Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-05-01T08:06:25.903Z Has data issue: false hasContentIssue false

Chapter Eight - The ecophysiology of responding to change in polar marine benthos

from Part III - Life in extreme environments and the responses to change: the example of polar environments

Published online by Cambridge University Press:  28 September 2020

Guido di Prisco
Affiliation:
National Research Council of Italy
Howell G. M. Edwards
Affiliation:
University of Bradford
Josef Elster
Affiliation:
University of South Bohemia, Czech Republic
Ad H. L. Huiskes
Affiliation:
Royal Netherlands Institute for Sea Research
Get access

Summary

Environments vary over all spatial scales, from subatomic to the universe and all timescales from a Plank unit to eternity. There have been very large changes in environments in the past, notably over hundreds to thousands of years during glaciations driven by Milankovitch cycles (Zachos et al., 2001). Life on Earth has survived and evolved over time to adapt to those changes and to exploit nearly all of the available environments on Earth. The very small exceptions include the insides of hydrothermal vents

Type
Chapter
Information
Life in Extreme Environments
Insights in Biological Capability
, pp. 184 - 217
Publisher: Cambridge University Press
Print publication year: 2020

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

Abele, D., Tesch, C., Wencke, P., Pörtner, H.O. (2001). How does oxidative stress relate to thermal tolerance in the Antarctic bivalve Yoldia eightsi? Antarctic Science, 13, 111118.Google Scholar
Amaral Zettler, L.A., Messerli, M.A., Laatsch, A.D., Smith, P.J.S., Sogin, M.L. (2003). From genes to genomes: beyond biodiversity in Spain’s Rio Tinto. The Biological Bulletin, 204, 205209.CrossRefGoogle ScholarPubMed
Arnaud, P.M. (1974). Contribution à la bionomie marine benthique des régions antarctiques et subantarctiques. Téthys, 6, 567653.Google Scholar
Arntz, W.E., Brey, T., Gallardo, V.A. (1994). Antarctic zoobenthos. Oceanography and Marine Biology: An Annual Review, 32, 241304.Google Scholar
Aronson, R.B., Thatje, S., Clarke, A., et al. (2007). Climate change and invisibility of the Antarctic benthos. Annual Review of Ecology, Evolution, and Systematics, 38, 129154.Google Scholar
Ashton, G., Morley, S.A., Barnes, D.K.A., Clark, M.S., Peck, L.S. (2017). Warming by 1°C drives species and assemblage level responses in Antarctica’s marine shallows. Current Biology, 27, 26982705.Google Scholar
Bailey, A., Thor, P., Browman, H.I., et al. (2016). Early life stages of the Arctic copepod Calanus glacialis are unaffected by increased seawater pCO2. ICES Journal of Marine Science, 74, 9961004.Google Scholar
Barnes, D.K.A. (2016). Iceberg killing fields limit huge potential for benthic blue carbon in Antarctic shallows. Global Change Biology, 23, 26492659.Google Scholar
Barnes, D.K.A., Conlan, K.E. (2012). The dynamic mosaic. In: Rