahashi, F., Yamagata, D., Ishikawa, M. et al. (2007). Aureochrome, a photoreceptor required for photomorphogenesis in stramenopiles. Proceedings of the National Academy of Sciences USA 104: 1962519630.CrossRefGoogle ScholarPubMed
Takahashi, F. (2016). Blue-light-regulated transcription factor, Aureochrome, in photosynthetic stramenopiles. Journal of Plant Research 129: 189197.CrossRefGoogle ScholarPubMed
Takizawa, K., Cruz, J. A., Kanazawa, A. et al. (2007). The thylakoid proton motive force in vivo. Quantitative, non-invasive probes, energetics, and regulatory consequences of light-induced pmf. Biochimica et Biophysica Acta-Bioenergetics 1767: 12331244.CrossRefGoogle ScholarPubMed
Ünlü, C., Drop, B., Croce, R. et al. (2014). State transitions in Chlamydomonas reinhardtii strongly modulate the functional size of photosystem II but not of photosystem I. Proceedings of the National Academy of Sciences USA 111: 34603465.CrossRefGoogle Scholar
Vant, W. N. (1990). Causes of light attenuation in nine New Zealand estuaries. Estuarine, Coastal and Shelf Science. 31: 125137.CrossRefGoogle Scholar
Virtanen, O., Khorobrykh, S. & Tyystjärvi, E. (2021). Acclimation of Chlamydomonas reinhardtii to extremely strong light. Photosynthesis Research 147: 91106.CrossRefGoogle ScholarPubMed
Wagner, H., Jakob, T. & Wilhelm, C. (2006). Balancing the energy flow from captured light to biomass under fluctuating light conditions. New Phytologist 169: 95108.CrossRefGoogle ScholarPubMed
Wilhelm, C. & Jakob, T. (2006). Uphill energy transfer from long-wavelength absorbing chlorophylls to PS II in Ostreobium sp. is functional in carbon assimilation. Photosynthesis Research 87: 323329.CrossRefGoogle Scholar
Wilhelm, C., Jungandreas, A., Jakob, T. et al. (2014). Light acclimation in diatoms: From phenomenology to mechanisms. Marine Genomics 16: 515.CrossRefGoogle ScholarPubMed
Willows, R. D. (2020). Biosynthesis of chlorophyll and bilins in algae. In: Larkum, A. W. D., Grossman, A. and Raven, J. A. (eds.) Photosynthesis in Algae: Biochemical and Physiological Mechanisms, Advances in Photosynthesis and Respiration 45. Springer, Cham, pp. 83103.CrossRefGoogle Scholar
Wolf, B. M., Niedzwiedzki, D. M., Magdaong, N. C. M. et al. (2018). Characterization of a newly isolated freshwater Eustigmatophyte alga capable of utilizing far-red light as its sole light source. Photosynthesis Research 135: 177189.CrossRefGoogle ScholarPubMed
Yang, Y., Lam, V., Adomako, M. et al. (2018). Phototaxis in a wild isolate of the cyanobacterium Synechococcus elongatus. Proceedings of the National Academy of Sciences USA 115: E12878E12387.CrossRefGoogle Scholar
Zhu, S.-H. & Green, B. R. (2010). Photoprotection in the diatom Thalassiosira pseudonana: Role of LI818-like proteins in response to high light stress. Biochimica et Biophysica Acta- Bioenergetics 1797: 14491457.CrossRefGoogle ScholarPubMed

References

Allakhverdiev, S. I., Los, D. A., Mohanty, P. et al. (2007). Glycinebetaine alleviates the inhibitory effect of moderate heat stress on the repair of photosystem II during photoinhibition. Biochimica Biophysica Acta BBA – Bioenergetics 1767: 13631371.CrossRefGoogle ScholarPubMed
Alric, J. & Johnson, X. (2017). Alternative electron transport pathways in photosynthesis: A confluence of regulation. Current Opinions in Plant Biology 37: 7886.CrossRefGoogle ScholarPubMed
Andersson, A., Haecky, P. & Hagström, Å. (1994). Effect of temperature and light on the growth of micro- nano- and pico-plankton: Impact on algal succession. Marine Biology 120: 511520.CrossRefGoogle Scholar
Asada, K. (2000). The water-water cycle as alternative photon and electron sinks. Philosophical Transactions of the Royal Society B 355: 14191431.CrossRefGoogle ScholarPubMed
Banse, K. (1976). Rates of growth, respiration and photosynthesis of unicellular algae as related to cell size – a review. Journal of Phycology 12: 135140.Google Scholar
Beauchemin, M., Roy, S., Pelletier, S. et al. (2016). Characterization of two dinoflagellate cold shock domain proteins. Msphere 1: e00034–15. https://doi.org/10.1128/mSphere.00034–15.CrossRefGoogle ScholarPubMed
Bělehrádek, J. (1926). Influence of temperature on biological processes. Nature 118: 117118.CrossRefGoogle Scholar
Bieri, P., Leibundgut, M., Saurer, M. et al. (2017). The complete structure of the chloroplast 70S ribosome in complex with translation factor pY. EMBO Journal 36: 475486.CrossRefGoogle ScholarPubMed
Bligh, J. & Johnson, K. G. (1973). Glossary of terms for thermal physiology. Journal of Applied Physiology 35: 941961.CrossRefGoogle ScholarPubMed
Bozzato, D., Jakob, T. & Wilhelm, C. (2019). Effects of temperature and salinity on respiratory losses and the ratio of photosynthesis to respiration in representative Antarctic phytoplankton species. PLOS ONE 14: e0224101. https://doi.org/10.1371/journal.pone.0224101.CrossRefGoogle ScholarPubMed
Butterwick, C., Heaney, S. I. & Talling, J. F. (2005). Diversity in the influence of temperature on the growth rates of freshwater algae, and its ecological relevance. Freshwater Biology 50: 291300.CrossRefGoogle Scholar
Chandler, J. W., Lin, Y., Gainer, P. J. et al. (2016). Variable but persistent coexistence of Prochlorococcus ecotypes along temperature gradients in the ocean’s surface mixed layer. Environmental Microbiology Reports 8: 272284.CrossRefGoogle ScholarPubMed
Chaux, F., Peltier, G. & Johnson, X. (2015). A security network in PSI photoprotection: Regulation of photosynthetic control, NPQ and O2 photoreduction by cyclic electron flow. Frontiers in Plant Science 6: 875. https://doi.org/10.3389/fpls.2015.00875.CrossRefGoogle ScholarPubMed
Cvetkovska, M., Hüner, N. P. A. & Smith, D. R. (2017). Chilling out: The evolution and diversification of psychrophilic algae with a focus on Chlamydomonadales. Polar Biology 40: 11691184.CrossRefGoogle Scholar
D’Amico, S., Collins, T., Marx, J.-C. et al. (2006). Psychrophilic microorganisms: Challenges for life. EMBO Reports. 7: 385389.CrossRefGoogle ScholarPubMed
Day, J. G., Benson, E. E., Harding, K. et al. (2005). Cryopreservation and conservation of microalgae: The development of a pan-European scientific and biotechnological resource (the COBRA project). Cryoletters 26: 231238.Google ScholarPubMed
Devos, N., Ingouff, M., Loppes, R. et al. (1998). Rubisco adaptation to low temperatures: A comparative study in psychrophilic and mesophilic unicellular algae. Journal of Phycology 34: 655660.CrossRefGoogle Scholar
Dolhi, J. M., Maxwell, D. P. & Morgan-Kiss, R. M. (2013). Review: The Antarctic Chlamydomonas raudensis: An emerging model for cold adaptation of photosynthesis. Extremophiles 17: 711722.CrossRefGoogle ScholarPubMed
Fanesi, A., Wagner, H. & Wilhelm, C. (2016). Temperature affects the partitioning of absorbed light energy in freshwater phytoplankton. Freshwater Biology 61: 13651378.CrossRefGoogle Scholar
Fanesi, A., Wagner, H. & Wilhelm, C. (2017). Phytoplankton growth rate modelling: Can spectroscopic cell chemotyping be superior to physiological predictors? Proceedings of the Royal Society of London B 284: 20161956. https://doi.org/10.1098/rspb.2016.1956.Google ScholarPubMed
Galmés., J., Capó-Bauçà, S., Niinmets, Ü. et al. (2019). Potential improvement of photosynthetic CO2 assimilation in crops with the natural variation in temperature response of Rubisco catalytic traits. Current Opinion in Plant Biology 49: 6067.CrossRefGoogle ScholarPubMed
Gao, G., Shi, Q., Xu, Z. et al. (2018). Global warming interacts with ocean acidification to alter PSII function and protection in the diatom Thalassiosira weissflogii. Environmental and Experimental Botany 147: 95103.CrossRefGoogle Scholar
Gierz, S. L., Forêt, S. & Leggat, W. (2017). Transcriptomic analysis of thermally stressed Symbiodinium reveals differential expression of stress and metabolism genes. Frontiers in Plant Science 8: 271. https://doi.org/10.3389/fpls.2017.00271.CrossRefGoogle ScholarPubMed
Gómez, I., Wulff, A., Roleda, M. Y. et al. (2009). Light and temperature demands of marine benthic microalgae and seaweeds in polar regions. Botanica Marina 52: 593608.CrossRefGoogle Scholar
Gong, L., Gao, M., Zang, X. et al. (2015). Peptidase: A novel member of a calmodulin-binding protein of Gracilaria lemaneiformis under heat shock. Journal of Applied Phycology 27: 563570.CrossRefGoogle Scholar
Goss, R. & Jakob, T. (2010). Regulation and function of xanthophyll cycle-dependent photoprotection in algae. Photosynthesis Research 106: 103122.CrossRefGoogle ScholarPubMed
Goss, R. & Lepetit, B. (2015). Biodiversity of NPQ. Journal of Plant Physiology 172: 1332. https://doi.org/10.1016/j.jplph.2014.03.004.CrossRefGoogle ScholarPubMed
Grouneva, I., Gollan, P. J., Kangasjärvi, S. et al. (2013). Phylogenetic viewpoints on regulation of light harvesting and electron transport in eukaryotic photosynthetic organisms. Planta 237: 399412.CrossRefGoogle ScholarPubMed
Han, J. W. & Kim, G. H. (2013). An ELIP-like gene in the freshwater green alga, Spirogyra varians (Zygnematales), is regulated by cold stress and CO2 influx. Journal of Applied Phycology 25: 12971307.CrossRefGoogle Scholar
Hemme, D., Veyel, D., Mühlhaus, T. et al. (2014). Systems-wide analysis of acclimation responses to long-term heat stress and recovery in the photosynthetic model organism Chlamydomonas reinhardtii. Plant Cell 26: 42704297.CrossRefGoogle ScholarPubMed
Hoegh-Guldberg, O. (1999). Climate change, coral bleaching and the future of the world’s coral reefs. Marine and Freshwater Research 50: 839886.Google Scholar
Huner, N. P. A., Öquist, G. & Sarhan, F. (1998). Energy balance and acclimation to light and cold. Trends in Plant Science 3: 224230.CrossRefGoogle Scholar
Kremer, C. T., Thomas, M. K. & Litchman, E. (2017). Temperature- and size-scaling of phytoplankton population growth rates: Reconciling the Eppley curve and the metabolic theory of ecology. Limnology and Oceanography 62: 16581670.CrossRefGoogle Scholar
Kotajima, T., Shiraiwa, Y. & Suzuki, I. (2014). Functional screening of a novel Delta 15 fatty acid desaturase from the coccolithophorid Emiliania huxleyi. Biochimica et Biophysica Acta-Molecular and Cell Biology of Lipids 1841: 14511458.CrossRefGoogle Scholar
Legeret, B., Schulz-Raffelt, M., Nguyen, H. M. et al. (2016). Lipidomic and transcriptomic analyses of Chlamydomonas reinhardtii under heat stress unveil a direct route for the conversion of membrane lipids into storage lipids. Plant, Cell and Environment 39: 834847.CrossRefGoogle ScholarPubMed
Lu, N., Ding, Y., Zang, X.-N. et al. (2013). Molecular cloning and expression analysis of glutathione peroxidase and glutathione reductase from Gracilaria lemaneiformisunder heat stress. Journal of Applied Phycology 25: 19251931.CrossRefGoogle Scholar
Maikova, A., Zalutskaya, Z., Lapina, T. et al. (2016). The HSP70 chaperone machines of Chlamydomonas are induced by cold stress. Journal of Plant Physiology 204: 8591.CrossRefGoogle ScholarPubMed
Margesin, R., Schinner, F., Marx, J.-C. et al. (2008). Psychrophiles: From Biodiversity to Biotechnology. Springer, Berlin.CrossRefGoogle Scholar
Mathur, S., Agrawal, D. & Jajoo, A. (2014). Photosynthesis: Response to high temperature stress. Journal of Photochemistry and Photobiology B: Biology 137: 116126.CrossRefGoogle ScholarPubMed
Mock, T., Krell, A., Glöckner, G. et al. (2006). Analysis of expressed sequence tags (ESTs) from the polar diatom Fragilariopsis cylindrus. Journal of Phycology 42: 7885.CrossRefGoogle Scholar
Molen, T. A., Rosso, D., Piercy, S. et al. (2006). Characterization of the alternative oxidase of Chlamydomonas reinhardtii in response to oxidative stress and a shift in nitrogen source. Physiologia Plantarum 127: 7486.CrossRefGoogle Scholar
Moreno, A. R. & Martiny, A. C. (2018). Ecological stoichiometry of ocean plankton. Annual Review of Marine Science 10: 4369.CrossRefGoogle ScholarPubMed
Mou, S., Xu, D., Ye, N. et al. (2012). Rapid estimation of lipid content in an Antarctic ice alga (Chlamydomonas sp.) using the lipophilic fluorescent dye BODIPY505/515. Journal of Applied Phycology 24: 11691176.CrossRefGoogle Scholar
Murata, N. & Los, D. A. (1997). Membrane fluidity and temperature perception. Plant Physiology 115: 875879.CrossRefGoogle ScholarPubMed
Neale, P. J. & Priscu, J. C. (1995). The photosynthetic apparatus of phytoplankton from a perennially ice-covered Antarctic lake: Acclimation to an extreme shade environment. Plant and Cell Physiology 36: 253263.CrossRefGoogle Scholar
Pearson, G. A., Lago-Leston, A., Canovas, F. et al. (2015). Metatranscriptomes reveal functional variation in diatom communities from the Antarctic Peninsula. ISME Journal 9: 22752289.CrossRefGoogle ScholarPubMed
Phadtare, S. (2004). Recent developments in bacterial cold-shock response. Current Issues in Molecular Biology 6: 125136.Google ScholarPubMed
Raven, J. A. & Kübler, J. E. (2002). New light on the scaling of metabolic rate with the size of algae. Journal of Phycology 38: 1116.CrossRefGoogle Scholar
Renaut, J., Hausman, J.-F. & Wisniewski, M. E. (2006). Proteomics and low-temperature studies: Bridging the gap between gene expression and metabolism. Physiologia Plantarum 126: 97109.CrossRefGoogle Scholar
Rochaix, J.-D. (2014). Regulation and dynamics of the light-harvesting system. Annual Review of Plant Biology 65: 287309.CrossRefGoogle ScholarPubMed
Rothschild, L. J. (2008). The evolution of photosynthesis … again?. Philosophical Transactions of the Royal Society of London Series B Biological Sciences 363: 27872801.CrossRefGoogle ScholarPubMed
Ryan, C. N., Thomas, M. K. & Litchman, E. (2017). The effects of phosphorus and temperature on the competitive success of an invasive cyanobacterium. Aquatic Ecology 51: 463472.CrossRefGoogle Scholar
Schramm, A., Jakob, T. & Wilhelm, C. (2016). The impact of the optical properties and respiration of algal cells with truncated antennae on biomass production under simulated outdoor conditions. Current Biotechnology 5: 142153.CrossRefGoogle Scholar
Schroda, M., Hemme, D. & Mühlhaus, T. (2015). The Chlamydomonas heat stress response. Plant Journal 82: 466480.CrossRefGoogle ScholarPubMed
Shi, J., Chen, Y., Xu, Y. et al. (2017). Differential proteomic analysis by iTRAQ reveals the mechanism of Pyropia haitanensis responding to high temperature stress. Scientific Reports 7: 44734. https://doi.org/10.1038/srep44734.CrossRefGoogle ScholarPubMed
Shikanai, T. (2007). Cyclic electron transport around Photosystem I: Genetic approaches. Annual Review of Plant Biology 58: 199217.CrossRefGoogle ScholarPubMed
Shin, H., Hong, S.-J., Yoo, C. et al. (2016). Genome-wide transcriptome analysis revealed organelle specific responses to temperature variations in algae. Scientific Reports 6: 37770. https://doi.org/10.1038/srep37770.CrossRefGoogle ScholarPubMed
Smith, R., Anning, J., Clement, P. et al. (1988). Abundance and production of ice algae in resolute passage, Canadian Arctic. Marine Ecology Progress Series 48: 251263.CrossRefGoogle Scholar
Tansey, M. R. & Brock, T. D. (1972). The upper temperature limit for eukaryotic organisms. Proceedings of the National Academy of Sciences USA. 69: 24262428.CrossRefGoogle ScholarPubMed
Tilzer, M. M. & Dubinsky, Z. (1987). Effects of temperature and day length on the mass balance of Antarctic phytoplankton. Polar Biology 7: 3542.CrossRefGoogle Scholar
Wagner, H., Fanesi, A. & Wilhelm, C. (2016). Freshwater phytoplankton responses to global warming. Journal of Plant Physiology 203: 127134.CrossRefGoogle ScholarPubMed
Wagner, H., Jakob, T. & Wilhelm, C. (2006). Balancing the energy flow from captured light to biomass under fluctuating light conditions. New Phytologist 169: 95108.CrossRefGoogle ScholarPubMed
Wilhelm, C. & Selmar, D. (2011). Energy dissipation is an essential mechanism to sustain the viability of plants: The physiological limits of improved photosynthesis. Journal of Plant Physiology 168:7987.CrossRefGoogle ScholarPubMed
Willette, S., Gill, S. S., Dungan, B. et al. (2018). Alterations in lipidome and metabolome profiles of Nannochloropsis salina in response to reduced culture temperature during sinusoidal temperature and light. Algal Research 32: 7992. https://doi.org/10.1016/j.algal.2018.03.001.CrossRefGoogle Scholar
Xiao, H., Chen, C., Xu, Y. et al. (2014). Cloning and expression analysis of the chloroplast fructose-1,6-bisphosphatase gene from Pyropia haitanensis. Acta Oceanologica Sinica 33: 92100. https://doi.org/10.1007/s13131–014–0455–0.CrossRefGoogle Scholar
Yang, J., Tang, H., Zhang, X. et al. (2018). High temperature and pH favor Microcystis aeruginosa to outcompete Scenedesmus obliquus. Environmental Science and Pollution Research 25: 47944802.CrossRefGoogle ScholarPubMed
Yu, C., Xu, K., Wang, W. et al. (2018). Detection of changes in DNA methylation patterns in Pyropia haitanensis under high-temperature stress using a methylation-sensitive amplified polymorphism assay. Journal of Applied Phycology 30: 20912100.CrossRefGoogle Scholar
Zheng, G., Tian, B., Zhang, F. et al. (2011). Plant adaptation to frequent alterations between high and low temperatures: Remodelling of membrane lipids and maintenance of unsaturation levels. Plant, Cell and Environment 34: 14311442.CrossRefGoogle ScholarPubMed

References

Abida, H., Dolch, D.-J., Meï, C. et al. (2015). Membrane glycerolipid remodeling triggered by nitrogen and phosphorus starvation in Phaeodactylum tricornutum. Plant Physiology 167: 118136.CrossRefGoogle ScholarPubMed
Abreu, M. H., Pereira, R., Buschmann, A. H. et al. (2011). Nitrogen uptake responses of Gracilaria vermiculophylla (Ohmi) Papenfuss under combined and single addition of nitrate and ammonium. Journal of Experimental Marine Biology and Ecology 407: 190199.CrossRefGoogle Scholar
Aksnes, D. L. & Egge, J. K. (1991). A theoretical model for nutrient uptake in phytoplankton. Marine Ecology Progress Series 70: 6572.CrossRefGoogle Scholar
Adler, P. B., Salguero-Gómez, R., Compagnoni, A. et al. (2014). Functional traits explain variation in plant life history strategies. Proceedings of the National Academy of Sciences USA 111: 740745.CrossRefGoogle ScholarPubMed
Ale, M. T., Mikkelsen, J. D. & Meyer, A. S. (2011). Differential growth response of Ulva lactuca to ammonium and nitrate assimilation. Journal of Applied Phycology 23: 345351.CrossRefGoogle Scholar
Alexandre, A., Silva, J. & Santos, R. (2010). Inorganic nitrogen uptake and related enzymatic activity of the seagrass Zostera noltii. Marine Biology 31: 539545.Google Scholar
Alexandre, A., Silva, J., Bouma, T. J. et al. (2011). Inorganic nitrogen uptake kinetics and whole-plant nitrogen budget in the seagrass Zostera noltii. Journal of Experimental Marine Biology and Ecology 401: 712.CrossRefGoogle Scholar
Alexandre, A., Georgiou, D. & Santos, R. (2014). Inorganic nitrogen acquisition by the tropical seagrass Halophila stipulacea. Marine Ecology 35: 387394.CrossRefGoogle Scholar
Alexandre, A. & Santos, R. (2020a). Competition for nitrogen between the seaweed Caulerpa prolifera and the seagrass Cymodocea nodosa. Marine Ecology Progress Series 648: 125134.CrossRefGoogle Scholar
Alexandre, A. & Santos, R. (2020b). High nitrogen and phosphorus acquisition by belowground parts of Caulerpa prolifera (Chlorophyta) contribute to the species’ rapid spread in the Ria Formosa saloon, Southern Portugal. Journal of Phycology 56: 608617.CrossRefGoogle Scholar
Alipanah, L., Rohloff, J., Winge, P. et al. (2015). Whole-cell response to nitrogen deprivation in the diatom Phaeodactylum tricornutum. Journal of Experimental Botany 66: 62816296.CrossRefGoogle ScholarPubMed
Allen, J. T, Brown, L., Sanders, R. et al. (2005). Diatom carbon export enhanced by silicate upwelling in the northeast Atlantic. Nature 437: 728732.CrossRefGoogle ScholarPubMed
Allen, A. E., Vardi, A. & Bowler, C. (2006). An ecological and evolutionary context for integrated nitrogen metabolism and related signalling pathways in marine diatoms. Current Opinions in Plant Biology 9: 264273.CrossRefGoogle ScholarPubMed
Allen, A. E., Dupont, C. L., Oborník, M. et al. (2011). Evolution and metabolic significance of the urea cycle in photosynthetic diatoms. Nature 473: 203207.CrossRefGoogle ScholarPubMed
Allen, W. J. & Collinson, I. (2020). Ammonium transporters: A molecular dual carriageway. eLife 9: e61148. https://doi.org/10.7554/eLife.61148CrossRefGoogle ScholarPubMed
Ammerman, J. W. (1991). Role of ecto-phosphohydrolases in phosphorus regeneration in estuarine and coastal ecosystems. In: Chróst, R. J. (ed.) Microbial Enzymes in Aquatic Environments. Springer Verlag, New York, pp. 165186.CrossRefGoogle Scholar
Ammerman, J. W. & Azam, F. (1985). Bacterial 5X-nucleotidase in aquatic ecosystems: A novel mechanism of phosphorus regeneration. Science 227: 13381340.CrossRefGoogle ScholarPubMed
Anbar, A. D. & Knoll, A. H. (2002). Proterozoic ocean chemistry and evolution: A bioinorganic bridge?. Science 297: 11371142.CrossRefGoogle ScholarPubMed
Andresen, E., Edgar, P. & Küpper, H. (2018). Trace metal metabolism in plants. Journal of Experimental Botany 69: 909954.CrossRefGoogle ScholarPubMed
Anderson, S. L. & Burris, J. E. (1987). Role of glutamine synthetase in ammonia assimilation by symbiotic marine dinoflagellates (zooxanthellae). Marine Biology 94: 451458.CrossRefGoogle Scholar
Andrews, M. (1987). Phosphate uptake by the component parts of Chara hispida. British Phycological Journal 22: 4953.CrossRefGoogle Scholar
Andrews, M., Maule, H. G., Raven, J. A. et al. (2005). Extension growth of Impatiens glandulifera at low irradiances: Importance of nitrate and potassium accumulation. Annals of Botany 95: 641658.CrossRefGoogle ScholarPubMed
Andrews, M., Raven, J. A. & Lea, P. J. (2013). Do plants need nitrate? The mechanisms by which nitrogen form affects plants. Annals of Applied Biology 163: 174199.CrossRefGoogle Scholar
Andrews, M. & Raven, J. A. (2022). Root or shoot nitrate assimilation in terrestrial vascular plants – does it matter? Plant and Soil 476: 3162.CrossRefGoogle Scholar
Armbrust, E. V., Berges, J. A., Bowler, C. et al. (2004). The genome of the diatom Thalassiosira pseudonana: Ecology, evolution, and metabolism. Science 306: 7986.CrossRefGoogle ScholarPubMed
Armin, G. & Inomura, K. (2022). Modeling the elemental stoichiometry and silicon accumulation in diatoms. Current Research in Microbial Sciences 3: 100164. https://doi.org/10.1016/j.crmicr.2022.100164.CrossRefGoogle ScholarPubMed
Arrigo, K. R., Dunbar, R. B., Lizotte, M. P. et al. (2002). Taxon-specific differences in C/P and N/P drawdown for phytoplankton in the Ross Sea, Antarctica. Geophysical Research Letters 29: 1938.CrossRefGoogle Scholar
Armin, G. & Inomura, K. (2022). Modeling the elemental stoichiometry and silicon accumulation in diatoms Current Research in Microbial Sciences 3: 100164. https://doi.org/10.1016/j.crmicr.2022.100164.CrossRefGoogle ScholarPubMed
Baines, S. B., Twining, B. S., Brzezinski, M. A. et al. (2012). Significant silicon accumulation by marine picocyanobacteria. Nature Geoscience 5: 886891.CrossRefGoogle Scholar
Baker, D. M., Andras, J. P., Jordá-Garza, A. G. et al. (2013). Nitrate competition in a coral symbiosis varies with temperature among Symbiodinium clades. The ISME Journal 7: 12481251.CrossRefGoogle Scholar
Balaguer, J., Koch, F., Hassler, C. et al. (2022). Iron and manganese co-limit the growth of two phytoplankton groups dominant at two locations of the Drake Passage. Communications Biology 5: 207. https://doi.org/10.1038/s42003-022-03148-8.CrossRefGoogle ScholarPubMed
Beardall, J., Young, E. & Roberts, S. (2001). Approaches for determining phytoplankton nutrient limitation. Aquatic Science 63: 4469.CrossRefGoogle Scholar
Beardall, J., Allen, D., Bragg, J. et al. (2009). Allometry and stoichiometry of unicellular, colonial and multicellular phytoplankton. New Phytologist 181: 295309.CrossRefGoogle ScholarPubMed
Behrenfeld, M. J. & Milligan, A. J. (2013) Photophysiological expressions of iron stress in phytoplankton. Annual Review of Marine Science 5: 217246.CrossRefGoogle ScholarPubMed
Bender, S. J., Parker, M. S. & Armbrust, E. V. (2012). Coupled effects of light and nitrogen source on the urea cycle and nitrogen metabolism over a diel cycle in the marine diatom Thalassiosira pseudonana. Protist 163: 232251.CrossRefGoogle Scholar
Bender, S. J., Durkin, C. A., Berthiaume, C. T. et al. (2014). Transcriptional responses of three model diatoms to nitrate limitation of growth. Frontiers in Marine Science 1: 3. https://doi.org/10.3389/fmars.2014.00003.CrossRefGoogle Scholar
Benitez-Nelson, C. R. (2000). The biogeochemical cycling of phosphorus in marine systems. Earth Science Reviews 51: 109135.CrossRefGoogle Scholar
Berges, J. (1997). Algal nitrate reductase. European Journal of Phycology 32: 38.CrossRefGoogle Scholar
Berges, J. A. & Mulholland, M. R. (2008). ‘Enzymes and nitrogen cycling’. In: Capone, D. G., Bronk, D. A., Mulholland, M. R. & Carpenter, E. J. (eds.) Nitrogen in the Marine Environment. Elsevier, Amsterdam, pp. 13851445.CrossRefGoogle Scholar
Bergman, B., Sandh, G., Lin, S. et al. (2013). Trichodesmium–a widespread marine cyanobacterium with unusual nitrogen fixation properties. FEMS Microbiology Review 37: 286302.CrossRefGoogle ScholarPubMed
Berman‐Frank, I., Lundgren, P. & Falkowski, P. (2003). Nitrogen fixation and photosynthetic oxygen evolution in cyanobacteria. Research in Microbiology 154: 157164.CrossRefGoogle ScholarPubMed
Berman-Frank, I., Quigg, A., Finkel, Z. V. et al. (2007). Cyanobacterial strategy of nitrogen-fixation influences diazotroph dependence on iron resources. Limnology and Oceanography 52: 22602269.CrossRefGoogle Scholar
Bertilsson, S., Berglund, O., Karl, D. M. et al. (2003). Elemental composition of marine Prochlorococcus and Synechococcus: Implications for the ecological stoichiometry of the sea. Limnology and Oceanography 48: 17211731.CrossRefGoogle Scholar
Björkman, K. M. & Karl, D. M. (1994). Bioavailability of inorganic and organic P compounds to natural assemblages of microorganisms in Hawaiian coastal waters. Marine Ecology Progress Series 111: 265273.CrossRefGoogle Scholar
Björkman, K. M. & Karl., D. M. (2005). Presence of dissolved nucleotides in the North Pacific Subtropical Gyre and their role in cycling of dissolved organic phosphorus. Aquatic Microbial Ecology 39: 193203.CrossRefGoogle Scholar
Bowler, C., Allen, A. E., Badger, J. H. et al. (2008). The Phaeodactylum genome reveals the evolutionary history of diatom genomes. Nature 456: 239244.CrossRefGoogle ScholarPubMed
Bowler, C., De Martino, A. & Falciatore, A. (2010). Diatom cell division in an environmental context. Current Opinion in Plant Biology 13: 623630.CrossRefGoogle Scholar
Boyd, P. W., Strzepek, R., Fu, F. et al. (2010). Environmental control of open-ocean phytoplankton groups: Now and in the future. Limnology and Oceanography 55: 13531376.CrossRefGoogle Scholar
Box, R. J., Andrews, M. & Raven, J. A. (1984). Intracellular transport and cytoplasmic streaming in Chara hispida. Journal of Experimental Botany 35: 10161021.CrossRefGoogle Scholar
Box, R. J., Boxer, M. & Boxer, D. (1985). Compartmentation of PO4-uptakeand 32P efflux in the rhizoid cells of Chara hispida L. Biochimie und Physiologie der Pflanzen 180: 551555.Google Scholar
Box, R. J. (1986). Quantitative short-term uptake of inorganic phosphate by the Chara hispida rhizoid. Plant Cell and Environment 9: 501506.CrossRefGoogle Scholar
Box, R. J. (1987). The uptake of nitrate and ammonium nitrogen in Chara hispida L.: Contribution of the rhizoid. Plant Cell and Environment 10: 169176.CrossRefGoogle Scholar
Branco, P., Stomp, M., Egas, M. et al. (2010). Evolution of nutrient uptake reveals a trade‐off in the ecological stoichiometry of plant‐herbivore interactions. The American Naturalist 176: E162E176. https://doi.org/10.1086/657036.CrossRefGoogle ScholarPubMed
Brembu, T., Mühlroth, A., Alipanah, L. et al. (2017). The effects of phosphorus limitation on carbon metabolism in diatoms. Philosophical Transactions of the Royal Society B 372: 20160406. https://doi.org/10.1098/rstb.2016.0406.CrossRefGoogle ScholarPubMed
Brewer, P. G. & Goldman, J. C. (1976). Alkalinity changes generated by phytoplankton growth. Limnology and Oceanography 21: 108117.CrossRefGoogle Scholar
Bristow, J. M. & Whitcombe, M. (1971). The role of roots in the nutrition of aquatic vascular plants. American Journal of Botany 58: 813.CrossRefGoogle Scholar
Bristow, L. A., Mohr, W., Ahmerkamp, S. et al. (2017). Nutrients that limit growth in the ocean. Current Biology 27: R431R510.CrossRefGoogle ScholarPubMed
Brodersen, K. E., Koren, K., Moβhanmer, M. et al. (2017). Seagrass-mediated phosphorus and iron solubilisation in tropical sediments. Science and Technology 51: 1415514163.CrossRefGoogle ScholarPubMed
Bromke, M. A., Giavalisco, P., Willmitzer, L. et al. (2013). Metabolic analysis of adaptation to short-term changes in culture conditions of the marine diatom Thalassiosira pseudonana. PLOS ONE 8: e67340. https://doi.org/10.1371/journal.pone.0067340.CrossRefGoogle ScholarPubMed
Bromke, M. A., Sabir, J. S., Alfassi, F. A. et al. (2015). Metabolomic profiling of 13 diatom cultures and their adaptation to nitrate limited growth conditions. PLOS ONE 10: e0138965. https://doi.org/10.1371/journal.pone.0138965.CrossRefGoogle ScholarPubMed
Bronk, D. A., See, J. H., Bradley, P. et al. (2007). DON as a source of bioavailable nitrogen for phytoplankton. Biogeosciences 4: 283296.CrossRefGoogle Scholar
Browning, T. J., Achterberg, E. P., Engel, A. et al. (2021). Manganese co-limitation of phytoplankton growth and major nutrient drawdown in the Southern Ocean. Nature Communications 12: 884.CrossRefGoogle ScholarPubMed
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
Brzezinski, M. A., Olson, R. & Chisholm, S. (1990). Silicon availability and cell-cycle progression in marine diatoms. Marine Ecology Progress Series 67: 8396.CrossRefGoogle Scholar
Brzezinski, M. A., Krause, J. W., Baines, S. B. et al. (2017). Patterns and regulation of silicon accumulation in Synechococcus spp. Journal of Phycology 53: 746761.CrossRefGoogle ScholarPubMed
Brzezinski, M. A., Closset, I., Jones, J. L. et al. (2021). New constraints on the physical and biological controls on the silicon isotopic composition of the Arctic Ocean. Frontiers in Marine Science 8: 699762. https://doi.org/10.3389/fmars.2021.699762.CrossRefGoogle Scholar
Cáceres, C., Spatharis, S., Kaiserli, E. et al. (2019). Temporal phosphate gradients reveal diverse acclimation responses in phytoplankton phosphate uptake. The ISME Journal 13: 28342845.CrossRefGoogle ScholarPubMed
Callon, O., Bulte, L., Kuras, R. et al. (1993). Extensive accumulation of an extracellular l-amino acid oxidase during gametogenesis of Chlamydomonas reinhardtii. European Journal of Biochemistry 215: 351360.Google Scholar
Capone, D. G., Zehr, J. P., Paerl, H. W. et al. (1997). Trichodesmium, a globally significant marine cyanobacterium. Science 276: 12211229.CrossRefGoogle Scholar
Capone, D. G. (2008). The marine nitrogen cycle. Microbe 3: 186192.Google Scholar
Carpenter, E., Montoya, J., Burns, J. et al. (1999). Extensive bloom of a N2-fixing diatom/cyanobacterial association in the tropical Atlantic Ocean. Marine Ecology Progress Series 185: 273283.CrossRefGoogle Scholar
Casey, J. R., Lomas, M. W., Michelou, V. K. et al. (2009). Phytoplankton taxon-specific orthophosphate (Pi) and ATP utilization in the western subtropical North Atlantic. Aquatic Microbial Ecology 58: 3144.CrossRefGoogle Scholar
Cedergreen, N. & Madsen, T. V. (2003a). Nitrate reductase activity in roots and shoots of aquatic macrophytes. Aquatic Botany 76: 203212.CrossRefGoogle Scholar
Cedergreen, N. & Madsen, T. V. (2003b). Light regulation of root and leaf NO3 uptake and reduction in the floating macrophyte Lemna minor. New Phytologist 161: 449457.CrossRefGoogle ScholarPubMed
Cembella, A. D., Antia, N. J. & Harrison, P. J. (1982). The utilization of inorganic and organic phosphorous compounds as nutrients by eukaryotic microalgae: A multidisciplinary perspective. CRC Critical Reviews in Microbiology 10: 317391.CrossRefGoogle Scholar
Chen, X.-H., Li, Y.-Y., Zhang, H. et al. (2018). Quantitative proteomics reveals common and specific responses of a marine diatom Thalassiosira pseudonana to different macronutrient deficiencies. Frontiers in Microbiology 9: 2761. https://doi.org/10.3389/fmicb.2018.02761.CrossRefGoogle ScholarPubMed
Chisholm, S. W. & Stross, R. G. (1976). Phosphate uptake kinetics in Euglena gracilis (Euglenophyceae) in light-dark cycles. 2. Phased PO4-limited cultures. Journal of Phycology 12: 217222.Google Scholar
Chisholm, J. R. M., Douga, C., Ageron, E. et al. (1996). ‘Roots’ in mixotrophic algae. Nature 381: 382.CrossRefGoogle Scholar
Christiansen, N. H., Andersen, F. Ø. & Jensen, H. S. (2016). Phosphate uptake kinetics for four species of submerged freshwater macrophytes measured by a 33P phosphate radioisotope technique. Aquatic Botany 128: 5867.CrossRefGoogle Scholar
Chung, C. C., Hwang, S. P. L. & Chang, J. (2003). Identification of a high-affinity phosphate transporter gene in a prasinophyte alga, Tetraselmis chui, and its expression under nutrient limitation. Applied Aquatic Botany Environmental Microbiology 69: 754759.CrossRefGoogle Scholar
Cochlan, W. P. & Harrison, P. J. (1991). Uptake of nitrate, ammonium, and urea by nitrogen-starved cultures of Micromonas pusilla (Prasinophyceae): Transient responses. Journal of Phycology 27: 673679.CrossRefGoogle Scholar
Codispoti, L. A. (1989). ‘Phosphorus vs. nitrogen limitation in new and export production’. In: Berger, W. H., Smetacek, V. S. & Wefer, G. (eds.) Productivity of Oceans: Present and Past. Wiley, Chichester, pp. 372394.Google Scholar
Codispoti, L. A. (1995). Is the ocean losing nitrate? Nature 376: 724.CrossRefGoogle Scholar
Cohen, R. A. & Fong, P. (2006). Using opportunistic green macroalgae as indicators of nitrogen supply and sources to estuaries. Ecological Applications 16: 14051420.CrossRefGoogle ScholarPubMed
Collos, Y. & Slawyk, G. (1980). ‘Nitrogen uptake and assimilation by marine phytoplankton’. In: , P. G. Falkowski, (ed.) Primary Productivity in the Sea. Plenum Press, New York, pp. 195211.CrossRefGoogle Scholar
Collos, Y. (1986). Time-lag algal growth dynamics: Biological constraints on primary production in aquatic environments. Marine Ecology Progress Series 33: 193206.CrossRefGoogle Scholar
Conti-Jerpe, I. E., Thompson, P. D., Wong, C. W. M. et al. (2020). Trophic strategy and bleaching resistance in reef-building corals. Science Advances 6: z5443. https://doi.org/10.1126/sciadv.aaz5443.CrossRefGoogle ScholarPubMed
Corbridge, D. E. C. (1990). Phosphorus, Studies in Inorganic Chemistry. Elsevier, Oxford.Google Scholar
Corcoran, A. A. & Boeing, W. J. (2012). Biodiversity increases the productivity and stability of phytoplankton communities. PLOS ONE 7: 19.CrossRefGoogle ScholarPubMed
Cullen, J. J. (2015). Subsurface chlorophyll maximum layers: Enduring enigma or mystery solved? Annual Review of Marine Science 7: 207239.CrossRefGoogle ScholarPubMed
Dagenais-Bellefeuille, S. & Morse, D. (2013). Putting the N in dinoflagellates. Frontiers in Microbiology 4: 369. https://doi.org/10.3389/micb.2013.00369.CrossRefGoogle Scholar
Dagestad, D., Lien, T. & Knutsen, G. (1981). Degradation and compartmentalization of urea in Chlamydomonas reinhardii. Archives of Microbiology 129: 261264.CrossRefGoogle Scholar
D’Elia, C. F., Domotor, S. L. & Webb, K. L. (1983). Nutrient uptake kinetics of freshly isolated zooxanthellae. Marine Biology 75: 157167.CrossRefGoogle Scholar
Denny, P. (1972). Sites of nutrient absorption in aquatic macrophytes. Journal of Ecology 60: 819829.CrossRefGoogle Scholar
Derelle, E., Ferraz, C., Rombautz, S. et al. (2006). Genome analysis of the smallest free-living eukaryote Ostreococcus tauri unveils many unique features. Proceedings of the National Academy of Sciences USA 103: 1164711652.CrossRefGoogle ScholarPubMed
Deutsch, C. & Weber, T. (2012). Nutrient ratios as a tracer and driver of ocean biogeochemistry. Annual Review of Marine Science 4: 113141.CrossRefGoogle ScholarPubMed
Diaz, J., Ingall, E., Benitez-Nelson, C. et al. (2008). Marine polyphosphate: A key player in geo-logic phosphorus sequestration. Science 320: 652655.CrossRefGoogle Scholar
Dorado, S., Booe, T., Steichen, J. et al. (2015). Towards an understanding of the interactions between freshwater inflows and phytoplankton communities in subtropical estuaries. PLOS ONE 10: e0130931. https://doi.org/10.1371/journal.pone.0130931.CrossRefGoogle Scholar
Donald, K. M., Scanlan, D. J., Carr, N. G. et al. (1997). Comparative phosphorus nutrition of the marine cyanobacterium Synechococcus WH7803 and the marine diatom Thalassiosira weissflogii. Journal of Plankton Research 19: 17931813.CrossRefGoogle Scholar
Dortch, Q., Clayton, J. R., Thoreson, S. S. et al. (1982). Response of marine phytoplankton to nitrogen deficiency: Decreased nitrate uptake vs enhanced ammonium uptake. Marine Biology 70: 1319.CrossRefGoogle Scholar
Dortch, Q., Thompson, P. A. & Harrison, P. J. (1991). Short-term interaction between nitrate and ammonium uptake in Thalassiosira pseudonana: Effect of preconditioning nitrogen source and growth rate. Marine Biology 110: 183193.CrossRefGoogle Scholar
Droop, M. R. (1968). Vitamin B12 and marine ecology. IV. The kinetics of uptake, growth and inhibition in Monochrysis lutheri. Journal of the Marine Biological Association of the UK 48: 689733.CrossRefGoogle Scholar
Droop, M. R. (1973). Some thoughts on nutrient limitation in algae. Journal of Phycology 9: 264272.CrossRefGoogle Scholar
Droop, M. R. (1979). On the definition of X and Q in the cell quota model. Journal of Experimental Marine Biology and Ecology 39: 203. https://doi.org/10.1016/0022–0981(79)90014–5.CrossRefGoogle Scholar
Du, C., Liang, J. R., Chen, D. D. et al. (2014). iTRAQ-based proteomic analysis of the metabolism mechanism associated with silicon response in the marine diatom Thalassiosira pseudonana. Journal of Proteome Research 13: 720734.CrossRefGoogle Scholar
Duarte, C. M. (1992). Nutrient concentration of aquatic plants: Patterns across species. Limnology and Oceanography 37: 882889.CrossRefGoogle Scholar
Duarte, C. M., Borum, J., Short, F. T. et al. (2008). Seagrass ecosystems: Their global status and prospects. In: Polunin, N. V. C. (ed.) Aquatic ecosystems: Trends and Global Prospects. Cambridge University Press, Cambridge, pp. 281294.CrossRefGoogle Scholar
Duarte, C. M., Marbá, N., Gacia, E. et al. (2010). Seagrass community metabolism: Assessing the carbon sink capacity of seagrass meadows. Global Biogeochemical Cycles 24: GB4032. https://doi.org/10.1029/2010GB003793.CrossRefGoogle Scholar
Ducklow, H. W. & Doney, S. C. (2013). What is the metabolic state of the oligotrophic ocean? A debate. Annual Review of Marine Science 5: 525533.CrossRefGoogle ScholarPubMed
Dugdale, R. C. & Goering, J. J. (1967). Uptake of new and regenerated forms of nitrogen in primary productivity. Limnology and Oceanography 12: 196206.CrossRefGoogle Scholar
Duhamel, S., Björkman, K. M. & Karl, D. M. (2012). Light dependence of phosphorus uptake by microorganisms in the subtropical North and South Pacific Ocean. Aquatic Microbial Ecolog 67: 225238.CrossRefGoogle Scholar
Dyhrman, S. T. & Palenik, B. P. (1997). The identification and purification of a cell-surface alkaline phosphatase from the dinoflagellate Prorocentrum minimum (Dinophycaeae). Journal of Phycology 33: 602612.CrossRefGoogle Scholar
Dyhrman, S. T., & Palenik, B. P. (2003). A characterization of ectoenzyme activity and phosphate-regulated proteins in the coccolithophorid Emiliania huxleyi. Journal of Plankton Research 25: 111.CrossRefGoogle Scholar
Dyhrman, S. T., Haley, S. T., Birkeland, S. R. et al. (2006). Long serial analysis of gene expression for gene discovery and transcriptome profiling in the widespread marine coccolithophore Emiliania huxleyi. Applied Environmental Microbiology 72: 252260.CrossRefGoogle ScholarPubMed
Dyhrman, S. T., Ammerman, J. W. & Van Mooy, B. A. (2007). Microbes and the marine phosphorus cycle. Oceanography 20: 110116.CrossRefGoogle Scholar
Dyhrman, S., Benitez-Nelson, C., Orchard, E. et al. (2009). A microbial source of phosphonates in oligotrophic marine systems. Nature Geoscience 2: 696699.CrossRefGoogle Scholar
Dyhrman, S. T., Jenkins, B. D., Rynearson, T. A. et al. (2012). The transcriptome and proteome of the diatom Thalassiosira pseudonana reveal a diverse phosphorus stress response. PLOS ONE 7: e33768.CrossRefGoogle ScholarPubMed
Dyhrman, S. T. (2016). Nutrients and their acquisition: Phosphorus physiology in microalgae. In: Borowitzka, M. A., Beardall, J. & Raven, J. A. (eds.) The Physiology of Microalgae. Springer, Dordrecht, The Netherlands, pp. 155183.CrossRefGoogle Scholar
Edwards, K. F., Thomas, M. K., Klausmeier, C. A. et al. (2012). Allometric scaling and taxonomic variation in nutrient utilization traits and maximum growth rate of phytoplankton. Limnology and Oceanography 57: 554566.CrossRefGoogle Scholar
Ehrlich, H., Konstantinos, D. D., Pokrovsky, O. S. et al. (2010). Modern views on desilicification: Biosilica and abiotic silica dissolution in natural and artificial environments. Chemical Reviews 110: 46564689.CrossRefGoogle ScholarPubMed
Elser, J. J., Dobberfuhl, D., MacKay, N. A. et al. (1996). Organism size, life history, and N:P stoichiometry: Towards a unified view of cellular and ecosystem processes. BioScience 46: 674684. https://doi.org/10.2307/1312897.CrossRefGoogle Scholar
Elser, J. J., Fagan, W. F., Denno, R. F. et al. (2000). Nutrient constraints in terrestrial and freshwater food webs. Nature 408: 578580.CrossRefGoogle ScholarPubMed
Elser, J. J., Acharya, M., Kyle, J. et al. (2003). Growth rate-stoichiometry couplings in diverse biota. Ecology Letters 6: 936943.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: 18.CrossRefGoogle ScholarPubMed
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. & Stone, D. P. (1975). Nitrate uptake in marine phytoplankton: Energy sources and the interaction with carbon fixation. Marine Biology 32: 7784.CrossRefGoogle Scholar
Falkowski, P. G. (1997). Evolution of the nitrogen cycle and its influence on the biological CO2 pump in the ocean. Nature 387: 272275.CrossRefGoogle Scholar
Falkowski, P. G., Barber, R. T. & Smetacek, V. (1998). Biogeochemical controls and feedbacks on ocean primary production. Science 281: 200206.CrossRefGoogle ScholarPubMed
Falkowski, P. G. (2000). Rationalizing elemental ratios in unicellular algae. Journal of Phycology 36: 36.CrossRefGoogle Scholar
Falkowski, P. G., Katz, M. E., Knoll, A. H. et al. (2004). The evolutionary history of eukaryotic phytoplankton. Science 305: 354360. https://doi.org/10.1126/science.1095964.CrossRefGoogle Scholar
Falkowski, P. G. & Raven, J. A. (2007). Aquatic Photosynthesis, 2nd ed. Princeton University Press, Princeton, NJ.CrossRefGoogle Scholar
Ferdie, M. & Fourqurean, J. W. (2004). Responses of seagrass communities to fertilization along a gradient of relative availability of nitrogen and phosphorus in a carbonate environment. Limnology and Oceanography 49: 20822094.CrossRefGoogle Scholar
Fernandez, E., Llamas, A. & Galvàn, A. (2009). Nitrogen assimilation and its regulation. In: Stern, D. B. & Harris, E. H. (eds.) The Chlamydomonas Source Book, Vol. 2, 2nd ed. Elsevier, Amsterdam, pp. 69113.CrossRefGoogle Scholar
Field, C. B., Behrenfeld, M. J., Randerson, J. T. et al. (1998). Primary production of the biosphere: Integrating terrestrial and oceanic components. Science 281: 237240.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
Finkel, Z. V., Follows, M. J., Liefer, J. et al. (2016). Phylogenetic diversity in the macromolecular composition of microalgae. PLOS ONE 11: e0155977. https://doi.org/10.1371/journal.pone.0155977.CrossRefGoogle ScholarPubMed
Flores, E. & Herrero, A. (1994). ‘Assimilatory nitrogen metabolism and its regulation’. In: , A. Bryant, (ed.) The Molecular Biology of Cyanobacteria. Kluwer Academic Publications, Dordrecht, pp. 487517.CrossRefGoogle Scholar
Flores, E., Frías, J. E., Rubio, L. M. et al. (2005). Photosynthetic nitrate assimilation in cyanobacteria. Photosynthesis Research 83: 117133.CrossRefGoogle ScholarPubMed
Flynn, K. J. (1998). Estimation of kinetic parameters for the transport of nitrate and ammonium into marine phytoplankton. Marine Ecology Progress Series 169: 1328.CrossRefGoogle Scholar
Flynn, K. J., Raven, J. A., Rees, T. A. K. et al. (2010). Is the growth rate hypothesis applicable to microalgae? Journal of Phycology 46: 112.CrossRefGoogle Scholar
Fong, P., Boyer, K. E. & Zedler, J. B. (1998). Developing an indicator of nutrient enrichment in coastal estuaries and lagoons using tissue nitrogen content of opportunistic alga, Enteromorpha intestinalis (L. Link). Journal of Experimental Marine Biology and Ecology 231: 6370.CrossRefGoogle Scholar
Fourqurean, J. W., Duarte, C. M., Kennedy, H. et al. (2012). Seagrass ecosystems as a globally significant carbon stock. Nature Geoscience 5: 505509.CrossRefGoogle Scholar
Fowler, D., Coyle, M., Skiba, U. et al. (2013). The global nitrogen cycle in the twenty-first century. Philosophical Transactions of the Royal Society B 368: 20130164. https://doi.org/10.1098/rstb.2013.0164.CrossRefGoogle ScholarPubMed
Fraser, M. W., Gleesons, D. B., Grierson, P. F. et al. (2018). Metagenomic evidence of microbial community responsiveness to phosphorus and salinity gradients ion seagrass sediments. Frontiers in Microbiolog 9: 1703.CrossRefGoogle Scholar
Frausto da Silva, J. J. R. & Williams, R. J. P. (2001). The Biological Chemistry of the Elements: The Inorganic Chemistry of Life. Oxford University Press, Oxford, UK.CrossRefGoogle Scholar
Freeman, L. A., Corbett, D. R., Fitzgerald, A. et al. (2019). Impacts of urbanization on estuarine ecosystems and water quality. Estuaries and Coasts 42: 18211838.CrossRefGoogle Scholar
Frigeri, L. G., Radabaugh, T. R., Haynes, P. A., et al. (2006). Identification of proteins from a cell wall fraction of the diatom Thalassiosira pseudonana: Insights into silica structure formation. Molecular and Cell Proteomics 5: 182193.CrossRefGoogle ScholarPubMed
Fu, F. X., Zhang, Y. H., Leblanc, K. et al. (2005). The biological and biogeochemical consequences of phosphate scavenging onto phytoplankton cell surfaces. Limnology and Oceanography 50: 14591472.CrossRefGoogle Scholar
Gallon, J. R. (2001). N2 fixation in phototrophs: Adaptation to a specialized way of life. Plant Soil 230: 3948.CrossRefGoogle Scholar
Geider, R. J. & La Roche, J. (2002). Redfield revisited: Variability of C:N:P in marine microalgae and its biochemical basis. European Journal of Phycology 37: 117.CrossRefGoogle Scholar
Glibert, P. M., Wilkerson, F. P., Dugdale, R. C. et al. (2016). Pluses and minuses of ammonium and nitrate uptake and assimilation by phytoplankton and implications for productivity and community composition, with emphasis on nitrogen-enriched conditions. Limnology and Oceanography 61: 165197.CrossRefGoogle Scholar
Glibert, P. M. (2019). Phytoplankton in the aqueous ecological theater: Changing conditions, biodiversity, and evolving ecological concepts. Journal of Marine Research 77: 83137.CrossRefGoogle Scholar
Giordano, M. & Raven, J. A. (2014). Nitrogen and sulfur assimilation in plants and algae. Aquatic Botany 118: 4561.CrossRefGoogle Scholar
Goldman, J. C., McCarthy, J. J. & Peavey, D. G. (1979). Growth rate influence on the chemical composition of phytoplankton in oceanic waters. Nature 279: 21102215.CrossRefGoogle Scholar
Goldman, J. C. & Glibert, P. M. (1983). Kinetics of inorganic nitrogen uptake by phytoplankton. In: Carpenter, E. J. & Capone, D. G. (eds.) Nitrogen in the Marine Environment. Academic Press, Cambridge, pp. 233274.CrossRefGoogle Scholar
Gotham, I. J. & Rhee, G. (1981). Comparative kinetic studies of phosphate-limited growth and phosphate uptake in phytoplankton in continuous culture. Journal of Phycology 17: 257265.CrossRefGoogle Scholar
Graziano, L. M., La Roche, J. & Geider, R. J. (1996). Physiological response to phosphorus limitation in batch and steady-state cultures of Dunaliella tertiolecta (Chlorophyta): A unique stress protein as an indicator of phosphate deficiency. Journal of Phycology 32: 825838.CrossRefGoogle Scholar
Grossman, A. & Takahashi, H. (2001). Macronutrient utilization by photosynthetic eukaryotes and the fabric of interactions. Annual Review of Plant Physiology and Plant Molecular Biology 52: 163210.CrossRefGoogle ScholarPubMed
Grossman, A. R. & Aksoy, M. (2015). Algae in a phosphorus-limited landscape. Annual Plant Reviews 48: 337374.Google Scholar
Grover, J. P. (1991). Resource competition in a variable environment – Phytoplankton growing according to the variable-internal-stores model. American Naturalist 138: 811835.CrossRefGoogle Scholar
Gruber, N. & Sarmiento, J. L. (1997). Global patterns of marine nitrogen fixation and denitrification. Global Biogeochemical Cycles 11: 235266.CrossRefGoogle Scholar
Gruber, N. & Galloway, J. N. (2008). An Earth system perspective of the global nitrogen cycle. Nature 451: 293296.CrossRefGoogle ScholarPubMed
Guerra, L. T., Levitan, O., Frada, M. J. et al. (2013). Regulatory branch points affecting protein and lipid biosynthesis in the diatom Phaeodactylum tricornutum. Biomass and Bioenergy 59: 306315.CrossRefGoogle Scholar
Gunnersen, J., Yellowlees, D. & Miller, D. J. (1988). The ammonium/methylammonium uptake system of Symbiodinium microadriaticum. Marine Biology 97: 593596.CrossRefGoogle Scholar
Haines, K. C. & Wheeler, P. A. (1978). Ammonium and nitrate uptake by the marine macrophytes Hypnea musciformis (Rhodophyta) and Macrocystis pyrifera (Phaeophyta). Journal of Phycology 14: 319324.CrossRefGoogle Scholar
Hanisak, M. D. & Harlin, M. M. (1978). Uptake of inorganic nitrogen by Codium fragile subsp. tomentosoides (Chlorophyta). Journal of Phycology 14: 450454.CrossRefGoogle Scholar
Harrison, G. I., Harris, L. R. & Irwin, B. D. (1996). The kinetics of nitrogen utilization in the oceanic mixed layer: Nitrate and ammonium interactions at nanomolar concentrations. Limnology and Oceanography 41: 1632.CrossRefGoogle Scholar
Harrison, P. J. & Hurd, C. L. (2001). Nutrient physiology of seaweeds: Application of concepts to aquaculture. Cahiers de Biologie Marine 42: 7182.Google Scholar
Harke, M. J., Juhl, A. R., Haley, S. T. et al. (2017). Conserved transcriptional responses to nutrient stress in bloom-forming algae. Frontiers in Microbiology 8: 1279. https://doi.org/10.3389/fmicb.2017.01279.CrossRefGoogle ScholarPubMed
Healey, P. F. (1979). Short-term responses of nutrient-deficient algae to nutrient addition. Journal of Phycology 15: 289299.Google Scholar
Healey, F. P. (1980). Slope of the Monod equation as an indicator of advantage in nutrient competition. Microbial Ecology 5: 281286.CrossRefGoogle ScholarPubMed
Hecky, R. E. & Kilham, P. (1988). Nutrient limitation of phytoplankton in freshwater and marine environments: A review of recent evidence on the effects of enrichment. Limnology and Oceanography 33: 796822.Google Scholar
Held, N. A., Webb, E. A., McIlvin, M. M. et al. (2020). Co-occurrence of Fe and P stress in natural populations of the marine diazotroph Trichodesmium. Biogeosciences 17: 25372551.CrossRefGoogle Scholar
Hildebrand, M., Volcani, B. E., Gassmann, W. et al. (1997). A gene family of silicon transporters. Nature 385: 688689.CrossRefGoogle ScholarPubMed
Hildebrand, M., Dahlin, K. & Volcani, B. E. (1998). Characterization of a silicon transporter gene family in Cylindrotheca fusiformis: Sequences, expression analysis, and identification of homologs in other diatoms. Molecular and General Genetics 260: 480486.CrossRefGoogle ScholarPubMed
Ho, T.-Y., Quigg, A., Finkel, Z. V. et al. (2003). On the elemental composition of some marine phytoplankton. Journal of Phycology 39: 115.CrossRefGoogle Scholar
Hockin, N. L., Mock, T., Mulholland, F. et al. (2012). The response of diatom central carbon metabolism to nitrogen starvation is different from that of green algae and higher plants. Plant Physiology 158: 299312.CrossRefGoogle ScholarPubMed
Howarth, R. W. Marino, R. & Cole, J. J. (1988). Nitrogen fixation in freshwater, estuarine and marine ecosystems. 2. Biogeochemical controls. Limnology and Oceanography 33: 688701.Google Scholar
Howarth, R. W. & Marino, R. (2006). Nitrogen as the limiting nutrient for eutrophication in coastal marine ecosystems: Evolving views over three decades. Limnology and Oceanography 51: 364376.CrossRefGoogle Scholar
Hughes, A. D. & Grottoli, A. G. (2013). Heterotrophic compensation: A possible mechanism for resilience of coral reefs to global warming or a sign of prolonged Stress? PLOS ONE 8: e81172. https://doi.org/10.1371/journal.pone.0081172.CrossRefGoogle ScholarPubMed
Hughes, A. R., Stacowicz, J. J. & Williams, S. L. (2009). Morphological and physiological variation among seagrass (Zostera marina) genotypes. Oecologia 159: 725733.CrossRefGoogle ScholarPubMed
Huppe, H. C. & Turpin, D. H. (1994). Integration of carbon and nitrogen metabolism in plant and algal cells. Annual Review of Plant Physiology and Plant Molecular Biology 45: 577607.CrossRefGoogle Scholar
Hurd, C. L., Harrison, P. J. & Druehl, L. D. (1996). The effect of seawater flow velocity on nutrient uptake by morphologically distinct forms of Macrocystis integrifolia from sheltered and exposed sites. Marine Biology 126: 205214.CrossRefGoogle Scholar
Hurd, C. L., Harrison, P. J., Bischof, K. et al. (2014). Seaweed Ecology and Physiology. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Hwang, Y. S., Jung, G. & Jin, E. (2008). Transcriptome analysis of acclimatory responses to thermal stress in Antarctic algae. Biochemical and Biophysical Research Communications 367: 635641.CrossRefGoogle ScholarPubMed
Igarashi, R. & Seefeldt, L. (2003). Nitrogen fixation: The mechanism of the Mo-dependent nitrogenase. Critical Reviews in Biochemistry and Molecular Biology 38: 351384. https://doi.org/10.1080/10409230391036766.CrossRefGoogle ScholarPubMed
Irwin, A. J., Finkel, Z. V., Schofield, O. M. et al. (2006). Scaling-up from nutrient physiology to the size-structure of phytoplankton communities. Journal of Plankton Research 28: 459471.CrossRefGoogle Scholar
Jackson, A. E. & Yellowlees, D. (1990). Phosphate uptake by zooxanthellae isolated from corals. Proceedings of the Royal Society: Biological Sciences 242: 201204.Google Scholar
Jacobson, L. and Halmann, M. (1982). Polyphosphate metabolism in the blue-green alga Microcystis aeruginosa. Journal of Plankton Research 4: 481488.CrossRefGoogle Scholar
Janauer, G. A. (1981a). Distribution of organic and mineral components in leaves of Ranunculus fluitans Lam. Hydrobiologia 80: 193204.CrossRefGoogle Scholar
Janauer, G. A. (1981b). Elodea canadensis and its dormant apices: An investigation of organic and mineral constituents. Aquatic Botany 11: 231243.CrossRefGoogle Scholar
Janauer, G. A. (1981c). Divergence of organic anio and amino compound concentrations in different parts of leaves of Posidonia oceanica (L.) Delile. Biochimie und Physiologie der Pflanzen 176: 314321.CrossRefGoogle Scholar
Jones, G. J. & Morel, F. M. (1988). Plasmalemma redox activity in the diatom Thalassiosira: A possible role for nitrate reductase. Plant Physiology 87: 143147.CrossRefGoogle ScholarPubMed
Kamalanathan, M., Pierangelini, M., Shearman, L. A. et al. (2016). Impacts of nitrogen and phosphorus starvation on the physiology of Chlamydomonas reinhardtii. Journal of Applied Phycology 28: 15091520.CrossRefGoogle Scholar
Karl, D. M., Letelier, R., Tupas, L. et al. (1997). The role of nitrogen fixation in biochemical cycling in the subtropical North Pacific Ocean. Nature 388: 533538.CrossRefGoogle Scholar
Karl, D., Michaels, A., Bergman, B. et al. (2002). Dinitrogen fixation in the world’s oceans. Biogeochemistry 57: 4798.CrossRefGoogle Scholar
Karl, D. M. (2014). Microbially mediated transformations of phosphorus in the sea: New views of an old cycle. Annual Review of Marine Science 6: 279337.CrossRefGoogle ScholarPubMed
Kevekordes, K. (2001). Toxicity tests using developmental stages of Hormosira banksii (Phaeophyta) identify ammonium as a damaging component of secondary treated sewage effluent discharged into Bass Strait, Victoria, Australia. Marine Ecology Progress Series 219: 139148.CrossRefGoogle Scholar
Kevekordes, K., Holland, D., Häubner, N. et al. (2006). Inorganic carbon acquisition by eight species of Caulerpa (Caulerpales, Chlorophyta). Phycologia 45: 442448.CrossRefGoogle Scholar
Kirkman, H., Griffiths, F. B. & Parker, R. R. (1979). The release of reactive phosphate by a Posidionia australia seagrass community. Aquatic Botany 6: 329337.CrossRefGoogle Scholar
Klausmeier, C. A., Litchman, E., Daufresne, T. et al. (2004). Optimal nitrogen-to-phosphorus stoichiometry of phytoplankton. Nature 429: 171174.CrossRefGoogle ScholarPubMed
Knight, M., Senior, L., Nancolas, B. et al. (2016). Direct evidence of the molecular basis for biological silicon transport. Nature Communications 7: 11926. https://doi.org/10.1038/ncomms11926.CrossRefGoogle ScholarPubMed
Krumhardt, K. M., Callnan, K., Roache-Johnson, K. et al. (2013). Effects of phosphorus starvation versus limitation on the marine cyanobacterium Prochlorococcus MED4 I: Uptake physiology. Environmental Microbiology 15: 21142128.CrossRefGoogle ScholarPubMed
Kudela, R. M. & Dugdale, R. C. (2000). Nutrient regulation of phytoplankton productivity in Monterey Bay, California. Deep-Sea Research Part II Topical Studies in Oceanography 47: 10231053.CrossRefGoogle Scholar
Lapointe, B. E., Barile, P. J., Littler, M. M. et al. (2005). Macroalgal blooms on southeast Florida coral reefs I. Nutrient stoichiometry of the invasive green alga Codium isthmacladum in the wider Caribbean indicates nutrient enrichment. Harmful Algae 4: 10921105.CrossRefGoogle Scholar
Larkum, A. W. D., Kendrick, G. A., Ralph, P. J. (2018). Seagrasses of Australia: Structure, Ecology and Evolution. Springer, Cham.CrossRefGoogle Scholar
Laws, E. A., Pei, S. F., Bienfang, P. et al. (2011). Phosphate-limited growth and uptake kinetics of the marine prasinophyte Tetraselmis suecia (Kylin) Butcher. Aquaculture 322: 117121.CrossRefGoogle Scholar
Lehninger, A. L., Nelson, D. L. & Cox, M. M. (1993). Principles of Biochemistry. 2nd ed. Worth Publishers, New York.Google Scholar
Levitan, O., Dinamarca, J., Zelzion, E. et al. (2015). Remodeling of intermediate metabolism in the diatom Phaeodactylum tricornutum under nitrogen stress. Proceedings of the National Academy of Sciences USA 112: 412417.CrossRefGoogle ScholarPubMed
Liefer, J. D., Garg, A., Fyfe, M. H. et al. (2019). The macromolecular basis of phytoplankton C:N:P under nitrogen starvation. Frontiers in Microbiology 10: 763. https://doi.org/10.3389/fmicb.2019.00763.CrossRefGoogle ScholarPubMed
Lin, S., Zhang, Y., Zhang, H. et al. (2015). The Symbiodinium kawagutii genome illuminates dinoflagellate gene expression and coral symbiosis. Science 350: 691694.CrossRefGoogle ScholarPubMed
Lin, S., Litaker, R. W. & Sunda, W. G. (2016). Phosphorus physiological ecology and molecular mechanisms in marine phytoplankton. Journal of Phycology 52: 1036.CrossRefGoogle ScholarPubMed
Litchman, E., Klausmeier, C. A., Schofield, O. M. et al. (2007). The role of functional traits and trade-offs in structuring phytoplankton communities: Scaling from cellular to ecosystem level. Ecology Letters 10: 11701181.CrossRefGoogle ScholarPubMed
Litchman, E. & Klausmeier, C. A. (2008). Trait-based community ecology of phytoplankton. Annual Review of Ecology, Evolution, and Systematics 39: 615639.CrossRefGoogle Scholar
Lobus, N. V. & Kulikovskiy, M. S. (2023). The co-evolution aspects of the biogeochemical role of phytoplankton in aquatic ecosystems: A review. Biology 12: 92. https://doi.org/10.3390/biology12010092.CrossRefGoogle ScholarPubMed
Loladze, I. & Elser, J. (2011). The origins of the Redfield nitrogen-to phosphorus ratio are in a homoeostatic protein-to-rRNA ratio. Ecology Letters 14: 244250.CrossRefGoogle Scholar
Lomas, M. W. & Glibert, P. M. (1999). Temperature regulation of nitrate uptake: A novel hypothesis about nitrate uptake and reduction in cool-water diatoms. Limnology and Oceanography 44: 556572.CrossRefGoogle Scholar
Longworth, J., Wu, D., Huete-Ortega, M. et al. (2016). Proteome response of Phaeodactylum tricornutum, during lipid accumulation induced by nitrogen depletion. Algal Research 18: 213224.CrossRefGoogle ScholarPubMed
Lopez-Ruiz, A., Verbelen, J. P., Roldan, J. M. et al. (1985). Nitrate reductase of green algae is located in the pyrenoid. Plant Physiology 79: 10061010.CrossRefGoogle ScholarPubMed
Luo, Y. W., Doney, S. C., Anderson, L. A. et al. (2012). Database of diazotrophs in global ocean: Abundance, biomass and nitrogen fixation rates. Earth System Science Data 4: 4773.CrossRefGoogle Scholar
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. https://doi.org/10.1016/j.scitotenv.2020.137837.CrossRefGoogle ScholarPubMed
Mackereth, F. J. (1953). Phosphorus utilization by Asterionella formosa Hass. Journal of Experimental Botany 4: 296313.CrossRefGoogle Scholar
Maloney, B., Iliffe, T. M., Gelwick, F. et al. (2011). Effect of nutrient enrichment on naturally occurring algal species in six cave pools of Bermuda. Phycologia 50: 132143.CrossRefGoogle Scholar
Martin, P., Van Mooy, B. A. S., Heithoff, A. et al. (2011). Phosphorus supply drives rapid turnover of membrane phospholipids in the diatom Thalassiosira pseudonana. ISME Journal 5: 10571060.CrossRefGoogle ScholarPubMed
Martin, R. E. & Quigg, A. (2012). Evolving phytoplankton stoichiometry fueled diversification of the marine biosphere. Geosciences. Special Issue on Paleontology and Geo/Biological Evolution 2: 130146.Google Scholar
Martin, R. E. & Quigg, A. (2013). Tiny plants that once ruled the seas. Scientific American 308: 4045.CrossRefGoogle ScholarPubMed
Martinez-Perez, C., Mohr, W., Loscher, C. R. et al. 2016. The small unicellular diazotrophic symbiont, UCYNA, is a key player in the marine nitrogen cycle. Nature Microbiology 1: 16163.CrossRefGoogle ScholarPubMed
Martiny, A. C., Coleman, M. L. & Chisholm, S. W. (2006). Phosphate acquisition genes in Prochlorococcus ecotypes: Evidence for genome-wide adaptation. Proceedings of the National Academy of Sciences USA 103: 552–12.CrossRefGoogle ScholarPubMed
Martiny, A. C., Kathuria, S. & Berube, P. M. (2009). Widespread metabolic potential for nitrite and nitrate assimilation among Prochlorococcus ecotypes. Proceedings of the National Academy of Sciences USA 106: 1078710792.CrossRefGoogle ScholarPubMed
Martiny, A. C., Vrugt, J. A. & Lomas, M. W. (2014). Concentrations and ratios of particulate organic carbon, nitrogen, and phosphorus in the global ocean. Scientific Data 1: 140048. https://doi.org/10.1038/sdata.2014.48.CrossRefGoogle ScholarPubMed
McInnes, A. S., Shepard, A. K., Raes, E. J. et al. (2014). Simultaneous quantification of active carbon and nitrogen fixing communities and estimation of rates using fluorescence in situ hybridization and flow cytometry. Applied and Environmental Microbiology 80: 67506759.CrossRefGoogle ScholarPubMed
McKew, B. A., Metodieva, G., Raines, C. A. et al. (2015). Acclimation of Emiliania huxleyi (1516) to nutrient limitation involves precise modification of the proteome to scavenge alternative sources of N and P. Environmental Microbiology 17: 40504062.CrossRefGoogle ScholarPubMed
Mills, M. M., Ridame, C., Davie, M. et al. (2004). Iron and phosphorus co-limit nitrogen fixation in the eastern tropical North Atlantic. Nature 429: 292294.CrossRefGoogle ScholarPubMed
Mock, T., Samanta, M. P., Iverson, V. et al. (2008). Whole-genome expression profiling of the marine diatom Thalassiosira pseudonana identifies genes involved in silicon bioprocesses. Proceedings of the National Academy of Sciences USA 105: 15791584.CrossRefGoogle ScholarPubMed
Mock, T. (2021). Silicon drives the evolution of complex crystal morphology in calcifying algae. New Phytologist 231: 18451857.CrossRefGoogle ScholarPubMed
Monod, J. (1942). Recherches sur la croissance des cultures bactériennes, 2nd ed. Hermann, Paris.Google 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., Hudson, R. J. M. & Price, N. M. (1991). Limitation of productivity by trace metals in the sea. Limnology and Oceanography 36: 17421755.CrossRefGoogle Scholar
Morel, F. M. M. & Price, N. M. (2003). The biogeochemical cycles of trace metals in the oceans. Science 300: 944947.CrossRefGoogle ScholarPubMed
Morey, J. S., Monroe, E. A., Kinney, A. L. et al. (2011). Transcriptomic response of the red tide dinoflagellate, Karenia brevis, to nitrogen and phosphorus depletion and addition. BMC Genomics 12: 346. https://doi.org/10.1186/1471–2164–12–346.CrossRefGoogle ScholarPubMed
Moseley, J. L., Chang, C.-W. & Grossman., A. R. (2006). Genome-based approaches to understanding phosphorus deprivation responses and PSR1 control in Chlamydomonas reinhardtii. Eukaryotic Cell 5: 2644.CrossRefGoogle ScholarPubMed
Mữnoz-Blanco, J., Moyano, E. & Cardenas, J. (1990). Extracellular deamination of amino acids by Chlamydomonas reinhardtii cells. Planta 182: 194198.CrossRefGoogle ScholarPubMed
Nayar, S., Loo, M. G. K., Tanner, J. E. et al. (2018). Nitrogen acquisition and resource allocation strategies in temperate seagrass Zostera nigricaulis: Uptake, assimilation and translocation processes. Scientific Reports 8: 17151.CrossRefGoogle ScholarPubMed
Nelson, D. M., Tréguer, P., Brzezinski, M. A. et al. (1995). Production and dissolution of biogenic silica in the ocean – revised global estimates, comparison with regional data and relationship to biogenic sedimentation. Global Biogeochemical Cycles 9: 359372.CrossRefGoogle Scholar
Navarro, M. T., Prieto, R., Fernandez, E. et al. (1996). Constitutive expression of nitrate reductase changes the regulation of nitrate and nitrite transporters in Chlamydomonas reinhardii. Plant Journal 9: 819827.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. 225272.CrossRefGoogle Scholar
Orchard, E. D., Benitez-Nelson, C. R., Pellechia, P. J. et al. (2010). Polyphosphate in Trichodesmium from the low-phosphorus Sargasso Sea. Limnology and Oceanography 55: 21612169.CrossRefGoogle Scholar
Ottosen, L. D. M., Risgaard-Petersen, N. & Nielsen, L. P. (1999). Direct and indirect measurements of nitrification and denitrification in the rhizosphere of aquatic macrophytes. Aquatic Microbiological Ecology 19: 8191.CrossRefGoogle Scholar
Paasche, E. (1973). Silicon and the ecology of marine plankton diatoms. II. Silicate-uptake kinetics in five diatom species. Marine Biology 19: 262269.CrossRefGoogle Scholar
Paerl, H. W., Rudek, J. & Mallin, M. A. (1987). Limitation of N2 fixation in coastal marine waters: Relative importance of molybdenum, iron, phosphorus and organic matter availability. Limnology and Oceanography 32: 525536.CrossRefGoogle Scholar
Palenik, B. (2014). Molecular mechanisms by which marine phytoplankton respond to their dynamic chemical environment. Annual Review of Marine Science 7: 325340.CrossRefGoogle ScholarPubMed
Parslow, J. S., Harrison, P. J. & Thompson, P. A. (1984). Saturated uptake kinetics: Transient response of the marine diatom Thalassiosira pseudonana to ammonium, nitrate, silicate or phosphate starvation. Marine Biology 83: 5159.CrossRefGoogle Scholar
Parsons, T. R., Stephens, K. & Strickland, J. D. H. (1961). On the chemical composition of eleven species of marine phytoplankters. Journal of the Fisheries Research Board of Canada 18: 10011016.CrossRefGoogle Scholar
Pausch, F., Bischof, K. & Trimborn, S. (2019). Iron and manganese co-limit growth of the Southern Ocean diatom Chaetoceros debilis. PLOS ONE 14: e0221959.CrossRefGoogle ScholarPubMed
Pederson, U., Jørgensen, L. B. & Sand-Jensen, K. (1997). Through-flow of water in leaves of a submerged plant is influenced by the apical opening. Planta 202: 4350.CrossRefGoogle Scholar
Perry, M. J. (1972). Alkaline phosphatase activity in subtropical Central North Pacific waters using a sensitive fluorometric method. Marine Biology 15: 113119.CrossRefGoogle Scholar
Phillips, J. C. & Hurd, C. L. (2003). Nitrogen ecophysiology of intertidal seaweeds from New Zealand: N uptake, storage and utilization in relation to shore position and season. Marine Ecology Progress Series 264: 3148.CrossRefGoogle Scholar
Pitt, F. D., Mazard, S., Humphreys, L. et al. (2010). Functional characterization of Synechocystis sp. Strain PCC 6803 pst1 and pst2 gene clusters reveals a novel strategy for phosphate uptake in a freshwater cyanobacterium. Journal of Bacteriology 192: 35123523CrossRefGoogle Scholar
Pirc, H. (1985). Growth dynamics in Posidonia oceanica (L.) Delile. I. Seasonal changes of soluble carbohydrates, starch, free amino acids, nitrogen and organic anions in different parts of the plant. Marine Ecology 6: 141165.CrossRefGoogle Scholar
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. (2008). Trace elements. In: Jørgensen, S. E. & Fath, B. D. (eds.) Ecological Stoichiometry in the Encyclopedia of Ecology, Vol. 5. Elsevier, Oxford, pp. 35643573.CrossRefGoogle Scholar
Quigg, A., Irwin, A. J. & Finkel, Z. V. (2011). Evolutionary inheritance of elemental stoichiometry in phytoplankton. Proceedings of the Royal Society: Biological Sciences 278: 526534.Google ScholarPubMed
Quigg, A., Al-Anasi, M., Nour El Din, N. et al. (2013). Phytoplankton along the coastal shelf of an oligotrophic hypersaline environment in a semi-enclosed marginal sea: Qatar (Arabian Gulf). Continental Shelf Research 60: 116.CrossRefGoogle Scholar
Quigg, A. (2016). Micronutrients. In: Borowitzka, M. A., Beardall, J. & Raven, J. A. (eds.) The Physiology of Microalgae. Developments in Applied Phycology Series, Vol. 6. Springer, Dordrecht, pp. 211231.Google Scholar
Rascio, N. & La Rocca, N. (2013). Biological Nitrogen Fixation, Reference Module in Earth Systems and Environmental Sciences. Elsevier, Amsterdam. https://doi.org/10.1016/B978–0-12–409548–9.09470–7.Google Scholar
Raven, J. A. & Smith, F. A. (1976). Nitrogen assimilation and transport in vascular land plants in relation to intracellular pH regulation. New Phytologist 76: 415431.CrossRefGoogle Scholar
Raven, J. A. (1980). Nutrient transport in microalgae. Advances in Microbiology and Physiology 21: 47226.CrossRefGoogle ScholarPubMed
Raven, J. A. (1981). Nutrient strategies of submerged benthic plants: The acquisition of C, N and P by rhizophytes and haptophytes. New Phytologist 88: 130.CrossRefGoogle Scholar
Raven, J. A. (1984). Energetics and Transport in Aquatic Plants. A. R. Liss Inc., New York.Google Scholar
Raven, J. A. (1985). Regulation of pH and generation of osmolarity in vascular plants: A cost-benefit analysis in relation to efficiency of use of energy, nitrogen and water. New Phytologist 101: 2577.CrossRefGoogle ScholarPubMed
Raven, J. A. & Farquhar, G. D. (1990). The influence of N metabolism and organic acid synthesis on the natural abundance of isotopes of carbon in plants. New Phytologist 116: 505529.CrossRefGoogle Scholar
Raven, J. A., Wollenweber, B. & Handley, L. L. (1992). A comparison of ammonium and nitrate as nitrogen sources for photolithotrophs. New Phytologist 121: 1932. https://doi.org/10.1111/nph.1992.121.issue-1.CrossRefGoogle Scholar
Raven, J. A., Evans, M. C. W. & Korb, R. E. (1999). The role of trace metals in photosynthetic electron transport in O2-evolving organisms. Photosynthesis Research 60: 111149.CrossRefGoogle Scholar
Raven, J. A. & Knoll, A. H. (2010). Non-skeletal biomineralization by eukaryotes: Matters of moment and gravity. Geomicrobiology Journal 27: 572584.CrossRefGoogle Scholar
Raven, J. A. (2013a). RNA function and phosphorus use by photosynthetic organisms. Frontiers in Plant Science 4: 536. https://doi.org/10.3389/fpls.2013.00536.CrossRefGoogle ScholarPubMed
Raven, J. A. (2013b). The evolution of autotrophy in relation to phosphorus requirement. Journal of Experimental Botany 64: 40234046.CrossRefGoogle ScholarPubMed
Raven, J. A. (2017). Evolution and palaeophysiology of the vascular system and other means of long distance transport. Philosophical Transactions of the Royal Society B 373: 20160497.CrossRefGoogle Scholar
Raven, J. A., Knight, C. A. & Beardall, J. (2019). Cell size has gene expression and biophysical consequences for cellular function. Perspectives in Phycology 6: 8194.CrossRefGoogle 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 United Kingdom 100: 333346.CrossRefGoogle Scholar
Raven, J. A., Beardall, J. & Quigg, A. (2020). Light-driven oxygen consumption in the water-water cycles and photorespiration, and light stimulated mitochondrial respiration. In: Larkum, A. W. D., Raven, J. A. & Douglas, S. (eds.) Photosynthesis in the Algae. Biochemical and Physiological Mechanisms. Advances in Photosynthesis and Respiration 45, Springer Nature Publishing, Cham, pp. 161178. https://doi.org/10.1007/978-3-030-33397-3_8.CrossRefGoogle Scholar
Redfield, A. C. (1934). On the proportions of organic derivatives in sea water and their relation to the composition of plankton. In: Daniel, R. J. (ed.) James Johnstone Memorial Volume. Liverpool University Press, Liverpool, UK, pp. 176192.Google Scholar
Rees, T. A. V. (2007). Metabolic and ecological constraints imposed by similar rates of ammonium and nitrate uptake per unit surface area at low substrate concentrations in marine phytoplankton and macroalgae. Journal of Phycology 43: 197207.Google Scholar
Rees, T. A. V. & Raven, J. A. (2021). The maximum growth rate hypothesis is correct for eukaryotic photosynthetic organisms but not cyanobacteria. New Phytologist 230: 601611.CrossRefGoogle Scholar
Reid, R. J., Mimura, T., Ohsumi, Y. et al. (2000). Phosphate uptake in Chara: Membrane transport via Na/Pi cotransport. Plant Cell and Environment 23: 223228.CrossRefGoogle Scholar
Rhee, G.-Y. (1978). Effects of N:P atomic ratios and nitrate limitation on algal growth, cell composition, and nitrate uptake. Limnology and Oceanography 23: 1025.CrossRefGoogle Scholar
Richardson, T. L. & Jackson, G. A. (2007). Small phytoplankton and carbon export from the surface ocean. Science 315: 838840.CrossRefGoogle ScholarPubMed
Risgaard-Petersen, N. & Jensen, K. (1997). Nitrification and denitrification in the rhizosphere of the aquatic macrophyte Lobelia dortmanna L. Limnology and Oceanography 42: 529537.CrossRefGoogle Scholar
Roberts, S. C. (1997). Physiological effects of phosphorus limitation on photosynthesis in two green algae. PhD thesis, Monash University, Melbourne, Australia. pp. 116.Google Scholar
Roberts, S., Shelly, K. & Beardall, J. (2008). Interactions among phosphate uptake, photosynthesis, and chlorophyll fluorescence in nutrient-limited cultures of the chlorophyte microalga Dunaliella tertiolecta. Journal of Phycology 44: 662669.CrossRefGoogle ScholarPubMed
Rogato, A., Amato, A., Iudicone, D. et al. (2015). The diatom molecular toolkit to handle nitrogen uptake. Marine Genomics 24: 95108.CrossRefGoogle ScholarPubMed
Roleda, M. Y. & Hurd, C. L. (2019). Seaweed nutrient physiology: Application of concepts to aquaculture and bioremediation. Phycologia 58: 552562.CrossRefGoogle Scholar
Rose, A., Padovan, A., Christian, K. et al. (2021). The diversity of nitrogen-cycling microbial genes in a waste stabilization pond reveals changes over space and time that is uncoupled to changing nitrogen chemistry. Microbial Ecology 8: 1102911041.Google Scholar
Rosenburg, G. & Ramus, J. (1984). Uptake of inorganic nitrogen and seaweed surface area: volume ratios. Aquatic Botany 19: 6572.CrossRefGoogle Scholar
Roth, N. C. & Pregnall, A. M. (1988). Nitrate reductase activity in Zostera marina. Marine Biology 99: 457463.CrossRefGoogle Scholar
Sakshaug, E., Granéli, E., Elbrächter, M. et al. (1984). Chemical composition and alkaline phosphatase activity of nutrient-saturated and P-deficient cells of four marine dinoflagellates. Journal of Experimental Marine Biology and Ecology 77: 241254.CrossRefGoogle Scholar
Sand-Jensen, K., Martinson, T. K., Jakobs, A. L. et al. (2021). Large pools and fluxes of carbon, calcium and phosphorus in dense charophyte stands in ponds. Science of the Total Environment 765: 142792. https://doi.org/10.1016/j.scitoenv.2020.142792.CrossRefGoogle ScholarPubMed
Sanchez-Baracaldo, P., Ridgwell, A. & Raven, J. A. (2014). A neoproterozoic transition in the marine nitrogen cycle. Current Biology 6: 652657.CrossRefGoogle Scholar
Sañudo-Wilhelmy, S. A., Kustka, A. B., Gobler, C. J. et al. (2001). Phosphorus limitation of nitrogen fixation by Trichodesmium in the central Atlantic Ocean. Nature 411: 6669.CrossRefGoogle ScholarPubMed
Sapriel, G., Quinet, M., Heijde, M. et al. (2009). Genome-wide transcriptome analyses of silicon metabolism in Phaeodactylum tricornutum reveal the multilevel regulation of silicic acid transporters. PLOS ONE 4: e7458. https://doi.org/10.1371/journal.pone.0007458.CrossRefGoogle ScholarPubMed
Scanlan, D. J. & Post, A. F. (2008). Aspects of marine cyanobacterial nitrogen physiology and connection to the nitrogen cycle. In: Capone, D. G., Bronk, D. A., Mulholland, M. R. & Carpenter, E. J. (eds.) Nitrogen in the Marine Environment. Elsevier, Amsterdam, pp. 10731096.CrossRefGoogle Scholar
Scanlan, D. J., Ostrowski, M., Mazard, S. et al. (2009). Ecological genomics of marine picocyanobacteria. Microbiology and Molecular Biology Reviews 73: 249299.CrossRefGoogle ScholarPubMed
Schindler, D. W. (1975). Whole-lake eutrophication experiments with phosphorus, nitrogen and carbon. Internationale Vereinigung für Theoretische und Angewandte Limnologie: Verhandlungen 19: 32213231.Google Scholar
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
Sculthorpe, C. D. (1967). The Biology of Aquatic Vascular Plants. Edward Arnold, London.Google Scholar
Short, F. T., Dennison, W. C. & Capone, D. G. (1990) Phosphorus-limited growth of the tropical seagrass Syringodium filiforme in carbonate sediments. Marine Ecology Progress Series 62: 169174.CrossRefGoogle Scholar
Shrestha, R. P., Tesson, B., Krichmar, T. N. et al. (2012). Whole transcriptome analysis of the silicon response of the diatom Thalassiosira pseudonana. BMC Genomics 13: 499.CrossRefGoogle ScholarPubMed
Shrestha, R. P. & Hildebrand, M. (2015). Evidence for a regulatory role of diatom silicon transporters in cellular silicon responses. Eukaryotic Cell 14: 2940.CrossRefGoogle ScholarPubMed
Siaut, M., Heijde, M., Mangogna, M. et al. (2007). Molecular toolbox for studying diatom biology in Phaeodactylum tricornutum. Gene 406: 2335.CrossRefGoogle ScholarPubMed
Smith, V. H. (1983). Low nitrogen to phosphorus ratios favor dominance by blue-green algae in lake phytoplankton. Science 221: 669671.CrossRefGoogle ScholarPubMed
Smith, S., Yamanaka, Y., Pahlow, M. et al. (2009). Optimal uptake kinetics: Physiological acclimation explains the patterns of nitrate uptake by phytoplankton in the ocean. Marine Ecology Progress Series 384: 112.CrossRefGoogle Scholar
Sohm, J. A., Webb, E. A. & Capone, D. G. (2011). Emerging patterns of marine nitrogen fixation. Nature Reviews in Microbiology 9: 499508.CrossRefGoogle ScholarPubMed
Solomon, C. M., Collier, J. L., Berg, G. M. et al. (2010). Role of urea in microbial metabolism in aquatic systems: A biochemical and molecular review. Aquatic Microbial Ecology 59: 6788.CrossRefGoogle Scholar
Solovchenko, A. E., Ismagulova, T. T., Lukyanov, A. A. et al. (2019). Luxury phosphorus uptake in microalgae. Journal of Applied Phycology 31: 27552770.CrossRefGoogle Scholar
Sproles, A. E., Kirk, N. L. & Kitchen, S. A. (2018). Phylogenetic characterization of transporter proteins in the cnidarian-dinoflagellate symbiosis. Molecular and Phylogenetic Evolution 120: 307320.CrossRefGoogle ScholarPubMed
Sterner, R. W. & Elser, J. J. (2002). Ecological Stoichiometry: The Biology of the Elements From Molecules to the Biosphere. Princeton University Press, Princeton.Google Scholar
Staal, M. F., Meysman, J. R. & Staal, L. J. (2003). Temperature excludes N2-fixing heterocystous cyanobacteria in the tropical oceans. Nature 425: 504507.CrossRefGoogle ScholarPubMed
Stapel, J., Aerts, T. L., Dutnhoven, B. H. M. et al. (1999). Nutrient uptake by leaves and roots of the seagrass, Thalassia hemprichii in the Spermods Archipelago, Indonesia. Marine Ecology Progress Series 134: 195206.CrossRefGoogle Scholar
Su, Z., Olman, V., Mao, F. et al. (2005). Comparative genomics analysis of NtcA regulons in cyanobacteria: Regulation of nitrogen assimilation and its coupling to photosynthesis. Nucleic Acids Research 33: 51565171.CrossRefGoogle ScholarPubMed
Suggett, D. J., Warner, M. E. & Leggat, W. (2017). Symbiotic dinoflagellate functional diversity mediates coral survival under ecological crisis. Trends in Ecology and Evolution 32: 735745.CrossRefGoogle ScholarPubMed
Sunda, W. G. (1994). Trace metal/phytoplankton interactions in the sea. In: Bidoglio, G. & Stumm, W. (eds.) Chemistry of Aquatic Systems: Local and Global Perspectives. Springer, Dordrecht, The Netherlands, pp. 213247.CrossRefGoogle Scholar
Sylvan, J. B., Quigg, A., Tozzi, S. et al. (2011). Mapping phytoplankton community physiology on a river impacted continental shelf: Testing a multifaceted approach. Estuaries and Coasts 34: 12201233.CrossRefGoogle Scholar
Tabita, F. R., Hanson, T. E., Li, H. et al. (2007). Function, structure, and evolution of the RubisCO-like proteins and their RubisCO homologs. Microbiology and Molecular Biology Reviews 71: 576599.CrossRefGoogle ScholarPubMed
Takabayashi, M., Wilkerson, F. P. & Robertson, D. L. (2005). Response of glutamine synthetase gene transcription and enzyme activity to external nitrogen sources in the diatom Skeletonema costatum (Bacillariophyceae). Journal of Phycology 41: 8494.CrossRefGoogle Scholar
Tanaka, R. & Tanaka, A. (2007). Tetrapyrrole biosynthesis in higher plants. Annual Review of Plant Biology 58: 321346.CrossRefGoogle ScholarPubMed
Taylor, M. W. & Rees, T. A. V. (1999). Kinetics of ammonium assimilation in two seaweeds, Enteromorpha sp. (Chlorophyceae) and Osmundaria colensoi (Rhodophyceae). Journal of Phycology 35: 740746.CrossRefGoogle Scholar
Thamatrakoln, K. & Hildebrand, M. (2008). Silicon uptake in diatoms revisited: A model for saturable and nonsaturable uptake kinetics and the role of silicon transporters. Plant Physiology 146: 13971407.CrossRefGoogle Scholar
Thompson, A. W., Huang, K., Saito, M. A. et al. (2011). Transcriptome response of high- and low-light adapted Prochlorococcus strains to changing iron availability. The ISME Journal 5: 15801594.CrossRefGoogle ScholarPubMed
Tilman, D. (1982). Resource Competition and Community Structure. Princeton University Press, Princeton.Google ScholarPubMed
Tolonen, A. C., Aach, J., Lindell, D. et al. (2006). Global gene expression of Prochlorococcus ecotypes in response to changes in nitrogen availability. Molecular Systems Biology 2: 53. https://doi.org/10.1038/msb4100087.CrossRefGoogle ScholarPubMed
Touchette, B. W. & Burkholder, J. M. (2000) Reviews of nitrogen and phosphorus metabolism in seagrasses. Journal of Experimental Marine Biology and Ecology 250: 133167.CrossRefGoogle Scholar
Tréguer, P., Bowler, C., Moriceau, B. et al. (2018). Influence of diatom diversity on the ocean biological carbon pump. Nature Geoscience 11: 2737.CrossRefGoogle Scholar
Twilley, R. R., Chen, R. H. & Hargis, T. (1992). Carbon sinks in mangroves and their implications to carbon budget of tropical coastal ecosystems. Water Air and Soil Pollution 64: 265265.CrossRefGoogle Scholar
Twining, B. S. & Baines, S. B. (2013). The trace metal composition of marine phytoplankton. Annual Review of Marine Science 5: 191215.CrossRefGoogle ScholarPubMed
Tyrrell, T. (1999). The relative influences of nitrogen and phosphorus on oceanic primary production. Nature 400: 525531.CrossRefGoogle Scholar
Van de Waal, D. B. & Litchman, E. (2020). Multiple global change stressor effects on phytoplankton nutrient acquisition in a future ocean. Philosophical Transactions of the Royal Society B 375: 20190706. https://doi.org/10.1098/rstb.2019.0706.CrossRefGoogle Scholar
Van Mooy, B. A., 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
Vermeer, C. P., Escher, M., Portielje, R. et al. (2003). Nitrogen uptake and translocation by Chara. Aquatic Botany 76: 245258.CrossRefGoogle Scholar
Vitousek, P. M. & Howarth, R. W. (1991). Nitrogen limitation on land and in the sea: How can it occur? Biogeochemistry 13: 87115.CrossRefGoogle Scholar
Vitousek, P. M., Menge, D. N. L., Reed, S. C. et al. (2013). Biological nitrogen fixation: Rates, patterns and ecological controls in terrestrial ecosystems. Philosophical Transactions of the Royal Society B 368: 20130119. https://doi.org/10.1098/rstb.2013.0119.CrossRefGoogle ScholarPubMed
Vogels, G. D. & Van Der Drift, C. (1976). Degradation of purines and pyrimidines by microorganisms. Bacteriology Review 40: 403468.CrossRefGoogle ScholarPubMed
Voss, M., Bange, H. W., Dippner, J. W. et al. (2013). The marine nitrogen cycle: Recent discoveries, uncertainties and the potential relevance of climate change. Philosophical Transactions of the Royal Society B 368: 20130121. https://doi.org/10.1098/rstb.2013.0121.CrossRefGoogle ScholarPubMed
Walker, N. A., Reid, R. J. & Smith, F. A. (1993). The uptake and metabolism of urea by Chara australis. IV. Symport with sodium – a slip model for the high and low affinity systems. Journal of Membrane Biology 136: 263271.CrossRefGoogle ScholarPubMed
Waraich, E. A., Amad, R., Ashraf, M. Y. et al. (2011). Improving agricultural water use efficiency by nutrient management. Acta Agriculturae Scandinavica – Soil & Plant Science 61: 291304.Google Scholar
Wang, X., Huang, B. & Zhang, H. (2014). Phosphorus deficiency affects multiple macromolecular biosynthesis pathways of Thalassiosira weissflogii. Acta Oceanology Sinica 33: 8591.CrossRefGoogle Scholar
Waycott, M., Duarte, C. M., Carruthers, T. J. B. et al. (2009). Accelerating loss of seagrasses across the globe threatens coastal ecosystems. Proceedings of the National Academy of Sciences USA 106: 1237712381. https://doi.org/10.1073/pnas.0905620106.CrossRefGoogle ScholarPubMed
Wetz, M. S., Cira, E. K., Sterba-Boatwright, B. et al. (2017). Exceptionally high organic nitrogen concentrations in a semi-arid South Texas estuary susceptible to brown tide blooms. Estuarine, Coastal and Shelf Science 188: 2737.CrossRefGoogle Scholar
Wilkerson, F. P. & Trench, R. K. (1986). Uptake of dissolved inorganic nitrogen by the symbiotic clam Tridacna gigas and the coral Acropora sp. Marine Biology 93: 237246.CrossRefGoogle Scholar
Wilkerson, F. P., Dugdale, R. C., Hogue, V. E. et al. (2006). Phytoplankton blooms and nitrogen productivity in the San Francisco Bay. Estuaries Coasts 29: 401416.CrossRefGoogle Scholar
Williams, S. K. & Hodson, R. C. (1977). Transport of urea at low concentrations in Chlamydomonas reinhardtii. Journal of Bacteriology 130: 266273.CrossRefGoogle Scholar
Williams, S. L. (1981). Uptake of sediment ammonium and translocation in a marine green macroalga Caulerpa cuppressoides. Limnology and Oceanography 29: 374379.CrossRefGoogle Scholar
Williams, S. L. & Fisher, T. R. (1985). Kinetics of nitrogen-15 labelled ammonium uptake by Caulerpa cupressoides (Chlorophyta). Journal of Phycology 21: 287296.CrossRefGoogle Scholar
Wong, J. C. Y., Enríquez, S. & Baker, D. M. (2021). Towards a trait-based understanding of Symbiodiniaceae nutrient acquisition strategies. Coral Reefs 40: 625639.CrossRefGoogle Scholar
Wu, J. F., Sunda, W., Boyle, E. A. et al. (2000). Phosphate depletion in the western North Atlantic Ocean. Science 289: 759762.CrossRefGoogle ScholarPubMed
Wurch, L. L., Haley, S. T., Orchard, E. D. et al. (2011). Nutrient regulated transcriptional responses in the brown tide forming alga Aureococcus anophagefferens. Environment and Microbiology 13: 468481.CrossRefGoogle ScholarPubMed
Wynne, D. & Berman, T. (1990). The influence of environmental factors on nitrate reductase activity in freshwater phytoplankton. I. Field studies. Hydrobiologia 194: 235245.CrossRefGoogle Scholar
Yacano, M. R., Foster, S. Q., Ray, N. E. et al. (2022). Marine macroalgae are an overlooked sink of silicon in coastal systems. New Phytologist 233: 23302336.CrossRefGoogle ScholarPubMed
Yamaguchi, H., Yamaguchi, M., Fukami, K. et al. (2005). Utilization of phosphate diester by the marine diatom Chaetoceros ceratosporus. Journal of Plankton Research 27: 603606.CrossRefGoogle Scholar
Yang, Z. K., Niu, Y.-F. & Ma, Y.-H. (2013). Molecular and cellular mechanisms of neutral lipid accumulation in diatom following nitrogen deprivation. Biotechnology for Biofuels 6: 67. https://doi.org/10.1186/1754–6834–6-67.CrossRefGoogle ScholarPubMed
Yang, Z. K., Ma, Y. H., Zheng, J. W. et al. (2014). Proteomics to reveal metabolic network shifts towards lipid accumulation following nitrogen deprivation in the diatom Phaeodactylum tricornutum. Journal of Applied Phycology 26: 7382.CrossRefGoogle ScholarPubMed
Yin, K., Liu, H. & Harrison, P. J. (2017). Sequential nutrient uptake as a potential mechanism for phytoplankton to maintain high primary productivity and balanced nutrient stoichiometry. Biogeosciences 14: 24692480.CrossRefGoogle Scholar
Young, E. B., Dring, M. J., Savidge, G. et al. (2007). Seasonal variations in nitrate reductase activity and internal N pools in intertidal brown algae are correlated with ambient nitrate concentrations. Plant Cell and Environment 30: 764774.CrossRefGoogle ScholarPubMed
Yu, C., Pan, Y. & Hu, H. (2023). NmrA acts as a positive regulator of nitrate assimilation in Phaeodactylum tricornutum. Algal Research 69: 102960. https://doi.org/10.1016/j.algal.2022.102960.CrossRefGoogle Scholar
Zehr, J. P. & Kudela, J. P. (2011). Nitrogen cycle of the open ocean: From genes to ecosystem. Annual Review of Marine Science 3: 197225.CrossRefGoogle Scholar
Zehr, J. P. & Capone, D. G. (2020). Changing perspectives in marine nitrogen fixation. Science 368: 9514.CrossRefGoogle ScholarPubMed
Zhang, Q., Bendif, E. M., Zhou, Y. et al. (2022). Declining metal availability in the Mesozoic seawater reflected in phytoplankton succession. Nature Geoscience 15: 932941.CrossRefGoogle Scholar
Zimmerman, A. E., Allison, S. D. & Martiny, A. C. (2014). Phylogenetic constraints on elemental stoichiometry and resource allocation in heterotrophic marine bacteria. Environmental Microbiology 16: 13981410.CrossRefGoogle ScholarPubMed

References

Allakhverdiev, S. I. & Murata, N. (2008). Salt stress inhibits photosystems II and I in cyanobacteria. Photosynthesis Research 98: 529539.CrossRefGoogle ScholarPubMed
Bird, C. J. & McLachlan, J. (1986). The effect of salinity on distribution of species of Gracilaria (Rhodophyta, Gigartinales): An experimental assessment. Botanica Marina 29: 231238.CrossRefGoogle Scholar
Bisson, M. A. & Gutknecht, J. (1975). Osmotic regulation in the marine alga, Codium decorticatum: Regulation of turgor pressure by control of ionic composition. The Journal of Membrane Biology 24:183200.CrossRefGoogle ScholarPubMed
Bisson, M. A. & Kirst, G. O. (1983). Osmotic adaptations of charophyte algae in the Coorong, South Australia and other Australian lakes. Hydrobiologia 105: 4551.CrossRefGoogle Scholar
Bisson, M. A. & Kirst, G. O. (1995). Osmotic acclimation and turgor pressure regulation in algae. Naturwissenschaften 82: 461471.CrossRefGoogle Scholar
Bolton, J. J. (1979). Estuarine adaptation in populations of Pilayella littoralis (L.) Kjellm. (Phaeophyta, Ectocarpales). Estuarine Coastal Marine Science 9: 273280.CrossRefGoogle Scholar
Chudek, J. A., Foster, R., Davison, I. R. et al. (1984). Altritol in the brown alga Himanthalia elongata. Phytochemistry 23: 10811082.CrossRefGoogle Scholar
Davison, I. R. & Reed, R. H. (1985a). Osmotic adjustment in Laminaria digitata (Phaeophyta) with particular reference to seasonal changes in internal solute concentrations. Journal of Phycology 21: 4150.CrossRefGoogle Scholar
Davison, I. R. & Reed, R. H. (1985b). The physiological significance of mannitol accumulation in brown algae: The role of mannitol as a compatible cytoplasmic solute. Phycologia 24: 449457.CrossRefGoogle Scholar
Dickson, D. M. J. & Kirst, G. O. (1986). The role of dimethylsulphoniopropionate, glycine betaine and homarine in the osmoacclimation of Platymonas subcordiformis. Planta 155: 409415.CrossRefGoogle Scholar
Eggert, A. & Karsten, U. (2010). Low molecular weight carbohydrates in red algae – an ecophysiological and biochemical perspective. In: Seckbach, J., Chapman, D. & Weber, A. (eds.) Cellular Origins, Life in Extreme Habitats and Astrobiology Red Algae in the Genomics Age. Springer, Berlin, pp. 445456.Google Scholar
Eppley, R. W. & Bovell, C. R. (1958). Sulphuric acid in Desmarestia. The Biological Bulletin 115: 101106.CrossRefGoogle Scholar
Gessner, F. & Schramm, W. (1971). Salinity: Plants In: Kinne, O. (ed.) Marine Ecology, Vol. 1(2) Environmental Factors. Wiley Interscience, London, pp. 7051083.Google Scholar
Gimmler, H., Kaaden, R., Kirchner, U. et al. (1984). The chloride sensitivity of Dunaliella parva enzymes. Zeitschrift für Pflanzenphysiologie 114: 131150.CrossRefGoogle Scholar
Groszmann, M., Osborn, H. L. & Evans, J. R. (2017). Carbon dioxide and water transport through plant aquaporins. Plant, Cell and Environment 40: 938961.CrossRefGoogle ScholarPubMed
Gustavs, L., Schumann, R., Eggert, A. et al. (2009) In vivo growth fluorometry: accuracy and limits of microalgal growth rate measurements in ecophysiological investigations. Aquatic Microbial Ecology 55: 95104.CrossRefGoogle Scholar
Hagemann, M. (2011). Molecular biology of cyanobacterial salt acclimation. FEMS Microbiology Reviews 35: 87123.CrossRefGoogle ScholarPubMed
Hay, W. W., Migdisov, A., Balukhovsky, A. N. et al. (2006). Evaporites and the salinity of the ocean during the Phanerozoic: Implications for climate, ocean circulation and life. Palaeogeography Palaeoclimatology Palaeoecology Journal 240: 346.CrossRefGoogle Scholar
Herburger, K. & Holzinger, A. (2015). Localization and quantification of callose in the streptophyte green algae Zygnema and Klebsormidium: Correlation with desiccation tolerance. Plant Cell Physiology 56: 22592270.Google ScholarPubMed
Holzinger, A. & Karsten, U. (2013). Desiccation stress and tolerance in green algae: Consequences for ultrastructure, physiological, and molecular mechanisms. Frontiers in Plant Sciences 4: 327.Google ScholarPubMed
Holzinger, A., Herburger, K., Kaplan, F. et al. (2015). Desiccation tolerance in the chlorophyte green alga Ulva compressa: Does cell wall architecture contribute to ecological success? Planta 242: 477492.CrossRefGoogle ScholarPubMed
Jacob, A., Kirst, G. O., Wiencke, C. et al. (1991). Physiological responses of the Antarctic green alga Prasiola crispa ssp. antarctica to salinity stress. Journal of Plant Physiology 139: 5762.CrossRefGoogle Scholar
Karsten, U., Wiencke, C. & Kirst, G. O. (1991a). The effect of salinity changes upon the physiology of eulittoral green macroalgae from Antarctica and Southern Chile. I. Cell viability, growth, photosynthesis and dark respiration. Journal of Plant Physiology 138: 667673.CrossRefGoogle Scholar
Karsten, U., Wiencke, C. & Kirst, G. O. (1991b). The effect of salinity changes upon the physiology of eulittoral green macroalgae from Antarctica and Southern Chile. II. Intracellular inorganic ions and organic compounds. Journal of Experimental Botany 245: 15331539.CrossRefGoogle Scholar
Karsten, U., Barrow, K. D. & King, R. J. (1993a). Floridoside, L-isofloridoside, and D-isofloridoside in the red alga Porphyra columbina. Plant Physiology 103: 485491.CrossRefGoogle ScholarPubMed
Karsten, U., West, J. A. & Ganesan, E. K. (1993b). Comparative physiological ecology of Bostrychia moritziana (Ceramiales, Rhodophyta) from freshwater and marine habitats. Phycologia 32: 401409.CrossRefGoogle Scholar
Karsten, U., West, J. A., Zuccarello, G. et al. (1994). Physiological ecotypes in the marine red alga Bostrychia radicans (Ceramiales, Rhodophyta) from the east coast of the USA. Journal of Phycology 30: 174182.CrossRefGoogle Scholar
Karsten, U., Barrow, K. D., Nixdorf, O. et al. (1996). The compatibility of unusual organic osmolytes from mangrove red algae with enzyme activity. Australian Journal of Plant Physiology 23: 577582.Google Scholar
Karsten, U. (2007). Salinity tolerance of Arctic kelps from Spitsbergen. Phycological Research 55: 257262.CrossRefGoogle Scholar
Karsten, U., Görs, S., Eggert, A. et al. (2007). Trehalose, digeneaside and floridoside in the Florideophyceae (Rhodophyta) – a re-evaluation of its chemotaxonomic value. Phycologia 46: 143150.CrossRefGoogle Scholar
Karsten, U. (2012). Seaweed acclimation to salinity and desiccation stress. In: Wiencke, C. & Bischof, K. (eds.) Seaweed Ecophysiology and Ecology. Ecological Studies, vol. 219, Springer, Berlin, pp. 87107.Google Scholar
Kauss, H. (1987). Some aspects of calcium-dependent regulation in plant metabolism. Annual Review of Plant Physiology 38: 4772.CrossRefGoogle Scholar
Kirst, G. O. (1990). Salinity tolerance of eukaryotic marine algae. Annual Review in Plant Physiology and Plant Molecular Biology 41: 2153.CrossRefGoogle Scholar
Kirst, G. O. & Wiencke, C. (1995). Ecophysiology of polar algae. Journal of Phycology 31: 181199.CrossRefGoogle Scholar
Kremer, B. P. (1976). Distribution and biochemistry of alditols in the genus Pelvetia (Phaeophyceae, Fucales). British Phycological Journal 11: 239243.CrossRefGoogle Scholar
Kremer, B. P. (1978). Patterns of photoassimilatory products in Pacific Rhodophyceae. Canadian Journal of Botany 56: 16551659.CrossRefGoogle Scholar
Lang, I., Sassmann, S., Schmidt, B. et al. (2014). Plasmolysis: Loss of turgor and beyond. Plants 3: 583593.CrossRefGoogle ScholarPubMed
Maathuis, F. J. M. & Amtmann, A. (1999). K+ nutrition and Na+ toxicity: The basis of cellular K+/Na+ ratios. Annals of Botany 84: 123133.CrossRefGoogle Scholar
Meng, J., Rosell, K. G. & Srivastava, L. M. (1987). Chemical characterization of floridosides from Porphyra perforata. Carbohydrate Research 161: 171180.CrossRefGoogle Scholar
Mostaert, A. S. & King, R. J. (1993). The cell wall of the halotolerant red alga Caloglossa leprieurii (Montagne) J. Agardh (Ceramiales, Rhodophyta) from freshwater and marine habitats: Effect of changing salinity. Cryptogamic Botany 4: 4046.Google Scholar
Nygard, C. A. & Dring, M. J. (2008). Influence of salinity, temperature and dissolved inorganic carbon and nutrient concentration on the photosynthesis and growth on Fucus vesiculosus from Baltic and Irish Seas. European Journal of Phycology 43: 253262.CrossRefGoogle Scholar
Oren, A. (2005). A hundred years of Dunaliella research: 1905–2005. Saline Systems 1: 2.CrossRefGoogle ScholarPubMed
Raven, J. A. (2020). Chloride involvement in the synthesis, functioning and repair of the photosynthetic apparatus in vivo. New Phytologist 227: 334342.CrossRefGoogle ScholarPubMed
Raven, J. A. & Beardall, J. (2020). Energizing the plasmalemma of photosynthetic organisms: The role of primary active transport. Journal of the Marine Biological Association UK 100: 333346.CrossRefGoogle Scholar
Reed, R. H. (1984). The effects of extreme hyposaline stress upon Polysiphonia lanosa (L.) Tandy from marine and estuarine sites. Journal of Marine Biology and Ecology 76: 131144.CrossRefGoogle Scholar
Rietema, H. (1993). Ecotypic differences between Baltic and North Sea populations of Delesseria sanguinea and Membranoptera alata. Botanica Marina 36: 1521.CrossRefGoogle Scholar
Ritchie, R. J. (1988). The ionic relations of Ulva lactuca. Journal of Plant Physiology 33: 183192.CrossRefGoogle Scholar
Roberts, M. F. (2005). Organic compatible solutes of halotolerant and halophilic microorganisms. Saline Systems 1: 5.CrossRefGoogle ScholarPubMed
Rueness, J. & Kornfeldt, R. A. (1992). Ecotypic differentiation in salinity responses of Ceramium strictum (Rhodophyta) from Scandinavian waters. Sarsia 77: 207212.CrossRefGoogle Scholar
Russell, G. (1987). Salinity and seaweed vegetation. In: Crawford, R. M. M. (ed.) The Physiological Ecology of Amphibious and Intertidal Plants. Blackwell, Oxford, pp. 3552.Google Scholar
Satoh, K., Smith, C. M. & Fork, D. C. (1983). Effects of salinity on primary processes of photosynthesis in the red alga Porphyra perforata. Plant Physiology 73: 643647.CrossRefGoogle ScholarPubMed
Smiatek, J., Harishchandra, R. K., Rubner, O. et al. (2012). Properties of compatible solutes in aqueous solution. Biophysical Chemistry 160: 6268.CrossRefGoogle ScholarPubMed
Smith, C. M., Satoh, K. & Fork, D. C. (1986). The effects of osmotic tissue dehydration and air drying in morphology and energy transfer in two species of Porphyra. Plant Physiology 80: 843847.CrossRefGoogle ScholarPubMed
Timasheff, S. N. (2002). Protein hydration, thermodynamic binding, and preferential hydration. Biochemistry 41: 1347313482.CrossRefGoogle ScholarPubMed
Turesson, G. (1922). The genotypical response of the plant species to the habitat. Hereditas 3: 211350.CrossRefGoogle Scholar
Verret, F., Wheeler, G., Taylor, A. R. et al. (2010). Calcium channels in photosynthetic eukaryotes: Implications for evolution of calcium-based signalling. New Phytologist 187: 2343.CrossRefGoogle ScholarPubMed
Wiencke, C. & Läuchli, A. (1980). Growth, cell volume, and fine structure of Porphyra umbilicalis in relation to osmotic tolerance. Planta 150: 303311.CrossRefGoogle ScholarPubMed
Wiencke, C. (1982). Effect of osmotic stress on thylakoid fine structure in Porphyra umbilicalis. Protoplasma 111: 215220.CrossRefGoogle Scholar
Wright, P. J., Clayton, M. N., Chudek, J. A. et al. (1987). Low molecular weight carbohydrates in marine brown algae from the southern hemisphere: The occurrence of altritol in Bifurcariopsis capensis, Hormosira banksii, Notheia anomala and Xiphophora chondrophylla. Phycologia 26: 429434.CrossRefGoogle Scholar
Wright, D. G., Pawlowicz, R., McDougall, T. J. et al. (2010). Absolute salinity, ‘density salinity’ and the reference-composition salinity scale: Present and future use in the seawater standard TEOS-10. Ocean Science Discussions 7: 15591625.Google Scholar
Yancey, P. H. (2005). Organic osmolytes as compatible, metabolic and counteracting cytoprotectants in high osmolarity and other stresses. Journal of Experimental Biology 208: 28192830.CrossRefGoogle ScholarPubMed
Young, A. J., Collins, J. C. & Russell, G. (1987). Solute regulation in the euryhaline marine alga Enteromorpha prolifera (O.F. Mull) J. Ag. Journal of Experimental Botany 38: 12981308.CrossRefGoogle Scholar
Zimmermann, U. & Steudle, E. (1978). Physical aspects of water relations of plant cells. Advances in Botanical Research 6: 45117.CrossRefGoogle Scholar

References

Alam, A., Dwivedi, A. & Emmanuel, I. (2019). Resurrection plants: Imperative resources in developing strategies to drought and desiccation pressure. Plant Science Today 6: 333341.CrossRefGoogle Scholar
Andreev, V. P., Maslov, Y. I. & Sorokoletova, E. F. (2012). Functional properties of photosynthetic apparatus in three Fucus species inhabiting the White Sea: Effect of dehydration. Russian Journal of Plant Physiology 59: 217223.CrossRefGoogle Scholar
Baker, S. M. (1901). On the causes of the zoning of brown seaweeds on the seashore. New Phytologist 8: 196202.CrossRefGoogle Scholar
Bell, E. C. (1993). Photosynthetic response to temperature and desiccation of the intertidal alga Mastocarpus papillatus. Marine Biology 117: 337346.CrossRefGoogle Scholar
Bell, E. C. (1995). Environmental and morphological influences on thallus temperature and desiccation of the intertidal alga Mastocarpus papillatus Kützing. Journal of Experimental Marine Biology and Ecology 191: 2955.CrossRefGoogle Scholar
Biebl, R. (1970). Comparative studies on temperature hardiness of marine algae along the Pacific Coast of North America. Protoplasma 69: 6183.CrossRefGoogle Scholar
Blouin, N. A., Brodie, J. A., Grossman, A. C. et al. (2011). Porphyra: A marine crop shaped by stress. Trends in Plant Science 16: 2937.CrossRefGoogle ScholarPubMed
Borowitzka, M. A. (2018). The ‘stress’ concept in microalgal biology – Homeostasis, acclimation and adaptation. Journal of Applied Phycolog 30: 28152825.CrossRefGoogle Scholar
Brown, M. T. (1987). Effects of desiccation on photosynthesis of intertidal algae from a southern New Zealand shore. Botanica Marina 30: 121127.CrossRefGoogle Scholar
Burnaford, J. L., Henderson, S. Y. & Van Alstyne, K. L. (2021). Linking physiology to ecological function: Environmental conditions affect performance and size of the intertidal kelp Hedophyllum sessile (Laminariales, Phaeophyceae). Journal of Phycology 57: 128142.CrossRefGoogle ScholarPubMed
Burnaford, J. L., Nielsen, K. J. & Williams, S. L. (2014). Celestial mechanics affects emersion time and cover patterns of an ecosystem engineer, the intertidal kelp Saccharina sessilis. Marine Ecology Progress Series 509: 127136.CrossRefGoogle Scholar
Burritt, D. J., Larkindale, J. & Hurd, C. L. (2002). Antioxidant metabolism in the intertidal red seaweed Stictosiphonia arbuscula following desiccation. Planta 215: 829838.CrossRefGoogle ScholarPubMed
Buschmann, A. H. (1990). The role of herbivory and desiccation on early successional patterns of intertidal macroalgae in southern Chile. Journal of Experimental Marine Biology and Ecology 139: 221230.CrossRefGoogle Scholar
Carmignani, J. R. & Roy, A. H. (2017). Ecological impacts of winter water level drawdowns on lake littoral zones: A review. Aquatic Science 79: 803824.CrossRefGoogle Scholar
Clark, J. S., Poore, A. G. B. & Doblin, M. A. (2018). Shaping up for stress: Physiological flexibility is key to survivorship in a habitat-forming macroalga. Journal of Plant Physiology 231: 346355.CrossRefGoogle Scholar
Collén, J. & Davison, I. R. (1999a). Reactive oxygen production and damage in intertidal Fucus spp. (Phaeophyceae). Journal of Phycology 35: 5461.CrossRefGoogle Scholar
Collén, J. & Davison, I. R. (1999b). Stress tolerance and reactive oxygen metabolism in the intertidal red seaweeds Mastocarpus stellatus and Chondrus crispus. Plant, Cell and Environment 22: 11431151.CrossRefGoogle Scholar
Contreras-Porcia, L., Callejas, S., Thomas, D. et al. (2012). Seaweeds early development: Detrimental effects of desiccation and attenuation by algal extracts. Planta 235: 337348.CrossRefGoogle ScholarPubMed
Contreras-Porcia, L., López-Cristoffanini, C., Lovazzano, C. et al. (2013). Differential gene expression in Pyropia columbina (Bangiales, Rhodophyta) under natural hydration and desiccation conditions. Latin American Journal of Aquatic Research 41: 933958.CrossRefGoogle Scholar
Contreras-Porcia, L., Thomas, D., Flores, V. et al. (2011). Tolerance to oxidative stress induced by desiccation in Porphyra columbina (Bangiales, Rhodophyta). Journal of Experimental Botany 62: 18151829.CrossRefGoogle ScholarPubMed
Cronin, G. & Hay, M. E. (1996). Susceptibility to herbivores depends on recent history of both the plant and animal. Ecology 77: 15311543.CrossRefGoogle Scholar
Datta, R. & Datta, B. (1999). Desiccation induced nitrate and ammonium uptake in the red alga Catenella repens (Rhodophyta, Gigartinales). Indian Journal of Geo-Marine Sciences 28: 458460.Google Scholar
Davison, I. R. & Pearson, G. A. (1996). Stress tolerance in intertidal seaweeds. Journal of Phycology 32: 197211.CrossRefGoogle Scholar
Denny, M. W. (1993). Air and Water: The Biology and Physics of Life’s Media. Princeton University Press, Princeton, NJ, p. 360.CrossRefGoogle Scholar
Dethier, M. N., Williams, S. L. & Freeman, A. (2005). Seaweeds under stress: Manipulated stress and herbivory affect critical life-history functions. Ecological Monographs 75: 403418.CrossRefGoogle Scholar
Dring, M. & Brown, F. (1982). Photosynthesis of intertidal brown algae during and after periods of emersion: A renewed search for physiological causes of zonation. Marine Ecology Progress Series 8: 301308.CrossRefGoogle Scholar
Dring, M. J. (2005). Stress resistance and disease resistance in seaweeds: The role of reactive oxygen metabolism. Advances in Botanical Research 43: 175207.CrossRefGoogle Scholar
Dromgoole, F. I. (1980). Desiccation resistance of intertidal and subtidal algae. Botanica Marina 23: 149159.CrossRefGoogle Scholar
Ferreira, J. G., Arenas, F., Martinez, B. et al. (2014). Physiological response of fucoid algae to environmental stress: comparing range centre and southern populations. New Phytologist 202: 11571172.CrossRefGoogle ScholarPubMed
Fierro, C., López-Cristoffanini, C., Latorre, N. et al. (2016). Methylglyoxal metabolism in seaweeds during desiccation. Revista de biología marina y oceanografía 51: 187191.CrossRefGoogle Scholar
Fierro, C., Lopez-Cristoffanini, C., Meynard, A. et al. (2017). Expression profile of desiccation tolerance factors in intertidal seaweed species during the tidal cycle. Planta 245: 11491164.CrossRefGoogle ScholarPubMed
Flores-Molina, M. R., Thomas, D., Lovazzano, C. et al. (2014). Desiccation stress in intertidal seaweeds: Effects on morphology, antioxidant responses and photosynthetic performance. Aquatic Botany 113: 9099.CrossRefGoogle Scholar
Fritsch, F. E. (1945). The Structure and Reproduction of the Algae, Vol. 2. Cambridge University Press, Cambridge, UK, p. 939.Google Scholar
Gao, S., Gu, W., Xiong, Q. et al. (2015). Desiccation enhances phosphorylation of PSII and affects the distribution of protein complexes in the thylakoid membrane. Physiologia Plantarum 153: 492502.CrossRefGoogle ScholarPubMed
Gao, S., Shen, S., Wang, G. et al. (2011). PSI-driven cyclic electron flow allows intertidal macro-algae Ulva sp. (Chlorophyta) to survive in desiccated conditions. Plant and Cell Physiology 52: 885893.CrossRefGoogle ScholarPubMed
Gao, S. & Wang, G. (2012). The enhancement of cyclic electron flow around photosystem I improves the recovery of severely desiccated Porphyra yezoensis (Bangiales, Rhodophyta). Journal of Experimental Botany 63: 43494358.CrossRefGoogle ScholarPubMed
Gasulla, F., vom Dorp, K., Dombrink, I. et al. (2013). The role of lipid metabolism in the acquisition of desiccation tolerance in Craterostigma plantagineum: A comparative approach. The Plant Journal 75: 726741.CrossRefGoogle ScholarPubMed
Green, T. G. A., Pintado, A., Raggio, J. et al. (2018). The lifestyle of lichens in soil crusts. The Lichenologist 50: 397410.CrossRefGoogle Scholar
Grime, J. P. (1974). Vegetation analysis by reference to strategies. Nature 50: 2631.CrossRefGoogle Scholar
Guajardo, E., Correa, J. A. & Contreras-Porcia, L. (2016). Role of abscisic acid (ABA) in activating antioxidant tolerance responses to desiccation stress in intertidal seaweed species. Planta 243: 767781.CrossRefGoogle ScholarPubMed
Guenther, R. J. & Martone, P. T. (2014). Physiological performance of intertidal coralline algae during a simulated tidal cycle. Journal of Phycology 50: 310321.CrossRefGoogle ScholarPubMed
Hodgson, L. M. (1981). Photosynthesis of the red alga Gastroclonium coulteri (Rhodophyta) in response to changes in temperature, light intensity, and desiccation. Journal of Phycology 17: 3742.CrossRefGoogle Scholar
Holzinger, A., Herburger, K., Kaplan, F. et al. (2015). Desiccation tolerance in the chlorophyte green alga Ulva compressa: Does cell wall architecture contribute to ecological success? Planta 242: 477492.CrossRefGoogle ScholarPubMed
Huan, L., Gao, S., Xie, X. J. et al. (2014). Specific photosynthetic and morphological characteristics allow macroalgae Gloiopeltis furcata (Rhodophyta) to survive in unfavorable conditions. Photosynthetica 52: 281287.CrossRefGoogle Scholar
Hunt, L. J. & Denny, M. W. (2008). Desiccation protection and disruption: A trade-off for an intertidal marine alga. Journal of Phycology 44: 11641170.CrossRefGoogle ScholarPubMed
Hurd, C. L. & Dring, M. J. (1991). Desiccation and phosphate uptake by intertidal fucoid algae in relation to zonation. British Phycological Journal 26: 327333.CrossRefGoogle Scholar
Hurd, C. L., , P. J. Harrison, , , K. Bischof, & , C. S. Lobban, (2014). Seaweed Ecology and Physiology. Cambridge University Press, Cambridge, UK.CrossRefGoogle Scholar
Im, S., Lee, H. N., Jung, H. S. et al. (2017). Transcriptome-based identification of the desiccation response genes in marine red algae Pyropia tenera (Rhodophyta) and enhancement of abiotic stress tolerance by PtDRG2 in Chlamydomonas. Marine Biotechnology (NY) 19: 232245.CrossRefGoogle ScholarPubMed
Ito, T., Borlongan, I. A., Nishihara, G. N. et al. (2021). The effects of irradiance, temperature, and desiccation on the photosynthesis of a brown alga, Sargassum muticum (Fucales), from a native distributional range in Japan. Journal of Applied Phycology 33: 17771791.CrossRefGoogle Scholar
Johnston, A. M. & Raven, J. A. (1986). The analysis of photosynthesis in air and water of Ascophyllum nodosum (L.) Le Jol. Oecologia 69: 288295.CrossRefGoogle ScholarPubMed
Kidron, G. J. & Starinsky, A. (2019). Measurements and ecological implications of non‐rainfall water in desert ecosystems – A review. Ecohydrology 12: e2121.CrossRefGoogle Scholar
Kim, J. K., Kraemer, G. P. & Yarish, C. (2008). Physiological activity of Porphyra in relation to eulittoral zonation. Journal of Experimental Marine Biology and Ecology 365: 7585.CrossRefGoogle Scholar
Kim, J. K., Kraemer, G. P. & Yarish, C. (2012). Metabolic plasticity of nitrogen assimilation by Porphyra umbilicalis (Linnaeus) Kützing. Journal of Ocean University of China 11: 517526.CrossRefGoogle Scholar
Kim, J. K., Kraemer, G. P. & Yarish, C. (2013). Emersion induces nitrogen release and alteration of nitrogen metabolism in the intertidal genus Porphyra. PLOS ONE 8: e69961.Google ScholarPubMed
Kim, K. Y. & Garbary, D. J. (2007). Photosynthesis in Codium fragile (Chlorophyta) from a Nova Scotia estuary: Responses to desiccation and hyposalinity. Marine Biology 151: 99107.CrossRefGoogle Scholar
Kumar, M., Gupta, V., Trivedi, N. et al. (2011). Desiccation induced oxidative stress and its biochemical responses in intertidal red alga Gracilaria corticata (Gracilariales, Rhodophyta). Environmental and Experimental Botany 72: 194201.CrossRefGoogle Scholar
Kumar, M., Kumari, P., Reddy, C. R. K. et al. (2014). Salinity and desiccation induced oxidative stress acclimation in seaweeds. In: , N. Bourgougnon, (ed.) Sea Plants. Advances in Botanical Research, Vol. 71. Elsevier, Amsterdam, pp. 91124.Google Scholar
Leedham Elvidge, E. C., Phang, S.-M., Sturges, W. T. et al. (2015). The effect of desiccation on the emission of volatile bromocarbons from two common temperate macroalgae. Biogeosciences 12: 387398.CrossRefGoogle Scholar
Leprince, O. & Buitink, J. (2010). Desiccation tolerance: From genomics to the field. Plant Science 179: 554564.CrossRefGoogle Scholar
Lin, A. P., Wang, G. C., Yang, F et al. (2009). Photosynthetic parameters of sexually different parts of Porphyra katadai var. hemiphylla (Bangiales, Rhodophyta) during dehydration and re-hydration. Planta 229: 803810.CrossRefGoogle ScholarPubMed
López-Cristoffanini, C., Zapata, J., Gaillard, F. et al. (2015). Identification of proteins involved in desiccation tolerance in the red seaweed Pyropia orbicularis (Rhodophyta, Bangiales). Proteomics 15: 39543968.CrossRefGoogle ScholarPubMed
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
Madsen, T. V. & Maberly, S. C. (1990). A comparison of air and water as environments for photosynthesis by the intertidal alga Fucus spiralis (Phaeophyta). Journal of Phycology 26: 2430.CrossRefGoogle Scholar
Oates, B. R. (1985). Photosynthesis and amelioration of desiccation in the intertidal saccate alga Colpomenia peregrina. Marine Biology 89: 109119.CrossRefGoogle Scholar
Pearson, G. A., Hoarau, G., Lago-Leston, A. et al. (2010). An expressed sequence tag analysis of the intertidal brown seaweeds Fucus serratus (L.) and F. vesiculosus (L.) (Heterokontophyta, Phaeophyceae) in response to abiotic stressors. Marine Biotechnology (NY) 12: 195213.CrossRefGoogle Scholar
Pearson, G. A., Lago-Leston, A. & Mota, C. (2009). Frayed at the edges: Selective pressure and adaptive response to abiotic stressors are mismatched in low diversity edge populations. Journal of Ecology 97: 450462.CrossRefGoogle Scholar
Proctor, M. C. F. & Tuba, Z. (2002). Poikilohydry and homoihydry: Antithesis or spectrum of possibilities? New Phytologist 156: 327349.CrossRefGoogle ScholarPubMed
Qian, F., Luo, Q., Yang, R. et al. (2015). The littoral red alga Pyropia haitanensis uses rapid accumulation of floridoside as the desiccation acclimation strategy. Journal of Applied Phycology 27: 621632.CrossRefGoogle Scholar
Quadir, A., Harrison, P. J. & DeWreede, R. E. (1979). The effects of emergence and submergence on the photosynthesis and respiration of marine macrophytes. Phycologia 18: 8388.CrossRefGoogle Scholar
Raven, J. A. (2008). Transpiration: How many functions? New Phytologist 179: 905907.CrossRefGoogle ScholarPubMed
Renaud, P. E., Hay, M. E. & Schmitt, T. M. (1990). Interactions of plant stress and herbivory: Intraspecific variation in the susceptibility of a palatable versus an unpalatable seaweed to sea urchin grazing. Oecologia 82: 217226.CrossRefGoogle ScholarPubMed
Schagerl, M. & Möstl, M. (2011). Drought stress, rain and recovery of the intertidal seaweed Fucus spiralis. Marine Biology 158: 24712479.CrossRefGoogle Scholar
Schmid, M, Fernández, P. A., Gaitán-Espitia, J-D. et al. (2020) Stress due to low nitrate availability reduces the biochemical acclimation potential of the giant kelp Macrocystis pyrifera to high temperature. Algal Research 47: 101895.CrossRefGoogle Scholar
Schonbeck, M. & Norton, T. A. (1978). Factors controlling the upper limits of fucoid algae on the shore. Journal of Experimental Marine Biology and Ecology 31: 303313.CrossRefGoogle Scholar
Schonbeck, M. W. & Norton, T. A. (1979a). Drought-hardening in the upper-shore seaweeds Fucus spiralis and Pelvetia canaliculata. The Journal of Ecology 67: 687696.CrossRefGoogle Scholar
Schonbeck, M. W. & Norton, T. A. (1979b). An investigation of drought avoidance in intertidal fucoid algae. Botanica Marina 22: 133144.CrossRefGoogle Scholar
Shalaby, E. A. (2017). Influence of a biotic stress on biosynthesis of alga-chemicals and its relation to biological activities. Indian Journal of Geo-Marine Sciences 46: 2332.Google Scholar
Skene, K. R. (2004). Key differences in photosynthetic characteristics of nine species of intertidal macroalgae are related to their position on the shore. Canadian Journal of Botany 82: 177184.CrossRefGoogle Scholar
Stark, L. R. (2017). Ecology of desiccation tolerance in bryophytes: A conceptual framework and methodology. The Bryologist 120: 130165.CrossRefGoogle Scholar
Stengel, D. & Dring, M. (1997). Morphology and in situ growth rates of plants of Ascophyllum nodosum (Phaeophyta) from different shore levels and responses of plants to vertical transplantation. European Journal of Phycology 32: 193202.CrossRefGoogle Scholar
Stirk, W. A., Novák, O., Hradecká, V. et al. (2009). Endogenous cytokinins, auxins and abscisic acid in Ulva fasciata (Chlorophyta) and Dictyota humifusa (Phaeophyta): Towards understanding their biosynthesis and homoeostasis. European Journal of Phycology 44: 231240.CrossRefGoogle Scholar
Terada, R., Nishihara, G. N., Arimura, K. et al. (2021). Photosynthetic response of a cultivated red alga, Neopyropia yezoensis f. narawaensis (=Pyropia yezoensis f. narawaensis; Bangiales, Rhodophyta) to dehydration stress differs with between two heteromorphic life-history stages. Algal Research 55: 102262.CrossRefGoogle Scholar
Thomas, T. E., Turpin, D. H. & Harrison, P. J. (1987). Desiccation enhanced nitrogen uptake rates in intertidal seaweeds. Marine Biology 94: 293298.CrossRefGoogle Scholar
Wang, W.-J., Sun, X.-T., Liu, F.-L. et al. (2016). Effect of abiotic stress on the gameophyte of Pyropia katadae var. hemiphylla (Bangiales, Rhodophyta). Journal of Applied Phycology 28: 469479.CrossRefGoogle Scholar
Wright, J. T., Williams, S. L. & Dethier, M. N. (2004). No zone is always greener: Variation in the performance of Fucus gardneri embryos, juveniles and adults across tidal zone and season. Marine Biology 145: 10611073.CrossRefGoogle Scholar
Xu, D., Zhang, X., Wang, Y. et al. (2016). Responses of photosynthesis and nitrogen assimilation in the green-tide macroalga Ulva prolifera to desiccation. Marine Biology 163: 18.CrossRefGoogle Scholar
Xu, J. & Gao, K. (2015). Photosynthetic performance of the red alga Pyropia haitanensis during emersion, with special reference to effects of solar UV radiation, dehydration and elevated CO2 concentration. Photochemistry and Photobiology 91: 13761381.CrossRefGoogle ScholarPubMed
Zhang, P., Shao, Z., Jin, W. et al. (2016). Comparative characterization of two GDP-mannose dehydrogenase genes from Saccharina japonica (Laminariales, Phaeophyceae). BMC Plant Biology 16: 110.CrossRefGoogle ScholarPubMed
Zhou, W., He, L., Yang, F. et al. (2014). Pyropia yezoensis can utilize CO2 in the air during moderate dehydration. Chinese Journal of Oceanology and Limnology 32: 358364.CrossRefGoogle Scholar

References

Adl, S. M., Simpson, A. G., Lane, C. E. et al. (2012). The revised classification of eukaryotes. Journal of Eukaryotic Microbiology 59: 429514.CrossRefGoogle ScholarPubMed
Adolf, J. E., Stoecker, D. K. & Harding, Jr. L. W. (2003). Autotrophic growth and photoacclimation in Karlodinium micrum (Dinophyceae) and Storeatula major (Cryptophyceae). Journal of Phycology 39: 11011108.CrossRefGoogle Scholar
Adolf, J. E., Stoecker, D. K. & Harding, Jr. L. W. (2006). The balance of autotrophy and heterotrophy during mixotrophic growth of Karlodinium micrum (Dinophyceae). Journal of Plankton Research 28: 737751.CrossRefGoogle Scholar
Altenburger, A., Cai, H., Li, Q. et al. (2021). Limits to the cellular control of sequestered cryptophyte prey in the marine ciliate Mesodinium rubrum. The ISME Journal 15: 10561072.CrossRefGoogle Scholar
Andersen, K. H., Berge, T., Gonçalves, R. J. et al. (2016). Characteristic sizes of life in the oceans, from bacteria to whales. Annual Review of Marine Science 8: 217241.CrossRefGoogle ScholarPubMed
Anschütz, A. A. & Flynn, K. J. (2020). Niche separation between different functional types of mixoplankton: Results from NPZ-style N-based model simulations. Marine Biology 167: 121.CrossRefGoogle Scholar
Avrahami, Y. & Frada, M. J. (2020). Detection of phagotrophy in the marine phytoplankton group of the coccolithophores (Calcihaptophycidae, Haptophyta) during nutrient-replete and phosphate-limited growth. Journal of Phycology 56: 11031108.CrossRefGoogle ScholarPubMed
Behrenfeld, M. J., Halsey, K. H., Boss, E. et al. (2021). Thoughts on the evolution and ecological niche of diatoms. Ecological Monographs 91: e01457.CrossRefGoogle Scholar
Biard, T., Stemmann, L., Picheral, M. et al. (2016). In situ imaging reveals the biomass of giant protists in the global ocean. Nature 532: 504507.CrossRefGoogle ScholarPubMed
Biddanda, B. & Benner, R. (1997). Carbon, nitrogen, and carbohydrate fluxes during the production of particulate and dissolved organic matter by marine phytoplankton. Limnology and Oceanography 42: 506518.CrossRefGoogle Scholar
Blossom, H. E. & Hansen, P. J. (2021). The loss of mixotrophy in Alexandrium pseudogonyaulax: Implications for trade-offs between toxicity, mucus trap production, and phagotrophy. Limnology and Oceanography 66: 528542.CrossRefGoogle Scholar
Burkholder, J. M., Glibert, P. M. & Skelton, H. M. (2008). Mixotrophy, a major mode of nutrition for harmful algal species in eutrophic waters. Harmful Algae 8: 7793.CrossRefGoogle Scholar
Caron, D. A., Porter, K. G. & Sanders, R. W. (1990). Carbon, nitrogen, and phosphorus budgets for the mixotrophic phytoflagellate Poterioochromonas malhamensis (Chrysophyceae) during bacterial ingestion. Limnology and Oceanography 35: 433443.CrossRefGoogle Scholar
Charalampous, E., Matthiessen, B. & Sommer, U. (2018). Light effects on phytoplankton morphometric traits influence nutrient utilization ability. Journal of Plankton Research 40: 568579.CrossRefGoogle Scholar
Clark, D. R., Flynn, K. J. & Owens, N. J. P. (2002). The large capacity for dark nitrate-assimilation in diatoms may overcome nitrate limitation of growth. New Phytologist 155: 101108.CrossRefGoogle ScholarPubMed
Crawford, D. W. (1989). Mesodinium rubrum: The phytoplankter that wasn’t. Marine Ecology Progress Series 79: 259265.CrossRefGoogle Scholar
Crawford, D. W. & Purdie, D. A. (1992). Evidence for avoidance of flushing from an estuary by a planktonic, phototrophic ciliate. Marine Ecology Progress Series 79: 259265.CrossRefGoogle Scholar
Davison, I. R. (1991). Environmental effects on algal photosynthesis: Temperature. Journal of Phycology 27: 28.CrossRefGoogle Scholar
de Castro, F., Gaedke, U. & Boenigk, J. (2009). Reverse evolution: Driving forces behind the loss of acquired photosynthetic traits. PLOS ONE 4: e8465.CrossRefGoogle ScholarPubMed
de Figueiredo, G. M., Nash, R. D. & Montagnes, D. J. (2007). Do protozoa contribute significantly to the diet of larval fish in the Irish Sea? Journal of the Marine Biological Association of the United Kingdom 87: 843850.CrossRefGoogle Scholar
de Vries, J. & Gould, S. B. (2018). The monoplastidic bottleneck in algae and plant evolution. Journal of Cell Science 131: jcs203414.CrossRefGoogle ScholarPubMed
Dolan, J. R. & Pérez, M. T. (2000). Costs, benefits and characteristics of mixotrophy in marine oligotrichs. Freshwater Biology 45: 227238.CrossRefGoogle Scholar
Edwards, K. F., Thomas, M. K., Klausmeier, C. A. et al. (2012). Allometric scaling and taxonomic variation in nutrient utilization traits and maximum growth rate of phytoplankton. Limnology and Oceanography 57: 554566.CrossRefGoogle Scholar
Ehrlich, E., Kath, N. J. & Gaedke, U. (2020). The shape of a defense-growth trade-off governs seasonal trait dynamics in natural phytoplankton. The ISME Journal 14: 14511462.CrossRefGoogle ScholarPubMed
Esteban, G. F., Fenchel, T. & Finlay, B. J. (2010). Mixotrophy in ciliates. Protist 161: 621641.CrossRefGoogle ScholarPubMed
Fenchel, T. & Hansen, P. J. (2006). Motile behaviour of the bloom-forming ciliate Mesodinium rubrum. Marine Biology Research 2: 3340.CrossRefGoogle Scholar
Flynn, K. J. (2009). Going for the slow burn: Why should possession of a low maximum growth rate be advantageous for microalgae? Plant Ecology and Diversity 2: 179189.CrossRefGoogle Scholar
Flynn, K. J. & Berry, L. S. (1999). The loss of organic nitrogen during marine primary production may be overestimated significantly when estimated using 15N substrates. Proceedings of the Royal Society of London. Series B: Biological Sciences 266: 641647.Google Scholar
Flynn, K. J. & Hansen, P. J. (2013). Cutting the canopy to defeat the ‘selfish gene’; conflicting selection pressures for the integration of phototrophy in mixotrophic protists. Protist 164: 811823.CrossRefGoogle ScholarPubMed
Flynn, K. J. & Skibinski, D. O. (2020). Exploring evolution of maximum growth rates in plankton. Journal of Plankton Research 42: 497513.CrossRefGoogle ScholarPubMed
Flynn, K. J. & Hipkin, C. R. (1999). Interactions between iron, light, ammonium, and nitrate: Insights from the construction of a dynamic model of algal physiology. Journal of Phycology 35: 11711190.CrossRefGoogle Scholar
Flynn, K. J. & Mitra, A. (2016). Why plankton modelers should reconsider using rectangular hyperbolic (Michaelis-Menten, Monod) descriptions of predator-prey interactions. Frontiers in Marine Science 3: 165.CrossRefGoogle Scholar
Flynn, K. J. & Skibinski, D. O. F., (2020). Exploring evolution of maximum growth rates in plankton. Journal of Plankton Research 42: 497513.CrossRefGoogle ScholarPubMed
Flynn, K. J., Blackford, J. C., Baird, M. E. et al. (2012). Changes in pH at the exterior surface of plankton with ocean acidification. Nature Climate Change 2: 510513.CrossRefGoogle Scholar
Flynn, K. J., Stoecker, D. K., Mitra, A. et al. (2013). Misuse of the phytoplankton–zooplankton dichotomy: The need to assign organisms as mixotrophs within plankton functional types. Journal of Plankton Research 35: 311.CrossRefGoogle Scholar
Flynn, K. J., St. John, M., Raven, J. A. et al. (2015a). Acclimation, adaptation, traits and trade-offs in plankton functional type models: Reconciling terminology for biology and modelling. Journal Plankton Research 37: 683691.CrossRefGoogle Scholar
Flynn, K. J., Clark, D. R., Mitra, A. et al. (2015b). Ocean acidification with (de)eutrophication will alter future phytoplankton growth and succession. Proceedings of the Royal Society of London. Series B: Biological Sciences 282: 20142604.CrossRefGoogle Scholar
Flynn, K. J., Skibinski, D. O. F. & Lindemann, C. (2018). Effects of growth rate, cell size, motion, and elemental stoichiometry on nutrient transport kinetics. PLOS Computational Biology 14: e1006118.CrossRefGoogle ScholarPubMed
Flynn, K. J., Mitra, A., Anestis, K. et al. (2019). Mixotrophic protists and a new paradigm for marine ecology: Where does plankton research go now? Journal of Plankton Research 41: 375391.CrossRefGoogle Scholar
Flynn, K. J., Kimmance, S. A., Clark, D. R. et al. (2021) Modelling the effects of traits and abiotic factors on viral lysis in phytoplankton. Frontiers in Marine Science 8: 667184.CrossRefGoogle Scholar
Fogg, G. E. (1991). The phytoplanktonic ways of life. New Phytologist 118: 191232.CrossRefGoogle ScholarPubMed
Gavelis, G. S., Wakeman, K. C., Tillmann, U. et al. (2017). Microbial arms race: Ballistic ‘nematocysts’ in dinoflagellates represent a new extreme in organelle complexity. Science Advances 3: e1602552.CrossRefGoogle ScholarPubMed
Glibert, P. M. (2016). Margalef revisited: A new phytoplankton mandala incorporating twelve dimensions, including nutritional physiology. Harmful Algae 55: 2530.CrossRefGoogle ScholarPubMed
Glibert, P. M. & Mitra, A. (2022). From webs, loops, shunts, and pumps to microbial multitasking: Evolving concepts of marine microbial ecology, the mixoplankton paradigm, and implications for a future ocean. Limnology and Oceanography 67: 585597.CrossRefGoogle Scholar
Gómez-Consarnau, L., Raven, J. A., Levine, N. M. et al. (2019). Microbial rhodopsins are major contributors to the solar energy captured in the sea. Science Advances 5: eaaw8855.CrossRefGoogle Scholar
Granéli, E. & Flynn, K. J. (2006). Chemical and physical factors influencing toxin production. In Granéli, E. & Turner, J. T. (eds.) Ecology of Harmful Algae, Springer-Verlag, Berlin, 189, pp. 229241.CrossRefGoogle Scholar
Hansen, P. J. (2002). Effect of high pH on the growth and survival of marine phytoplankton: Implications for species succession. Aquatic Microbial Ecology 28: 279288.CrossRefGoogle Scholar
Hansen, P. J., Skovgaard, A., Glud, R. N. et al. (2000). Physiology of the mixotrophic dinoflagellate Fragilidium subglobosum. II. Effects of time scale and prey concentration on photosynthetic performance. Marine Ecology Progress Series 201: 137146.CrossRefGoogle Scholar
Hartmann, M., Grob, C., Tarran, G. A. et al. (2012). Mixotrophic basis of Atlantic oligotrophic ecosystems. Proceedings of the National Academy of Sciences USA 109: 57565760.Google Scholar
Hausmann, K. (2002). Food acquisition, food ingestion and food digestion by protists. Japanese Journal of Protozoology 35: 8595.Google Scholar
Hughes, E. A., Maselli, M., Sørensen, H. et al. (2021). Metabolic reliance on photosynthesis depends on both irradiance and prey availability in the mixotrophic ciliate, Strombidium cf. basimorphum. Frontiers in Microbiology 12: 642600.CrossRefGoogle ScholarPubMed
Jeong, H. J., Du Yoo, Y., Kim, J. S. et al. (2010). Growth, feeding and ecological roles of the mixotrophic and heterotrophic dinoflagellates in marine planktonic food webs. Ocean Science Journal 45: 6591.CrossRefGoogle Scholar
Jiang, H. & Johnson, M. D. (2017). Jumping and overcoming diffusion limitation of nutrient uptake in the photosynthetic ciliate Mesodinium rubrum. Limnology and Oceanography 62: 421436.CrossRefGoogle Scholar
John, E. H. & Flynn, K. J. (2002). Modelling changes in paralytic shellfish toxin content of dinoflagellates in response to nitrogen and phosphorus supply. Marine Ecology Progress Series 225: 147160.CrossRefGoogle Scholar
Johnson, M. D. (2011). The acquisition of phototropy: Adaptive strategies of hosting endosymbionts and organelles. Photosynthesis Research 107: 117132.CrossRefGoogle Scholar
Johnson, M. D. & Stoecker, D. K. (2005). The role of feeding in growth and the photophysiology of Myrionecta rubra. Aquatic Microbial Ecology 39: 303312.CrossRefGoogle Scholar
Johnson, P. W., Donaghay, P. L., Small, E. B. et al. (1995). Ultrastructure and ecology of Perispira ovum (Ciliophora, Litostomatea) an aerobic, planktonic ciliate that sequesters the chloroplasts, mitochrondria, and paramylon of Euglenia proxima in a micro-oxic habitat. Journal of Eukaryotic Microbiology 42: 323335.CrossRefGoogle Scholar
Johnson, M. D., Stoecker, D. K., Tengs, T. et al. (2006). Sequestration and performance of cryptophyte plastids in Myrionecta rubra. Journal of Phycology 42: 12361246.CrossRefGoogle Scholar
Johnson, M. D., Beaudoin, D. J., Frada, M. J. et al. (2018). High grazing rates on cryptophyte algae in Chesapeake Bay. Frontiers in Marine Science 5: 241.CrossRefGoogle Scholar
Keeling, P. J., Burki, F., Wilcox, H. M. et al. (2014). The Marine Microbial Eukaryote Transcriptome Sequencing Project (MMETSP): Illuminating the functional diversity of eukaryotic life in the oceans through transcriptome sequencing. PLOS Biology 12: e1001889.CrossRefGoogle ScholarPubMed
Kemp, A. E. & Villareal, T. A. (2018). The case of the diatoms and the muddled mandalas: Time to recognize diatom adaptations to stratified waters. Progress in Oceanography 167: 138149.CrossRefGoogle Scholar
Kim, G. H., Han, J. H., Kim, B. et al. (2016). Cryptophyte gene regulation in the kleptoplastidic, karyokleptic ciliate Mesodinium rubrum. Harmful Algae 52: 2333.CrossRefGoogle ScholarPubMed
Lee, K. H., Jeong, H. J., Jang, T. Y. et al. (2014). Feeding by the newly described mixotrophic dinoflagellate Gymnodinium smaydae: Feeding mechanism, prey species, and effect of prey concentration. Journal of Experimental Marine Biology and Ecology 459: 114125.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 of London. Series B: Biological Sciences 284: 20170664.Google Scholar
Leles, S. C., Mitra, A., Flynn, K. J. et al. (2019). Sampling bias misrepresents the biogeographic significance of constitutive mixotrophs across global oceans. Global Ecology and Biogeography 28: 418428.CrossRefGoogle Scholar
Leles, S. G., Bruggeman, J., Polimene, L. et al. (2021). Differences in physiology explain succession of mixoplankton functional types and affect carbon fluxes in temperate seas. Progress in Oceanography 190: 102481.CrossRefGoogle Scholar
Li, A., Stoecker, D. K. & Adolf, J. E. (1999). Feeding, pigmentation, photosynthesis and growth of the mixotrophic dinoflagellate Gyrodinium galatheanum. Aquatic Microbial Ecology 19: 163176.CrossRefGoogle Scholar
Litchman, E. & Klausmeier, C. A. (2008). Trait-based community ecology of phytoplankton. Annual Review of Ecology, Evolution and Systematics 39: 615639.CrossRefGoogle Scholar
Litchman, E., Ohman, M. D. & Kiørboe, T. (2013). Trait-based approaches to zooplankton communities. Journal of Plankton Research 35: 473484.CrossRefGoogle Scholar
Maberly, S. C., Ball, L. A., Raven, J. A. et al. (2009). Inorganic carbon acquisition by chrysophytes. Journal of Phycology 45: 10521061.CrossRefGoogle ScholarPubMed
Malviya, S., Scalco, E., Audic, S. et al. (2016). Insights into global diatom distribution and diversity in the world’s ocean. Proceedings of the National Academy of Sciences USA 113: E1516E1525.Google Scholar
Margalef, R. (1978). Life-forms of phytoplankton as survival alternatives in an unstable environment. Oceanologia Acta 1: 493509.Google Scholar
McCue, M. D. (2006). Specific dynamic action: A century of investigation. Comparative Biochemistry and Physiology Part A: Molecular and Integrative Physiology 144: 381394.CrossRefGoogle Scholar
McKie-Krisberg, Z. M., Gast, R. J. & Sanders, R. W. (2015). Physiological responses of three species of Antarctic mixotrophic phytoflagellates to changes in light and dissolved nutrients. Microbial Ecology 70: 2129.CrossRefGoogle ScholarPubMed
McManus, G. B., Schoener, D. M. & Haberlandt, K. (2017). Chloroplast symbiosis in a marine ciliate: Ecophysiology and the risks and rewards of hosting foreign organelles. Frontiers in Microbiology 3: 321.Google Scholar
Millette, N. C., Pierson, J. J., Aceves, A. et al. (2017). Mixotrophy in Heterocapsa rotundata: A mechanism for dominating the winter phytoplankton. Limnology and Oceanography 62: 836845.CrossRefGoogle Scholar
Mitra, A. & Flynn, K. J. (2006). Promotion of harmful algal blooms by zooplankton predatory activity. Biology Letters 2: 194197.CrossRefGoogle ScholarPubMed
Mitra, A. & Flynn, K. J. (2010). Modelling mixotrophy in harmful algal blooms: More or less the sum of the parts? Journal of Marine Systems 83: 158169.CrossRefGoogle Scholar
Mitra, A. & Flynn, K. J. (2021). HABs and the mixoplankton paradigm. Harmful Algae News 67: 46.Google Scholar
Mitra, A. & Flynn, K. J. (2023). Low rates of bacterivory enhances phototrophy and competitive advantage for mixoplankton growing in oligotrophic waters. Scientific Reports 13: 6900.Google Scholar
Mitra, A., Flynn, K. J., Burkholder, J. M. et al. (2014). The role of mixotrophic protists in the biological carbon pump. Biogeosciences 11: 9951005.CrossRefGoogle Scholar
Mitra, A., Flynn, K. J., Tillmann, U. et al. (2016). Defining planktonic protist functional groups on mechanisms for energy and nutrient acquisition; incorporation of diverse mixotrophic strategies. Protist 167: 106120.CrossRefGoogle ScholarPubMed
Mitra, A., Caron, D. A., Faure, E. et al. (2023a). The Mixoplankton Database – diversity of photo-phago-trophic plankton in form, function and distribution across the global ocean. Journal of Eukaryotic Microbiology 70: e12972. https://doi.org/10.1111/jeu.12972.CrossRefGoogle ScholarPubMed
Mitra, A., Caron, D. A., Faure, E. et al. (2023b). The Mixoplankton Database (MDB). Zenodo. https://doi.org/10.5281/zenodo.7560583.CrossRefGoogle Scholar
Moestrup, Ø., Akselmann-Cardella, R., Churro, C. et al. (eds.) (2021). IOC-UNESCO Taxonomic Reference List of Harmful Micro Algae. Accessed at www.marinespecies.org/hab on 2021–02–08. https://doi.org/10.14284/362.CrossRefGoogle Scholar
Nishitani, G. O. H., Nagai, S., Sakiyama, S. et al. (2008). Successful cultivation of the toxic dinoflagellate Dinophysis caudata (Dinophyceae). Plankton and Benthos Research 3: 7885.CrossRefGoogle Scholar
Öpik, H. & Flynn, K. J. (1989). The digestive process of the dinoflagellate, Oxyrrhis marina Dujardin, feeding on the chlorophyte, Dunaliella primolecta Butcher: A combined study of ultrastructure and free amino acids. New Phytologist 113: 143151.CrossRefGoogle Scholar
Pančić, M. & Kiørboe, T. (2018). Phytoplankton defence mechanisms: Traits and trade-offs. Biological Reviews 93: 12691303. https://doi.org/10.1111/brv.12395.CrossRefGoogle ScholarPubMed
Park, M. G., Kim, S., Kim, H. S. et al. (2006). First successful culture of the marine dinoflagellate Dinophysis acuminata. Aquatic Microbial Ecology 45: 101106.CrossRefGoogle Scholar
Park, J. S., Myung, G., Kim, H. S. et al. (2007). Growth responses of the marine photosynthetic ciliate Myrionecta rubra to different cryptomonad strains. Aquatic Microbial Ecology 48: 8390.CrossRefGoogle Scholar
Pérez, M. T., Dolan, J. R. & Fukai, E. (1997). Planktonic oligotrich ciliates in the NW Mediterranean: Growth rates and consumption by copepods. Marine Ecology Progress Series 155: 89101.CrossRefGoogle Scholar
Ponce-Toledo, R. I., Deschamps, P., López-García, P. et al. (2017). An early-branching freshwater cyanobacterium at the origin of plastids. Current Biology 27: 368391.CrossRefGoogle ScholarPubMed
Princiotta, S. D., Smith, B. T. & Sanders, R. W. (2016). Temperature-dependent phagotrophy and phototrophy in a mixotrophic chrysophyte. Journal of Phycology 52: 432440.CrossRefGoogle Scholar
Raven, J. A., Beardall, J., Flynn, K. J. et al. (2009). Phagotrophy in the origins of photosynthesis in eukaryotes and as a complementary mode of nutrition in phototrophs: Relation to Darwin’s insectivorous plants. Journal of Experimental Botany 60: 39753987.CrossRefGoogle ScholarPubMed
Raven, J. A., Suggett, D. J. & Giordano, M. (2020). Inorganic carbon concentrating mechanisms in free-living and symbiotic dinoflagellates and chromerids. Journal of Applied Phycology 56: 13771397.CrossRefGoogle ScholarPubMed
Rottberger, J., Gruber, A., Boenigk, J. et al. (2013). Influence of nutrients and light on autotrophic, mixotrophic and heterotrophic freshwater chrysophytes. Aquatic Microbial Ecology 71: 179191.CrossRefGoogle Scholar
Sánchez-Baracaldo, P., Raven, J. A., Pisari, D. et al. (2017). Early photosynthetic eukaryotes inhabited low-salinity habitats. Proceedings of the National Academy of Sciences USA 114: E7737E7745.Google Scholar
Sato, N. (2020). Complex origins of chloroplast membranes with photosynthetic machineries: Multiple transfers of genes from divergent organisms at different times or a single endosymbiotic event? Journal of Plant Research 133: 1533.CrossRefGoogle ScholarPubMed
Schoener, D. M. & McManus, G. B. (2012). Plastid retention, use and replacement in a kleptoplastidic ciliate. Aquatic Microbial Ecology 67: 177187.CrossRefGoogle Scholar
Silkin, V., Fedorov, A., Flynn, K. J. et al. (2021). Protoplasmic streaming of chloroplasts enables rapid photoacclimation in large diatoms. Journal of Plankton Research 43: 831845.CrossRefGoogle Scholar
Sonntag, B., Summerer, M. & Sommaruga, R. (2011). Are freshwater mixotrophic ciliates less sensitive to solar ultraviolet radiation than heterotrophic ones? Journal of Eukaryotic Microbiology 58: 196202.CrossRefGoogle ScholarPubMed
Stoecker, D. K. (1999). Mixotrophy among dinoflagellates. Journal of Eukaryotic Microbiology 46: 397401.CrossRefGoogle Scholar
Stoecker, D. K., Silver, M. W. (1990). Replacement and aging of chloroplasts in Strombidium capitatum (Ciliata, Oligotrichida). Marine Biology 107: 491502.CrossRefGoogle Scholar
Stoecker, D. K., Silver, M. W., Michaels, A. E. et al. (1988). Obligate mixotrophy in Laboea strobila, a ciliate which retains chloroplasts. Marine Biology 99: 415423.CrossRefGoogle Scholar
Stoecker, D. K., Johnson, M. D., de Vargas, C. et al. (2009). Acquired phototrophy in aquatic protists. Aquatic Microbial Ecology 57: 279310.CrossRefGoogle Scholar
Stoecker, D. K., Hansen, P. J., Caron, D. A. et al. (2017). Mixotrophy in the marine plankton. Annual Review of Marine Science 9: 311335.CrossRefGoogle ScholarPubMed
Tillmann, U. (2003). Kill and eat your predator: A winning strategy of the planktonic flagellate Prymnesium parvum. Aquatic Microbial Ecology 32: 7384.CrossRefGoogle Scholar
van der Meeren, T. & Naess, T. (1993). How does cod (Gadus morhua) cope with variability in feeding conditions during early larval stages? Marine Biology 116: 637647.CrossRefGoogle Scholar
Ward, B. A. & Follows, M. J. (2016). Marine mixotrophy increases trophic transfer efficiency, mean organism size, and vertical carbon flux. Proceedings of the National Academy of Sciences USA 113: 29582963.Google Scholar
Wetz, M. S. & Wheeler, P. A. (2007). Release of dissolved organic matter by coastal diatoms. Limnology and Oceanography 52:798807.CrossRefGoogle Scholar
Yamasaki, Y., Nagasoe, S., Matsubara, T. et al. (2007). Allelopathic interactions between the bacillariophyte Skeletonema costatum and the raphidophyte Heterosigma akashiwo. Marine Ecology Progress Series 339: 8392.CrossRefGoogle Scholar
Zubkov, M. V. & Tarran, G. A. (2008). High bacterivory by the smallest phytoplankton in the North Atlantic Ocean. Nature 455: 224226.CrossRefGoogle ScholarPubMed

References

Abdel-Shafy, H. I. & Mansour, M. S. M. (2016). A review on polycyclic aromatic hydrocarbons: Source, environmental impact, effect on human health and remediation. Egyptian Journal of Petroleum 25: 107123.CrossRefGoogle Scholar
Abinandan, S., Subashchandrabose, S. R., Venkateswarlu, K. et al. (2020). Sustainable iron recovery and biodiesel yield by acid-adapted microalgae, Desmodesmus sp. MAS1 and Heterochlorella sp. MAS3, grown in synthetic acid mine drainage. ACS Omega 5: 68886894.CrossRefGoogle ScholarPubMed
Almeida, A. C., Gomes, T., Langford, K. et al. (2019). Oxidative stress potential of the herbicides bifenox and metribuzin in the microalgae Chlamydomonas reinhardtii. Aquatic Toxicology 210: 117128.CrossRefGoogle ScholarPubMed
Andrady, A. L. (2011). Microplastics in the marine environment. Marine Pollution Bulletin 62: 15961605.CrossRefGoogle ScholarPubMed
Anyasi, R. O. (2011). Biological remediation of polychlorinated biphenyls (PCB) in the environment by microorganisms and plants. African Journal of Biotechnology 10: 1891618938.CrossRefGoogle Scholar
Asghari, S., Rajabi, F., Tarrahi, R. et al. (2020). Potential of the green microalga Chlorella vulgaris to fight against fluorene contamination: Evaluation of antioxidant systems and identification of intermediate biodegradation compounds. Journal of Applied Phycology 32: 411419.CrossRefGoogle Scholar
Baghour, M. (2019). Algal degradation of organic pollutants. In Martínez, L., Kharissova, O. & Kharisov, B (eds.) Handbook of Ecomaterials, Vol. 1. Springer International Publishing, Cham, pp. 565586.CrossRefGoogle Scholar
Baker, L. F., Mudge, J. F., Thompson, D. G. et al. (2016). The combined influence of two agricultural contaminants on natural communities of phytoplankton and zooplankton. Ecotoxicology 25: 10211032.CrossRefGoogle ScholarPubMed
Batterton, J. C., Boush, G. M. & Matsumura, F. (1971). Growth response of blue-green algae to aldrin, dieldrin endrin and their metabolites. Bulletin of Environmental Contamination and Toxicology 6: 589591.CrossRefGoogle ScholarPubMed
Ben Chekroun, K., Sánchez, E. & Baghour, M. (2014). The role of algae in bioremediation of organic pollutants. International Research Journal of Public and Environmental Health 1: 1932.Google Scholar
Bergami, E., Pugnalini, S., Vannuccini, M. L. et al. (2017). Long-term toxicity of surface-charged polystyrene nanoplastics to marine planktonic species Dunaliella tertiolecta and Artemia franciscana. Aquatic Toxicology 189: 159169.CrossRefGoogle ScholarPubMed
Besseling, E., Wang, B., Lürling, M. et al. (2014). Nanoplastic affects growth of S. obliquus and reproduction of D. magna. Environmental Science and Technology 48: 1233612343.CrossRefGoogle Scholar
Beyer, A. & Biziuk, M. (2009). Environmental fate and global distribution of polychlorinated biphenyls. In Whitacre, D. M. (ed.) Reviews of Environmental Contamination and Toxicology, Vol. 201. Springer, Boston, MA, pp. 137158.Google Scholar
Bhattacharya, P., Lin, S., Turner, J. P. et al. (2010). Physical adsorption of charged plastic nanoparticles affects algal photosynthesis. Journal of Physical Chemistry C 114: 1655616561.CrossRefGoogle Scholar
Biggs, D. C., Rowland, R. G. & Wurster, C. F. (1979). Effects of trichloroethylene, hexachlorobenzene and polychlorinated biphenyls on the growth and cell size of marine phytoplankton. Bulletin of Environmental Contamination and Toxicology 21: 196201.CrossRefGoogle ScholarPubMed
Birungi, Z. S. & Chirwa, E. M. N. (2015). The adsorption potential and recovery of thallium using green micro-algae from eutrophic water sources. Journal of Hazardous Materials 299: 6777.CrossRefGoogle ScholarPubMed
Bodin, H., Asp, H. & Hultberg, M. (2017). Effects of biopellets composed of microalgae and fungi on cadmium present at environmentally relevant levels in water. International Journal of Phytoremediation 19: 500504.CrossRefGoogle ScholarPubMed
Bwapwa, J. K., Jaiyeola, A. T. & Chetty, R. (2017). Bioremediation of acid mine drainage using algae strains: A review. South African Journal of Chemical Engineering 24: 6270.CrossRefGoogle Scholar
Canniff, P. M. & Hoang, T. C. (2018). Microplastic ingestion by Daphnia magna and its enhancement on algal growth. Science of the Total Environment 633: 500507.CrossRefGoogle ScholarPubMed
Carder, J. P. & Hoagland, K. D. (1998). Combined effects of alachlor and atrazine on benthic algal communities in artificial streams. Environmental Toxicology and Chemistry 17: 14151420.CrossRefGoogle Scholar
Carrera-Martinez, D., Mateos-Sanz, A., Lopez-Rodas, V. et al. (2011). Adaptation of microalgae to a gradient of continuous petroleum contamination. Aquatic Toxicology 101: 342350.CrossRefGoogle ScholarPubMed
Chae, Y. & An, Y. J. (2017). Effects of micro- and nanoplastics on aquatic ecosystems: Current research trends and perspectives. Marine Pollution Bulletin 124: 624632.CrossRefGoogle Scholar
Chae, Y., Kim, D. & An, Y. J. (2019). Effects of micro-sized polyethylene spheres on the marine microalga Dunaliella salina: Focusing on the algal cell to plastic particle size ratio. Aquatic Toxicology 216: 105296.CrossRefGoogle ScholarPubMed
Chamsi, O., Pinelli, E., Faucon, B. et al. (2019). Effects of herbicide mixtures on freshwater microalgae with the potential effect of a safener. Annales de Limnologie – International Journal of Limnology 55: 19.CrossRefGoogle Scholar
Chaudhari, P. R., Jayangouder, I. & Krishnamoorthi, K. P. (1989). Response of some common freshwater algae to DDT applications. Proceedings of the lndian Academy of. Science (Plant Science) 99: 279285.Google Scholar
Chen, W., Jia, Y., Liu, A. et al. (2017). Simultaneous elimination of cyanotoxins and PCBs via mechanical collection of cyanobacterial blooms: An application of ‘green-bioadsorption concept’. Journal of Environmental Sciences (China) 57: 118126.CrossRefGoogle ScholarPubMed
Chetouhi, C., Masseret, E., Satta, C. T. et al. (2020). Intraspecific variability in membrane proteome, cell growth, and morphometry of the invasive marine neurotoxic dinoflagellate Alexandrium pacificum grown in metal-contaminated conditions. Science of the Total Environment 715: 136834.CrossRefGoogle ScholarPubMed
Chiellini, C., Guglielminetti, L., Pistelli, L. et al. (2020). Screening of trace metal elements for pollution tolerance of freshwater and marine microalgal strains: Overview and perspectives. Algal Research 45: 101751.CrossRefGoogle Scholar
Choix, F. J., de-Bashan, L. E. & Bashan, Y. (2012). Enhanced accumulation of starch and total carbohydrates in alginate-immobilized Chlorella spp. induced by Azospirillum brasilense: II. Heterotrophic conditions. Enzyme and Microbial Technology 51: 300309.CrossRefGoogle ScholarPubMed
Cole, M., Lindeque, P., Fileman, E. et al. (2013). Microplastic ingestion by zooplankton. Environmental Science and Technology 47: 66466655.CrossRefGoogle ScholarPubMed
DaSilva, E. J., Henriksson, L. E. & Henriksson, E. (1975). Effect of pesticides on blue-green algae and nitrogen-fixation. Archives of Environmental Contamination and Toxicology 3: 193204.CrossRefGoogle ScholarPubMed
Davis, J. A., Hetzel, F., Oram, J. J. et al. (2007). Polychlorinated biphenyls (PCBs) in San Francisco Bay. Environmental Research 105: 6786.CrossRefGoogle ScholarPubMed
De-Bashan, L. E., Bashan, Y., Moreno, M. et al. (2002). Increased pigment and lipid content, lipid variety, and cell and population size of the microalgae Chlorella spp. when co-immobilized in alginate beads with the microalgae-growth-promoting bacterium Azospirillum brasilense. Canadian Journal of Microbiology 48: 514521.CrossRefGoogle ScholarPubMed
Debenest, T., Silvestre, J., Coste, M., et al. (2008). Herbicide effects on freshwater benthic diatoms: Induction of nucleus alterations and silica cell wall abnormalities. Aquatic Toxicology 88: 8894.CrossRefGoogle ScholarPubMed
Debenest, T., Silvestre, J., Coste, M. et al. (2010). Effects of pesticides on freshwater diatoms. Reviews of Environmental Contamination and Toxicology 203: 87103.Google ScholarPubMed
Deng, J., Fu, D., Hu, W. et al. (2020). Physiological responses and accumulation ability of Microcystis aeruginosa to zinc and cadmium: Implications for bioremediation of heavy metal pollution. Bioresource Technology 303: 122963.CrossRefGoogle ScholarPubMed
Deng, L., Senseman, S. A., Gentry, T. J. et al. (2015). Effect of selected herbicides on growth and lipid content of Nannochloris oculata. Journal of Aquatic Plant Management 53: 2835.Google Scholar
Ebenezer, V. & Ki, J.-S. (2016). Toxic effects of Aroclor 1016 and bisphenol A on marine green algae Tetraselmis suecica, diatom Ditylum brightwellii and dinoflagellate Prorocentrum minimum. The Korean Journal of Microbiology 52: 306312.CrossRefGoogle Scholar
Eich, A., Mildenberger, T., Laforsch, C. et al. (2015). Biofilm and diatom succession on polyethylene (PE) and biodegradable plastic bags in two marine habitats: Early signs of degradation in the pelagic and benthic zone? PLOS ONE 10: e0137201.CrossRefGoogle ScholarPubMed
Elahi, A., Arooj, I., Bukhari, D. A. et al. (2020). Successive use of microorganisms to remove chromium from wastewater. Applied Microbiology and Biotechnology 104: 37293743.CrossRefGoogle ScholarPubMed
Eregie, S. B. & Jamal-Ally, S. F. (2019). Comparison of biodegradation of lubricant wastes by Scenedesmus vacuolatus vs a microalgal consortium. Bioremediation Journal 23: 277301.CrossRefGoogle Scholar
Erickson, M. D. (1997). Analytical Chemistry of PCBs, 2nd ed. CRC Press, New York, NY.Google Scholar
Fitzgerald, S. A. & Steuer, J. J. (2006). Association of polychlorinated biphenyls (PCBs) with live algae and total lipids in rivers – A field-based approach. Science of the Total Environment 354: 6074.CrossRefGoogle Scholar
Gonzalez-Rey, M., Tapie, N., Le Menach, K. et al. (2015). Occurrence of pharmaceutical compounds and pesticides in aquatic systems. Marine Pollution Bulletin 96: 384400.CrossRefGoogle ScholarPubMed
González, J., Figueiras, F. G., Aranguren-Gassis, M. et al. (2009). Effect of a simulated oil spill on natural assemblages of marine phytoplankton enclosed in microcosms. Estuarine, Coastal and Shelf Science 83: 265276.CrossRefGoogle Scholar
Groner, M. L. & Relyea, R. A. (2011). A tale of two pesticides: How common insecticides affect aquatic communities. Freshwater Biology 56: 23912404.CrossRefGoogle Scholar
Guo, Y., Ma, W., Li, J. et al. (2020). Effects of microplastics on growth, phenanthrene stress, and lipid accumulation in a diatom, Phaeodactylum tricornutum. Environmental Pollution 257: 113628.CrossRefGoogle Scholar
Halm-Lemeille, M. P., Abbaszadeh Fard, E., Latire, T. et al. (2014). The effect of different polychlorinated biphenyls on two aquatic models, the green alga Pseudokirchneriella subcapitata and the haemocytes from the European abalone Haliotis tuberculata. Chemosphere 110: 120128.CrossRefGoogle ScholarPubMed
Harding, L. W. & Phillips, J. H. (1978). Polychlorinated biphenyl (PCB) effects on marine phytoplankton photosynthesis and cell division. Marine Biology 49: 93101.CrossRefGoogle Scholar
Hernando, M. D., Ferrer, I., Agüera, A. et al. (2004). Evaluation of pesticides in wastewaters. A combined (chemical and biological) analytical approach. In Barcelo, D. (ed.) Water Pollution. The Handbook of Environmental Chemistry, Vol. 2. Springer-Verlag, Berlin, Heidelberg, pp. 5377.Google Scholar
Ibrahim, W. M., Karam, M. A., El-Shahat, R. M. et al. (2014). Biodegradation and utilization of organophosphorus pesticide malathion by cyanobacteria. BioMed Research International 2014: 392681. https://doi.org/10.1155/2014/392682.CrossRefGoogle ScholarPubMed
Jaafari, J. & Yaghmaeian, K. (2019). Optimization of heavy metal biosorption onto freshwater algae (Chlorella coloniales) using response surface methodology (RSM). Chemosphere 217: 447455.CrossRefGoogle ScholarPubMed
Jin, M., Xiao, X., Qin, L. et al. (2020). Physiological and morphological responses and tolerance mechanisms of Isochrysis galbana to Cr(VI) stress. Bioresource Technology 302: 122860.CrossRefGoogle ScholarPubMed
Karydis, M. & Fogg, G. E. (1980). Physiological effects of hydrocarbons on the marine diatom Cyclotella cryptica. Microbial Ecology 6: 281290.CrossRefGoogle Scholar
Khatiwada, B., Hasan, M. T., Sun, A. et al. (2020). Proteomic response of Euglena gracilis to heavy metal exposure – identification of key proteins involved in heavy metal tolerance and accumulation. Algal Research 45: 101764.CrossRefGoogle Scholar
Lal, S. (1984). Effects of insecticides on algae. In Lal, R. (ed.) Insecticide Microbiology. Springer, Berlin, Heidelberg, pp. 203236.CrossRefGoogle Scholar
Lara, R. J., Wiencke, C. & Ernst, W. (1989). Association between exudates of brown algae and polychlorinated biphenyls. Journal of Applied Phycology 1: 267270.CrossRefGoogle Scholar
Leong, Y. K. & Chang, J.-S. (2020). Bioremediation of heavy metals using microalgae: Recent advances and mechanisms. Bioresource Technology 303: 122886.CrossRefGoogle ScholarPubMed
Liu, G., Jiang, R., You, J. et al. (2020). Microplastic impacts on microalgae growth: Effects of size and humic acid. Environmental Science and Technology 54: 17821789.CrossRefGoogle ScholarPubMed
Long, M., Moriceau, B., Gallinari, M. et al. (2015). Interactions between microplastics and phytoplankton aggregates: Impact on their respective fates. Marine Chemistry 175: 3946.CrossRefGoogle Scholar
Lu, J., Ma, Y., Xing, G. et al. (2019). Revelation of microalgae’s lipid production and resistance mechanism to ultra-high Cd stress by integrated transcriptome and physiochemical analyses. Environmental Pollution 250: 186195.CrossRefGoogle ScholarPubMed
Luo, L., Xiao, Z., Zhou, X. et al. (2020). Quantum chemical calculation to elucidate the biodegradation pathway of methylphenanthrene by green microalgae. Water Research 173: 115598.CrossRefGoogle ScholarPubMed
Machado, M. D. & Soares, E. V. (2019). Sensitivity of freshwater and marine green algae to three compounds of emerging concern. Journal of Applied Phycology 31: 399408.CrossRefGoogle Scholar
Madadi, R., Pourbabaee, A. A., Tabatabaei, M. et al. (2016). Treatment of petrochemical wastewater by the green algae Chlorella vulgaris. International Journal of Environmental Research 10: 555560.Google Scholar
Mahanty, H. K. & Gresshoff, P. M. (1978). Influence of polychlorinated biphenyls (PCBs) on growth of freshwater algae. Botanical Gazette 139: 202206.CrossRefGoogle Scholar
Mansano, A. S., Moreira, R. A., Dornfeld, H. C. et al. (2017). Effects of diuron and carbofuran and their mixtures on the microalgae Raphidocelis subcapitata. Ecotoxicology and Environmental Safety 142: 312321.CrossRefGoogle ScholarPubMed
Moisset, S., Tiam, S. K., Feurtet-Mazel, A. et al. (2015). Genetic and physiological responses of three freshwater diatoms to realistic diuron exposures. Environmental Science and Pollution Research 22: 40464055.CrossRefGoogle ScholarPubMed
Mostafa, F. I. Y. & Helling, C. S. (2002). Impact of four pesticides on the growth and metabolic activities of two photosynthetic algae. Journal of Environmental Science and Health – Part B Pesticides, Food Contaminants and Agricultural Wastes 37: 417444.Google ScholarPubMed
Muñoz, R., Guieysse, B. & Mattiasson, B. (2003). Phenanthrene biodegradation by an algal-bacterial consortium in two-phase partitioning bioreactors. Applied Microbiology and Biotechnology 61: 261267.CrossRefGoogle ScholarPubMed
Nandimandalam, H. & Gude, V. G. (2019). Indigenous biosensors for in situ hydrocarbon detection in aquatic environments. Marine Pollution Bulletin 149: 110643.CrossRefGoogle Scholar
Nie, J., Sun, Y., Zhou, Y. et al. (2020). Bioremediation of water containing pesticides by microalgae: Mechanisms, methods, and prospects for future research. Science of the Total Environment 707: 136080.CrossRefGoogle ScholarPubMed
Niestroy, J., Martínez, A. B. & Band-Schmidt, C. J. (2014). Analysis of concentration-dependent effects of copper and PCB on different Chattonella spp. microalgae (Raphidophyceae) cultivated in artificial seawater medium. EXCLI Journal 13: 197211.Google ScholarPubMed
Nolte, T. M., Hartmann, N. B., Kleijn, J. M. et al. (2017). The toxicity of plastic nanoparticles to green algae as influenced by surface modification, medium hardness and cellular adsorption. Aquatic Toxicology 183: 1120.CrossRefGoogle ScholarPubMed
Nowicka, B., Fesenko, T., Walczak, J. et al. (2020). The inhibitor-evoked shortage of tocopherol and plastoquinol is compensated by other antioxidant mechanisms in Chlamydomonas reinhardtii exposed to toxic concentrations of cadmium and chromium ions. Ecotoxicology and Environmental Safety 191: 110241.CrossRefGoogle ScholarPubMed
Pandey, L. K. (2020). In situ assessment of metal toxicity in riverine periphytic algae as a tool for biomonitoring of fluvial ecosystems. Environmental Technology and Innovation 18: 100675.CrossRefGoogle Scholar
Patel, J. G., Kumar, J. I. N., Kumar, R. N. et al. (2016). Biodegradation capability and enzymatic variation of potentially hazardous polycyclic aromatic hydrocarbons – Anthracene and Pyrene by Anabaena fertilissima. Polycyclic Aromatic Compounds 36: 7287.CrossRefGoogle Scholar
Pham, T. L., Dao, T. S., Bui, H. N. et al. (2020). Lipid production combined with removal and bioaccumulation of Pb by Scenedesmus sp. green alga. Polish Journal of Environmental Studies 29: 17851791.CrossRefGoogle Scholar
Pistocchi, R., Dao, L. T. H., Mikulic, P. et al. (2019). Metal pollution in water: Toxicity, tolerance and use of algae as a potential remediation solution. In Hallmann, A. & Rampelotto, P. H. (eds.) Grand Challenges in Algae Biotechnology. Springer, Cham, pp. 471500.CrossRefGoogle Scholar
Plastics Europe (2018). Plastics– the Facts 2018: An analysis of European plastics production, demand and waste data. Brussels, Wemmel: Association for Plastics Manufacturers, European Association of Plastics Recycling. Reteived from: https://plasticseurope.org/wp-content/uploads/2021/10/2018-Plastics-the-facts.pdf.Google Scholar
Pradhan, D., Sukla, L. B., Mishra, B. B. et al. (2019). Biosorption for removal of hexavalent chromium using microalgae Scenedesmus sp. Journal of Cleaner Production 209: 617629.CrossRefGoogle Scholar
Radwan, E. K., Abdel-Aty, A. M., El-Wakeel, S. T. et al. (2020). Bioremediation of potentially toxic metal and reactive dye-contaminated water by pristine and modified Chlorella vulgaris. Environmental Science and Pollution Research 27: 2177721789.CrossRefGoogle ScholarPubMed
Ricart, M., Barceló, D., Geiszinger, A. et al. (2009). Effects of low concentrations of the phenylurea herbicide diuron on biofilm algae and bacteria. Chemosphere 76: 13921401.CrossRefGoogle ScholarPubMed
Ridley, S. M. & Horton, P. (1984). DCMU-induced fluorescence changes and photodestruction of pigments associated with an inhibition of photosystem I cyclic electron flow. Zeitschrift für Naturforschung C 39: 351353.CrossRefGoogle Scholar
Romero-Lopez, J., Lopez-Rodas, V. & Costas, E. (2012). Estimating the capability of microalgae to physiological acclimatization and genetic adaptation to petroleum and diesel oil contamination. Aquatic Toxicology 124–125: 227237.CrossRefGoogle ScholarPubMed
Sabater, S., Artigas, J., Corcoll, N. et al. (2016). Ecophysiology of river algae. In Necchi Jr, O (ed.) River Algae. Springer International Publishing, Cham, pp. 197217.CrossRefGoogle Scholar
Salehi, M., Biria, D., Shariati, M. et al. (2019). Treatment of normal hydrocarbons contaminated water by combined microalgae – Photocatalytic nanoparticles system. Journal of Environmental Management 243: 116126.CrossRefGoogle ScholarPubMed
Sarmah, P. & Rout, J. (2018). Efficient biodegradation of low-density polyethylene by cyanobacteria isolated from submerged polyethylene surface in domestic sewage water. Environmental Science and Pollution Research 25: 3350833520.CrossRefGoogle ScholarPubMed
Senger, H. & Fleischhacker, P. (1978). Adaptation of the photosynthetic apparatus of Scenedesmus obliquus to strong and weak light conditions. I. Differences in pigments, photosynthetic capacity, quantum yield and dark reactions. Physiologia Plantarum 43: 3542.CrossRefGoogle Scholar
Shen, N. & Chirwa, E. M. N. (2020). Live and lyophilized fungi-algae pellets as novel biosorbents for gold recovery: Critical parameters, isotherm, kinetics and regeneration studies. Bioresource Technology 306: 123041.CrossRefGoogle ScholarPubMed
Shivaji, S. & Dronamaraju, S. V. L. (2019). Scenedesmus rotundus isolated from the petroleum effluent employs alternate mechanisms of tolerance to elevated levels of cadmium and zinc. Scientific Reports 9: 115.CrossRefGoogle ScholarPubMed
Sjollema, S. B., Redondo-Hasselerharm, P., Leslie, H. A. et al. (2016). Do plastic particles affect microalgal photosynthesis and growth? Aquatic Toxicology 170:259261.CrossRefGoogle ScholarPubMed
Skoglund, R. S. (1996). A kinetics model for predicting the accumulation of PCBs in phytoplankton. Environmental Science and Technology 30: 21132120.CrossRefGoogle Scholar
Staley, Z. R., Harwood, V. J. & Rohr, J. R. (2015). A synthesis of the effects of pesticides on microbial persistence in aquatic ecosystems. Critical reviews in Toxicology 45: 813836.CrossRefGoogle ScholarPubMed
Stratton, G. W. & Corke, C. T. (1982). Toxicity of the insecticide permethrin and some degradation products towards algae and cyanobacteria. Environmental Pollution. Series A, Ecological and Biological 29: 7180.CrossRefGoogle Scholar
Subashchandrabose, S. R., Venkateswarlu, K., Venkidusamy, K. et al. (2019). Bioremediation of soil long-term contaminated with PAHs by algal–bacterial synergy of Chlorella sp. MM3 and Rhodococcus wratislaviensis strain 9 in slurry phase. Science of the Total Environment 659: 724731.CrossRefGoogle ScholarPubMed
Swackhamer, D. L. & Skoglund, R. S. (1993). Bioaccumulation of PCBs by algae: Kinetics versus equilibrium. Environmental Toxicology and Chemistry 12: 831838.CrossRefGoogle Scholar
Tang, X., He, L. Y., Tao, X. Q. et al. (2010). Construction of an artificial microalgal-bacterial consortium that efficiently degrades crude oil. Journal of Hazardous Materials 181: 11581162.CrossRefGoogle ScholarPubMed
Upadhyay, A. K., Mandotra, S. K., Kumar, N. et al. (2016). Augmentation of arsenic enhances lipid yield and defense responses in alga Nannochloropsis sp. Bioresource Technology 221: 430437.CrossRefGoogle ScholarPubMed
Urrutia, C., Yañez-Mansilla, E. & Jeison, D. (2019). Bioremoval of heavy metals from metal mine tailings water using microalgae biomass. Algal Research 43: 101659.CrossRefGoogle Scholar
Vargo, G., Hutchins, M. & Almquist, G. (1982). The effect of low, chronic levels of no. 2 fuel oil on natural phytoplankton assemblages in microcosms: 1. Species composition and seasonal succession. Marine Environmental Research 6: 245264.CrossRefGoogle Scholar
Vimal Kumar, R., Kanna, G. R. & Elumalai, S. (2017). Biodegradation of polyethylene by green photosynthetic microalgae. Journal of Bioremediation and Biodegradation 8: 18.CrossRefGoogle Scholar
Wright, S. L., Thompson, R. C. & Galloway, T. S. (2013). The physical impacts of microplastics on marine organisms: A review. Environmental Pollution 178: 483492.CrossRefGoogle ScholarPubMed
Xaaldi Kalhor, A., Movafeghi, A., Mohammadi-Nassab, A. D. et al. (2017). Potential of the green alga Chlorella vulgaris for biodegradation of crude oil hydrocarbons. Marine Pollution Bulletin 123: 286290.CrossRefGoogle ScholarPubMed
Xing, Y., Lu, Y., Dawson, R. W. et al. (2005). A spatial temporal assessment of pollution from PCBs in China. Chemosphere 60: 731739.CrossRefGoogle ScholarPubMed
Zeroual, S., El Bakkal, S. E., Mansori, M. et al. (2020). Cell wall thickening in two Ulva species in response to heavy metal marine pollution. Regional Studies in Marine Science 35: 101125.CrossRefGoogle Scholar
Zhang, C., Chen, X., Wang, J. et al. (2017). Toxic effects of microplastic on marine microalgae Skeletonema costatum: Interactions between microplastic and algae. Environmental Pollution 220: 12821288.CrossRefGoogle ScholarPubMed
Zhang, H., Jiang, X., Lu, L. et al. (2015). Biodegradation of polychlorinated biphenyls (PCBs) by the novel identified cyanobacterium Anabaena PD-1. PLOS ONE 10: e0131450.Google ScholarPubMed
Zhao, Y., Shang, D., Ning, J. et al. (2019). Subcellular distribution and chemical forms of lead in the red algae, Porphyra yezoensis. Chemosphere 227: 172178.CrossRefGoogle ScholarPubMed
Zhu, X., Sun, Y., Zhang, X. et al. (2016). Herbicides interfere with antigrazer defenses in Scenedesmus obliquus. Chemosphere 162: 243251.CrossRefGoogle ScholarPubMed
Zhu, Z.-L., Wang, S.-C., Zhao, F.-F. et al. (2019). Joint toxicity of microplastics with triclosan to marine microalgae Skeletonema costatum. Environmental Pollution 246: 509517.CrossRefGoogle ScholarPubMed

References

Adams, D. G. & Duggan, P. (2008). Cyanobacteria-bryophyte symbioses. Journal of Experimental Botany 59: 10471058.CrossRefGoogle ScholarPubMed
Albertano, P., Luongo, L. & Grilli Caiola, M. (1991). Influence of different lights of mixed cultures of microalgae from ancient frescoes. International Biodeterioration 27: 2738.CrossRefGoogle Scholar
Arima, H., Horiguchi, N., Takaichi, S. et al. (2011). Molecular, genetic and chemotaxonomic characterization of the terrestrial cyanobacterium Nostoc commune and its neighbouring species. FEMS Microbiology Ecology 79: 3445.CrossRefGoogle Scholar
Baker, J. A., Entsch, B. & McKay, D. B. (2003). The cyanobiont in an Azolla fern is neither Anabaena nor Nostoc. FEMS Microbiology Letters 229: 4347.CrossRefGoogle Scholar
Belnap, J. (2005). CRUSTS | Biological. In Hillel, D. (ed.) Encyclopedia of Soils in the Environment. Elsevier, Amsterdam, pp. 339347. ISBN 9780123485304, https://doi.org/10.1016/B0-12-348530-4/00131-4.CrossRefGoogle Scholar
Belnap, J. & Lange, O. L. (2003). Biological soil crusts: Structure, function, and management. In Baldwin, I. T., Caldwell, , , M. M., Heldmaier, G. et al. (eds.) Ecological Studies Series, Vol. 150. Springer-Verlag, Berlin, p. 503.Google Scholar
Belnap, J., Weber, B. (2013). Biological soil crusts as an integral component of desert environments. Ecological Processes 2: 11. https://doi.org/10.1186/2192-1709-2-11.CrossRefGoogle Scholar
Beraldi-Campesi, H. (2013). Early life on land and the first terrestrial ecosystems. Ecological Processes 2: 1. https://doi.org/10.1186/2192-1709-2-1.CrossRefGoogle Scholar
Bergman, B., Johanson, C. & Söderbock, E. (1992). The Nostoc-Gunnera symbiosis. New Phytologist 122: 379400.CrossRefGoogle ScholarPubMed
Bergman, B., Matveyev, A. & Rasmussen, U. (1996). Chemical signalling in cyanobacterial-plant symbioses. Trends in Plant Science 1: 191197.CrossRefGoogle Scholar
Bergman, B. & Osborne, B. (2002). The Nostoc-Gunnera symbiosis. Biology and Environment: Proceedings of the Royal Irish Academy 102B: 3539.CrossRefGoogle Scholar
Bhattacharya, D., Pelletreau, K. N., Price, D. C. et al. (2013). Genome analysis of Elysia chlorotica egg DNA provides no evidence for horizontal gene transfer into the germ line of this kleptoplastidic mollusc. Molecular Biology and Evolution 30: 18431852.CrossRefGoogle Scholar
Bi, Y.-H., Deng, Z.-Y., Hu, Z.-Y. et al. (2005). Response of Nostoc flagelliforme to salt stress. Acta Hydrobiologica Sinica 29: 125129.CrossRefGoogle Scholar
Björk, M., Axelsson, L. & Beer, S. (2004). Why is Ulva intestinalis the only macroalga inhabiting isolated rockpools along the Swedish Atlantic coast? Marine Ecology Progress Series 284: 109116.CrossRefGoogle Scholar
Brinkhuis, H., Schouten, S., Collinson, M. et al. (2006). Episodic fresh surface waters in the Eocene Arctic Ocean. Nature 441: 606609.CrossRefGoogle ScholarPubMed
Büdel, B., Dulić, T., Darienko, T. et al. (2016). Cyanobacteria and algae of biological soil crusts. In: Weber, B., Büdel, B. &. Belnap, J. (eds.) Biological Soil Crusts: An Organizing Principle in Drylands. Ecological Studies, Vol. 226. Springer, Cham, pp. 5580. https://doi.org/10.1007/978-3-319-30214-0_4.CrossRefGoogle Scholar
Büdel, B., Vivas, M. & Lange, O. L. (2013). Lichen species dominance and the resulting photosynthetic behavior of Sonoran Desert soil crust types (Baja California, Mexico). Ecological Processes 2: 6. https://doi.org/10.1186/2192-1709-2-6.CrossRefGoogle Scholar
Bustos-Díaz, E. D., Barona-Gómez, F. & Cibrián-Jaramillo, A. (2019). Cyanobacteria in nitrogen-fixing symbioses. In: Mishra, A. K., Tiwari, D. N. & Rai, A. N. (eds.) Cyanobacteria. Academic Press, London, pp. 2942.CrossRefGoogle Scholar
Castenholz, R. W. (1969). Thermophilic blue-green algae and the thermal environment. Bacteriological Reviews 33: 476504.CrossRefGoogle ScholarPubMed
Chang, A. C. G., Chen, T., Li, N. & Duan, J. (2019). Perspectives on endosymbiosis in coralloid roots: Association of cycads and cyanobacteria. Frontiers in Microbiology 10: 1888. https://doi.org/10.3389/fmicb.2019.01888.CrossRefGoogle ScholarPubMed
Chrismas, N. A. M., Allen, R., Hollingworth, A. C. et al. (2021). Complex photobiont diversity in the marine lichen Lichina pygmaea. Journal of the Marine Biological Association of the United Kingdom 101: 667674.CrossRefGoogle Scholar
Cockell, C. S. & Herrera, A. (2008). Why are some microorganisms boring? Trends in Microbiology 16: 101106.CrossRefGoogle ScholarPubMed
Costa, J. L. & Lindblad, P. (2002). Cyanobacteria in symbiosis with cycads. In: Rai, A. N., Bergman, B. & Rasmussen, U. (eds.) Cyanobacteria in Symbiosis. Springer, Dordrecht, pp. 195205.Google Scholar
Couradeau, E., Roush, D., Guida, B. S. et al. (2017). Diversity and mineral substrate preference in endolithic microbial communities from marine intertidal outcrops (Isla de Mona, Puerto Rico). Biogeosciences 14:311324.CrossRefGoogle Scholar
Czerwik-Marcinkowska, J. & Mrozińska, T. (2011). Algae and cyanobacteria in caves of the Polish Jura. Polish Botanical Journal 56: 203243.Google Scholar
Dasgupta, C. N., Singh, K. V., Nayaka, S. et al. (2020). Molecular phylogeny of a commercially important thermophilic microalga Chlorella sorokiniana LWG002615 and associated bacterium Aquimonas sp. NBRI01 isolated from Jeori thermal spring, Shimla, India. The Nucleus 63: 203210.CrossRefGoogle Scholar
Davey, M. C. (1989). The effect of freezing and desiccation on photosynthesis and survival of terrestrial Antarctic algae and cyanobacteria. Polar Biology 10: 2936.CrossRefGoogle Scholar
De Bary, A. (1879). Die erscheinung der symbiose: Vortrag gehalten auf der versammlung deutscher naturforscher und aerzte zu cassel. Trübner.Google Scholar
Dettweiler-Robinson, E., Ponzetti, J. M. & Bakker, J. D. (2013). Long-term changes in biological soil crust cover and composition. Ecological Processes 2: 5.CrossRefGoogle Scholar
Doemel, W. N. & Brock, T. D. (1971). The physiological ecology of Cyanidium caldarium. Microbiology 67: 1732.Google Scholar
Fayolle, S., Moriconi, C., Oursel, B. et al. (2016). Epizoic algae distribution on the carapace and plastron of the European pond turtle (Emys orbicularis, Linnaeus, 1758): A study from the Camargue, France. Cryptogamie Algologie 37: 221232.CrossRefGoogle Scholar
Ford, T. W. (1986). Thermostability of the photosynthetic system of the thermoacidophilic alga Cyanidium caldarium in continuous culture. Journal of Experimental Botany 37: 16981707.CrossRefGoogle Scholar
Foster, R. A. & Zehr, J. P. (2019). Diversity, genomics, and distribution of phytoplankton-cyanobacterium single-cell symbiotic associations. Annual Review of Microbiology 73: 435456.CrossRefGoogle ScholarPubMed
Fujii, M., Takano, Y., Kojima, H. et al. (2010). Microbial community structure, pigment composition, and nitrogen source of red snow in Antarctica. Microbial Ecology 59: 466475.CrossRefGoogle Scholar
Galtier, N. & Lobry, J. R. (1997). Relationships between genomic G+C content, RNA secondary structures, and optimal growth temperature in prokaryotes. Journal of Molecular Evolution 44: 632636.CrossRefGoogle ScholarPubMed
Gao, X. & Zou, D. (2001). Photosynthetic bicarbonate utilization by a terrestrial cyanobacterium, Nostoc flagelliforme (Cyanophyceae). Journal of Phycology 37: 768771.CrossRefGoogle Scholar
Garcia-Meza, J. V., Barranguet, C. & Admiraal, W. (2005). Biofilm formation by algae as a mechanism for surviving on mine tailings. Environmental Toxicology and Chemistry 24: 573581.CrossRefGoogle ScholarPubMed
Garcia-Pichel, F., Ramírez-Reinat, E. & Gao, Q. (2010). Microbial excavation of solid carbonates powered by P-type ATPase-mediated transcellular Ca2+ transport. Proceedings of the National Academy of Sciences USA. 107: 2174921754.CrossRefGoogle ScholarPubMed
Gehringer, M. M., Pengelly, J. J., Cuddy, W. S. et al. (2010). Host selection of symbiotic cyanobacteria in 31 species of the Australian cycad genus: Macrozamia (Zamiaceae). Molecular Plant Microbe Interactions 23: 811822.CrossRefGoogle ScholarPubMed
Gimmler, H. (2001). Acidophilic and acidotolerant algae. In: Rai, L. C. & Gaur, J. P. (eds.) Algal Adaptation to Environmental Stresses. Springer, Berlin, Heidelberg, pp. 259290.CrossRefGoogle Scholar
Gimmler, H. & Degenhardt, B. (2001). Alkaliphilic and alkali-tolerant algae. In: Rai, L. C. & Gaur, J. P. (eds.) Algal Adaptation to Environmental Stresses. Springer, Berlin, Heidelberg, pp. 291321.CrossRefGoogle Scholar
Gomaa, F., Kosakayan, A., Heger, T. J. et al. (2014). One alga to rule them all: Unrelated mixotrophic testate amobae (Amoebozoa, Rhizaria and stramenopiles) share the same symbiont (Trebouxiophycae). Protist 165: 111126.CrossRefGoogle Scholar
Gordon, B. R. & Leggat, W. (2010). Symbiodinium-invertebrate symbioses and the role of metabolomics. Marine Drugs 8: 25462568.CrossRefGoogle ScholarPubMed
Grant, W. D., Mwatha, W. E. & Jones, B. E. (1990). Alkaliphiles: Ecology, diversity and applications. FEMS Microbiology Reviews 75: 255270.CrossRefGoogle Scholar
Grimm, M., Grube, M., Schiefelbein, U. et al. (2021). The lichens’ microbiota, still a mystery? Frontiers in Microbiology 12: 623839. https://doi.org/10.3389/fmicb.2021.623839.CrossRefGoogle ScholarPubMed
Grobbelaar, J. U. (2000). Lithophytic algae: A major threat to the karst formation of show caves. Journal of Applied Phycology 12: 309315.CrossRefGoogle Scholar
Grobbelaar, N., Scott, W. E., Hattingh, W. et al. (1987). The identification of the coralloid root endophytes of the southern African cycads and the ability of the isolates to fix dinitrogen. South African Journal of Botany 53: 111118.CrossRefGoogle Scholar
Gubernator, B., Bartoszewski, R., Kroliczewski, J. et al. (2008). Ribulose-1,5-bisphosphate carboxylase/oxygenase from thermophilic cyanobacterium Thermosynechococcus elongatus. Photosynthesis Research 95: 101109.CrossRefGoogle ScholarPubMed
Guida, B. S. & Garcia-Pichel, F. (2016). Extreme cellular adaptations and cell differentiation by a cyanobacterium for carbonate excavation. Proceedings of the National Academy of Science USA 113: 57125717.CrossRefGoogle ScholarPubMed
Handa, S., Nakahara, M., Tsubota, H. et al. (2006). Choricystis minor (Trebouxiophyceae, Chlorophyta) as a symbiont of several species of freshwater sponge. Hikobia 14: 265373.Google Scholar
Händeler, K., Grzymbowski, Y. P., Krug, P. J. et al. (2009). Functional chloroplasts in metazoan cells – a unique evolutionary strategy in animal life. Frontiers in Zoology 6: 28.CrossRefGoogle Scholar
Hayes, F. E., Codde, S. & Allen, S. G. (2022). Epizoic cyanobacteria and algae on the pelage of pinnipeds: A literature review and new data for the harbor seal (Phoca vitulina). Pacific Science 76: 6978.CrossRefGoogle Scholar
Hirooka, S. & Miyagishima, S.-Y. (2016). Cultivation of acidophilic algae Galdieria sulphuraria and Pseudochlorella sp. YKT1 in media derived from acidic hot springs. Frontiers in Microbiology 7: Article 2022. https://doi.org/10.3389/fmicb.2016.02022.CrossRefGoogle ScholarPubMed
Hoham, R. W. & Remias, D. (2020). Snow and glacial algae: A review. Journal of Phycology 56: 264282.CrossRefGoogle ScholarPubMed
Holzinger, A. & Karsten, U. (2013). Desiccation stress and tolerance in green algae: Consequences for ultrastructure, physiological, and molecular mechanisms. Frontiers in Plant Science 4: 327. https://doi.org/10.3389/fpls.2013.00327.CrossRefGoogle ScholarPubMed
Honegger, R., Edwards, D. & Axe, L. (2013). The earliest records of internally stratified cyanobacterial and algal lichens from the Lower Devonian of the Welsh Borderland. New Phytologist 197: 264275.CrossRefGoogle ScholarPubMed
Honegger, R. (1996). Morphogenesis. In: Nash, T. H. (ed.) Lichen Biology. Cambridge University Press, Cambridge, UK, pp. 6587.Google Scholar
Hoshina, F., Kobayashi, M., Suzaki, T. et al. (2018). Brandtia ciliaticola gen.et sp. nov. (Chlorellaceae, Trebouxiophyceae), a common symbiotic coccid of various ciliate species. Phycological Research 66: 7681.CrossRefGoogle Scholar
Huang, H., Bai, K. Z., Zhong, Z. P. et al. (2005). Energy transfer from phycobilisomes to photosystems of N. flagelliforme Born. et Filah during the rewetting course and its physiological significance. Journal of Integrative Plant Biology 47: 703708.CrossRefGoogle Scholar
Ionescu, D., Oren, A., Hindiyeh, M. Y. et al. (2007). The thermophilic cyanobacteria of the Zerka Ma’in thermal springs in Jordan. In: Seckbach, J. (ed.) Algae and Cyanobacteria in Extreme Environments. Cellular Origin, Life in Extreme Habitats and Astrobiology, Vol. 11. Springer, Dordrecht, pp. 411424.CrossRefGoogle Scholar
Jayanthi, D., Meghana, S. J., Rao, H. G. et al. (2023). Cyanobacterial symbioses with bryophytes. In: Dharmaduirai, D. (ed.) Microbial Symbioses. Function and Molecular Interactions on Host. Academic Press, London, pp. 1527.Google Scholar
Johnson, D. B. (1998). Biodiversity and ecology of acidophilic microorganisms. FEMS Microbiology Ecology 27: 307317.CrossRefGoogle Scholar
Johnson, M. D., Oldach, D., Delwiche, C. F. et al. (2007). Retention of transcriptionally active cryptophyte nuclei by the ciliate Myrionecta rubra. Nature 445: 426428.CrossRefGoogle ScholarPubMed
Johnson, M. D. (2011). The acquisition of phototrophy: Adaptive strategies of hosting endosymbionts and organelles. Photosynthesis Research 107: 117132.CrossRefGoogle ScholarPubMed
Karsten, U. & Holzinger, A. (2014). Green algae in alpine biological soil crust communities: Acclimation strategies against ultraviolet radiation and dehydration. Biodiversity and Conservation 23: 18451858.CrossRefGoogle ScholarPubMed
Karatygin, L. V., Snigirevskaya, N. S. & Vikulin, S. V. (2009). The most ancient terrestrial lichen Winfrenatia reticulata: A new find and new interpretation. Palaeontological Journal 43: 107112.CrossRefGoogle Scholar
Kedem, I., Milvad, Y., Kaplan, A. et al. (2021). Juggling lightning: How Chlorella ohadii handles extreme energy inputs without damage. Photosynthesis Research 147: 329344.CrossRefGoogle ScholarPubMed
Kerney, R., Kim, E., Hangaster, R. P. et al. (2011). Intracellular invasion of green algae in a salamander host. Proceedings of the National Academy of Sciences USA 108: 64976502.CrossRefGoogle Scholar
Kershaw, K. A. (1985). Physiological Ecology of Lichens. Cambridge University Press, Cambridge, UK.Google Scholar
Krzewicka, B., Smykla, J., Galas, J. et al. (2017). Freshwater lichens and habitat zonation of mountain streams. Limnologica 63: 110.CrossRefGoogle Scholar
Lajeunesse, T. C., Parkinson, J. E., Gabrielson, P. J. et al. (2018). Systematic revision of the Symbiodiniaceae highlighting the antiquity and diversity of coral endosymbionts. Current Biology 28: 25702580.CrossRefGoogle ScholarPubMed
Lao, P. J. & Forsdyke, D. R. (2000). Thermophilic bacteria strictly obey Szybalski’s transcription direction rule and politely purine-load RNAs with both adenine and guanine. Genome Research 10: 228236. https://doi.org/10.1101/gr.10.2.228.CrossRefGoogle ScholarPubMed
Larsson, C., Axelsson, L., Ryberg, H. et al. (1997). Photosynthetic carbon utilization by Enteromorpha intestinalis (Chlorophyta) from a Swedish rockpool. European Journal of Phycology 32: 4954.CrossRefGoogle Scholar
Lemoine, Y. & Schoefs, B. (2010). Secondary ketocarotenoid astaxanthin biosynthesis in algae: a multifunctional response to stress. Photosynthesis Research 106: 155177.CrossRefGoogle ScholarPubMed
Lenton, T. M. & Daines, S. J. (2017). Matworld – the biogeochemical effects of early life on land. New Phytologist 215: 531537.CrossRefGoogle ScholarPubMed
Letsch, M. R., Muller-Parker, G., Friedl, T. et al. (2009). Elliptochloris marina sp. nov. (Trebouxiophyceae, Chlorophyta), symbiotic green algae of the temperate Pacific sea anemones Anthopleura xanthorammica and A. longatum (Anthozoa, Cnidaria). Journal of Phycology 45: 11271135.CrossRefGoogle Scholar
Leu, J. Y., Lin, T. H., Selvamani, M. J. P. et al. (2013). Characterization of a novel thermophilic cyanobacterial strain from Taian hot springs in Taiwan for high CO2 mitigation and C-phycocyanin extraction. Process Biochemistry 48: 4148.CrossRefGoogle Scholar
Lewin, R. A. & Robinson, P. (1979). The greening of polar bears in zoos. Nature (Lond.) 278: 445447.CrossRefGoogle ScholarPubMed
Lewin, R. A. (1980). Green polar-bears, and other associations of algae with higher vertebrates. Trends in Biochemical Science 5: 34.CrossRefGoogle Scholar
Lewin, R. A., Farnsworth, P. A. & Yamanaka, G. (1981). The algae of green polar bears. Phycologia 20: 303314.CrossRefGoogle Scholar
Liang, Y., Kaczmarek, M. B., Kasprzak, A. K. et al. (2018). Thermosynechococcaceae as a source of thermostable C-phycocyanins: Properties and molecular insights. Algal Research 35: 223235.CrossRefGoogle Scholar
Liu, Y. H., Yu, L., Ke, W. T. et al. (2010). Photosynthetic recovery of Nostoc flagelliforme (Cyanophyceae) upon rehydration after 2 years and 8 years dry storage. Phycologia 49: 429437.CrossRefGoogle Scholar
Ma, S., Han, B., Huss, V. A. et al. (2015). Chlorella thermophila (Trebouxiophyceae, Chlorophyta), a novel thermo-tolerant Chlorella species isolated from an occupied rooftop incubator. Hydrobiologia 760: 8189.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
Macedo, M. F., Miller, A. Z., Dionísio, A. et al. (2009). Biodiversity of cyanobacteria and green algae on monuments in the Mediterranean Basin: An overview. Microbiology (Reading) 155: 34763490.CrossRefGoogle ScholarPubMed
Maeda, T., Takahashi, S., Yoshida, T. et al. (2021). Chloroplast acquisition without the gene transfer in kleptoplastidic sea slugs, Plakobrankus ocellatus. eLife 10: e60176.Google Scholar
Malar, C. M., Krüger, M., Krüger, C. et al. (2021). The genome of Geosiphon pyriformis reveals ancestral traits linked to the arbuscular mycorrhizal symbiosis. Current Biology 31: 15701577.CrossRefGoogle Scholar
Mandal, S. & Rath, J. (2013). Algal colonization and its ecophysiology on the fine sculptures of terracotta monuments of Bishnupur, West Bengal, India. International Biodeterioration and Biodegradation 84: 291e299.CrossRefGoogle Scholar
Mann, D. G., Yamada, N., Bolton, J. J. et al. (2023). Nitzschia captivata sp. nov. (Bacillariophyta), the essential prey diatom of the kleptoplastidic dinoflagellate Durinska capensis, compared with N. agnita, N. kuetzungiioides and other species. Phycologia 62: 136151.CrossRefGoogle Scholar
Mansour, J. S. & Anstesia, K. (2021). Eco-evolutionary perspectives on mixoplankton. Frontiers in Marine Science 8: 666180.CrossRefGoogle Scholar
Marsh, J., Nouvet, S., Sanborn, P. et al. (2006). Composition and function of biological crust communities along topographic gradients in grasslands of central interior British Columbia (Chilcotin) and southwestern Yukon (Kluane). Canadian Journal of Botany 84: 713731.CrossRefGoogle Scholar
Maselli, M., Anastis, K., Klemm, K. et al. (2021). Retention of prey genetic material by the kleptoplastidic ciliate Strombidium cf. basimorphum. Frontiers in Microbiology 12: 694508.CrossRefGoogle ScholarPubMed
Mattson, M. D. (1999). Acid lakes and rivers. In: Environmental Geology. Encyclopedia of Earth Science. Springer, Dordrecht. https://doi.org/10.1007/1-4020-4494-1_4.Google Scholar
MeGraw, V. E., Brown, A. R., Boothman, C. et al. (2018). A novel adaptation mechanism underpinning algal colonization of a nuclear fuel storage pond. mBio 9: e02395–17. https://doi.org/10.1128/mBio.02395-17.CrossRefGoogle ScholarPubMed
Mikulic, P. & Beardall, J. (2014). Contrasting toxicity effects of zinc on growth and photosynthesis in a neutrophilic alga (Chlamydomonas reinhardtii) and an extremophilic alga (Cyanidium caldarium). Chemosphere 112: 402411.CrossRefGoogle Scholar
Mikulic, P. & Beardall, J. (2021). Oxidative and anti-oxidative responses to metal toxicity in an extremophilic alga (Cyanidium caldarium) and a neutrophilic alga (Chlamydomonas reinhardtii). Phycologia 60: 513523.CrossRefGoogle Scholar
Mock, T. & Junge, K. (2007). Psychrophilic diatoms. In: Seckbach, J. (ed.) Algae and Cyanobacteria in Extreme Environments. Cellular Origin, Life in Extreme Habitats and Astrobiology, Vol. 11. Springer, Dordrecht, pp. 344364.Google Scholar
Mohamed, A. R., Cumbo, V. R., Harii, S. et al. (2018). Deciphering the nature of the coral–Chromera association. ISME Journal 12: 776790.CrossRefGoogle ScholarPubMed
Møller, C. L., Vangsoe, M. T. & Sand-Jensen, K. (2014). Comparative growth and metabolism of gelatinous colonies of three phytoplankton, Nostoc commune, Nostoc pruniforme and Nostoc zetterstedtii, at different temperatures. Freshwater Biology 59: 21832193.CrossRefGoogle Scholar
Morgan-Kiss, R. M., Priscu, J. C., Pocock, T. et al. (2006). Adaptation and acclimation of photosynthetic microorganisms to permanently cold environments. Microbiology and Molecular Biology Reviews 70: 222–52.CrossRefGoogle ScholarPubMed
Mutalipassi, M., Riccio, G., Mazella, V. et al. (2021). Symbioses of cyanobacteria in marine environments: Ecological insights and biotechnological perspectives. Marine Drugs MDPI 19: 227.CrossRefGoogle ScholarPubMed
Nixdorf, B., Mischke, U. & Lessmann, D. (1998). Chrysophytes and chlamydomonads: Pioneer colonists in extremely acidic mining lakes (pH<3) in Lusatia (Germany). Hydrobiologia 369/370: 315327.CrossRefGoogle Scholar
Novis, P. M. & Harding, J. S. (2007). Extreme acidophiles. In: Seckbach, J. (ed.) Algae and Cyanobacteria in Extreme Environments. Cellular Origin, Life in Extreme Habitats and Astrobiology, Vol. 11. Springer, Dordrecht, pp. 443463.CrossRefGoogle Scholar
Nakayama, T., Ikegami, Y., Nakayama, T. et al. (2011). Spheroid bodies in rhopalodiacean diatoms were derived from a single endosymbiotic cyanobacterium. Journal of Plant Research 124: 9397.CrossRefGoogle ScholarPubMed
Nakayama, T., Kamikawak, R., Tanifujik, G. et al. (2014). Complete genome of a nonphotosynthetic cyanobacterium in a diatom reveals recent adaptation to an intracellular lifestyle. Proceedings of the National Academy of Sciences USA 111: 1140711412.CrossRefGoogle Scholar
Nelsen, M. P., Lücking, R., Boyce, C. K. et al. (2020). No support for emergence of lichens prior to the evolution of vascular plants. Geobiology 18: 313.CrossRefGoogle Scholar
Nordstrom, D. K., McCleskey, R. B. & Ball, J. W. (2008). Sulfur geochemistry of hydrothermal waters in Yellowstone National Park: IV Acid–sulfate waters. Applied Geochemistry 24: 191207.CrossRefGoogle Scholar
Oborník, M. (2020). Photoparasitism as an intermediate state in the evolution of apicomplexan parasites. Trends in Parasitology 36: 727734.CrossRefGoogle Scholar
Osawa, Y. & Tokeshi, M. (2020). Niche relations in a small world: Epizoic algae on an intertidal gastropod. Population Ecology 62: 395407.CrossRefGoogle Scholar
Pauli, J. N., Mendoza, J. E., Steffan, S. A. et al. (2014). A syndrome of mutualism reinforces the lifestyle of a sloth. Proceedings of the Royal Society of London B 281: 20133006. http://dx.doi.org/10.1098/rspb.2013.3006.Google ScholarPubMed
Peters, G. A. (1991). Azolla and other plant-cyanobacteria symbioses: Aspects of form and function. Plant Soil 137: 2536.CrossRefGoogle Scholar
Pistocchi, R., Dao, L. T. H., Mikulic, P. et al. (2019). Metal pollution in water: Toxicity, tolerance and use of algae as a potential remediation solution. In: Hallmann, A. & Rampellotto, P. H. (eds.) Grand Challenges in Algae Biotechnology, Springer, Cham, pp. 471500.CrossRefGoogle Scholar
Qiu, B. S., Zhong, A. H., Zhou, W. B. et al. (2004). Effects of potassium on the photosynthetic recovery of the terrestrial cyanobacterium, Nostoc flagelliforme (Cyanophyceae) during rehydration. Journal of Phycology 40: 323332.CrossRefGoogle Scholar
Quigg, A., Kotabová, E., Jarešová, J. et al. (2012). Photosynthesis in Chromera velia represents a simple system with high efficiency. PLOS ONE 7: e47036. https://doi.org/10.1371/journal.pone.0047036.CrossRefGoogle ScholarPubMed
Rajević, N., Kovačvić, M., Gould, S. B. et al. (2015). Algal endosymbionts in European Hydra strains reflect multiple origins of the zoochlorella symbiosis. Molecular Phylogenetics and Evolution 93: 5562.CrossRefGoogle ScholarPubMed
Raven, J. A. (2002). Evolution of cyanobacterial symbioses. Biology and environment. Proceedings of the Royal Irish Academy 102B: 36.CrossRefGoogle Scholar
Raven, J. A., Suggett, D. J. & Giordano, M. (2020). Inorganic carbon concentrating mechanisms in free-living and symbiotic dinoflagellates and chromerids. Journal of Phycology 56: 13771392.CrossRefGoogle ScholarPubMed
Raven, J. A. (2023). Distribution and function of calcium deposits in photosynthetic organisms. In Lüttge, U., Cánovas, F. M., Risueño, M. C., Leuschner, C. & Pretzsch, H. (eds.) Progress in Botany, Vol. 84. Springer, Cham, pp. 293326.Google Scholar
Raven, J. A. & Allen, J. F. (2003). Genomics and chloroplast evolution: What did cyanobacteria do for plants? Genome Biology, 4(3): 209215.CrossRefGoogle ScholarPubMed
Raymond, J. A. & Morgan-Kiss, R. (2013). Separate origins of ice-binding proteins in Antarctic Chlamydomonas species. PLoS One 8. https://doi.org/10.1371/journal.pone.0059186.CrossRefGoogle ScholarPubMed
Raymond, J. A., Morgan-Kiss, R. & Stahl-Rommel, S. (2020). Glycerol as an osmoprotectant in two Antarctic Chlamydomonas species from an ice-covered saline lake and is synthesized by an unusual bidomain enzyme. Frontiers in Plant Science 11: 1259. https://doi.org/10.3389/fpls.2020.01259.CrossRefGoogle ScholarPubMed
Raymond, J. A. & Remias, D. (2019). Ice-binding proteins in a chrysophycean snow alga: Acquisition of an essential gene by horizontal gene transfer. Frontiers in Microbiology 10:2697. https://doi.org/10.3389/fmicb.2019.02697.CrossRefGoogle Scholar
Rehman, M., Varshney, S., Ravi, L. et al. (2023). Cyanobacterial symbioses with angiosperms. In: Dharmaduirai, D. (ed.) Microbial Symbioses. Function and Molecular Interactions on Host. Academic Press, London, pp. 3956.Google Scholar
Rezayian, M., Niknam, V. & Ebrahimzadeh, H. (2019). Oxidative damage and antioxidative system in algae. Toxicology Reports 24: 13091313. https://doi.org/10.1016/j.toxrep.2019.10.001.CrossRefGoogle Scholar
Ritchie, R. J. & Sma-Air, S. (2022). Photosynthesis of an endolithic Chlorococcum alga (Chlorophyta, Chlorococcaceae) from travertine calcium carbonate rocks of a tropical limestone spring. Applied Phycology 3: 115.CrossRefGoogle Scholar
Rivasseau, C., Farhi, E., Compagnon, E. et al. (2016). Coccomyxa actinabiotis sp. nov. (Trebouxiophyceae, Chlorophyta), a new green microalga living in the spent fuel cooling pool of a nuclear reactor. Journal of Phycology 52: 689703.CrossRefGoogle ScholarPubMed
Rothschild, L. J. & Mancinelli, R. L. (2001). Life in extreme environments. Nature 409: 10921101.CrossRefGoogle ScholarPubMed
Roubeix, V., Attia, L., Chavaux, R. et al. (2021). Specificity of diatom communities attached on the carapace of the European pond turtle Emys orbicularis. Advances in Oceanography and Limnology 12: https://doi.org/10.4081/aiol.2021.9119.CrossRefGoogle Scholar
Saini, N., Pal, K., Sujata et al. (2021). Thermophilic algae: A new prospect towards environmental sustainability. Journal of Cleaner Production 324: 129277. https://doi.org/10.1016/j.jclepro.2021.129277.CrossRefGoogle Scholar
Samolov, E., Baumann, K., Büdel, B. et al. (2020). Biodiversity of algae and cyanobacteria in biological soil crusts collected along a climatic gradient in Chile using an integrative approach. Microorganisms. 8: 1047. https://doi.org/10.3390/microorganisms8071047.CrossRefGoogle ScholarPubMed
Sand-Jensen, K. & Jespersen, T. S. (2012). Tolerance of the widespread cyanobacterium Nostoc commune to extreme temperature variations (−269 to 105°C), pH and salt stress. Oecologia 169: 331339.CrossRefGoogle ScholarPubMed
Sato, N. (2020). Complex origins of chloroplast membranes with photosynthetic machineries: Multiple transfers of genes from divergent organisms at different times or a single endosymbiotic event? Journal of Plant Research 133: 1533.CrossRefGoogle ScholarPubMed
Saunders, R. M. K. & Fowler, K. (1993). The supraspecific taxonomy and evolution of the fern genus Azolla (Azollaceae). Plant Systematics and Evolution 184: 175193.CrossRefGoogle Scholar
Satoh, K., Hirai, M., Nishio, J. et al. (2002). Recovery of photosynthetic systems during rewetting is quite rapid in a terrestrial cyanobacterium. Plant and Cell Physiology 43: 170176.CrossRefGoogle Scholar
Scherer, W. S., Ernst, A., Chen, T.-W. et al. (1984). Rewetting of drought-resistant blue-green algae: Time-course of water uptake and reappearance of respiration, photosynthesis and nitrogen fixation. Oecologia 62: 418423.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
Simpson, C., Kiesling, W., Mewis, H. et al. (2011). Evolutionary diversification of reef corals: comparison of the molecular and fossil records. Evolution 65: 32743284.CrossRefGoogle ScholarPubMed
Souza-Egipsy, V., Altamirano, M., Amils, R. et al. (2011). Photosynthetic performance of phototrophic biofilms in extreme acidic environments. Environmental Microbiology 13: 23512358.CrossRefGoogle ScholarPubMed
Speelman, E. N., van Kempen, M. M. L., Barke, J. et al. (2009). The Eocene Arctic Azolla bloom: Environmental conditions, productivity and carbon drawdown. Geobiology 7: 155170.CrossRefGoogle ScholarPubMed
Spribille, T., Tagirdzhanova, G., Goyette, S. et al. (2020). 3D biofilms: In search of the polysaccharides holding together lichen symbioses. FEMS Microbiology Letters 367: 5. https://doi.org/10.1093/femsle/fnaa023.CrossRefGoogle ScholarPubMed
Stat, M., Carter, D. & Hoegh-Guldberg, O. (2006). The evolutionary history of Symbiodinium and scleractinian hosts – Symbiosis, diversity, and the effect of climate change. Perspectives in Plant Ecology, Evolution and Systematics 8: 2343.CrossRefGoogle Scholar
Steinberg, C. E. W., Schäfer, H. & Beisker, W. (1998). Do acid-tolerant cyanobacteria exist? Acta Hydrochimica et Hydrobiologica 26: 1319.3.0.CO;2-V>CrossRefGoogle Scholar
Szyja, M., Büdel, B. & Colesie, C. (2018). Ecophysiological characterization of early successional biological soil crusts in heavily human-impacted areas. Biogeosciences 15: 19191931.CrossRefGoogle Scholar
Takai, K., Nakamura, K., Toki, T., et al. (2008). Cell proliferation at 122°C and isotopically heavy CH4 production by a hyperthermophilic methanogen under high pressures cultivation. Proceedings of the National Academy of Sciences USA 105: 1094910954.CrossRefGoogle Scholar
Talling, J. F., Wood, R. B., Prosser, M. V. et al. (1973). The upper limit of photosynthetic productivity by phytoplankton: Evidence from Ethiopian soda lakes. Freshwater Biology 3: 5376.CrossRefGoogle Scholar
Tamaru, Y., Takani, Y., Yoshida, T. et al. (2005). Crucial role of extracellular polysaccharides in desiccation and freezing tolerance in the terrestrial cyanobacterium Nostoc commune. Applied and Environmental Microbiology 71: 73277333.CrossRefGoogle ScholarPubMed
Taranto, P. A., Keenan, T. W. & Potts, M. (1993). Rehydration induces rapid onset of lipid biosynthesis in desiccated Nostoc commune (Cyanobacteria). Biochimica et Biophysica Acta 168: 228237.CrossRefGoogle Scholar
Teugels, B., Bouillon, S., Veuger, B. et al. (2008). Kleptoplastids mediate nitrogen acquisition in the sea slug Elysia viridis. Aquatic Botany 4: 1521.CrossRefGoogle Scholar
Thomas, N. J., Coates, C. J., Kam, W. et al. (2023). Environmental constraints on the photosynthetic rate of the marine flatworm Symsagittifera roscoffensis. Journal of Experimental Marine Biology and Ecology 558: 51830. https://doi.org/10.1016/j.jembe.2022.151830.CrossRefGoogle Scholar
Treves, H., Raanan, H., Finkel, M. O. et al. (2013). A newly isolated Chlorella sp. from desert sand crusts exhibits a unique resistance to excess light intensity. FEMS Microbiology Ecology 86: 373380.CrossRefGoogle ScholarPubMed
Treves, H., Raanan, H., Kedem, I. et al. (2016). The mechanisms whereby the green alga Chlorella ohadii, isolated from desert soil crust, exhibits unparalleled photodamage resistance. New Phytologist 210: 12291243.CrossRefGoogle ScholarPubMed
Van Oppen, M. J. H. & Medina, M. (2020). Coral evolutionary responses to microbial symbioses. Philosophical Transactions of the Royal Society B 375: 20190591.CrossRefGoogle ScholarPubMed
Varshney, P., Beardall, J. & Wangikar, P. (2018). Isolation and biochemical characterisation of two novel thermophilic green algal species- Asterarcys quadricellulare and Chlorella sorokiniana, which are tolerant to high levels of carbon dioxide and nitric oxide. Algal Research 30: 2837.CrossRefGoogle Scholar
Varshney, P., Sohoni, S., Wangikar, P. P. et al. (2016). Effect of high CO2 concentrations on the growth and macromolecular composition of a heat- and high-light–tolerant microalga. Journal of Applied Phycology 28: 26312640.CrossRefGoogle Scholar
Varshney, P., Mikulic, P., Vonshak, A. et al. (2015). Extremophilic micro-algae and their potential contribution in biotechnology. Bioresource Technology 184: 363372.CrossRefGoogle ScholarPubMed
Wägele, H., Deusch, O., Händeler, K. et al. (2011). Transcription evidence that longevity of acquired plastids in the photosynthetic slugs Elysia timida and Plakybranchus ocellatus does not entail lateral transfer of algal nuclear genes. Molecular Biology and Evolution 28: 699706.CrossRefGoogle Scholar
Wagner, G. M. (1997). Azolla: A review of its biology and utilization. The Botanical Review 63: 126.CrossRefGoogle Scholar
Wyness, A. J., Roush, D. & McQuaid, C. D. (2022). Global distribution and diversity of marine euendolithic cyanobacteria. Journal of Phycology 58: 746759.CrossRefGoogle ScholarPubMed
Xue, S., Zang, Y., Chen, J. et al. (2020). Ultraviolet-B radiation stress triggers reactive oxygen species and regulates the antioxidant defense and photosynthesis systems of intertidal red algae Neoporphyra haitanensis. Frontiers in Marine Science 9. https://doi.org/10.3389/fmars.2022.1043462.Google Scholar
Yaguchi, T., Chung, S., Igarashi, Y. et al. (1992). Purification of RuBisCO from the thermophilic cyanobacterium Synechococcus sp. strain a-1. Journal of Fermentation Bioengineering 73: 348351.CrossRefGoogle Scholar
Yaguchi, T., Chung, S. Y., Igarashi, Y. et al. (1993). Cloning, sequence and overexpression of the thermophilic cyanobacterium gene for the ribulose-1,5-bisphosphate carboxylase/oxygenase. Journal of Fermentation Bioengineering 75: 18.CrossRefGoogle Scholar
Yang, H., Genot, B., Duhamel, S. et al. (2022). Organismal and cellular interactions in vertebrate-alga symbiosis. Biochemical Society Transactions 50: 609620.CrossRefGoogle Scholar
Yong-Hong, B., Zhong-Yang, D., Zheng-Yu, H., et al. (2005). Response of Nostoc flagelliforme to salt stress. Acta Hydrobiologica Sinica 29: 125129.Google Scholar
Zheng, Y., Xue, C., Chen, H. et al. (2020). Low-temperature adaptation of the snow alga Chlamydomonas nivalis is associated with the photosynthetic system regulatory process. Frontiers in Microbiology 11: 1233. https://doi.org/10.3389/fmicb.2020.01233.CrossRefGoogle ScholarPubMed

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
×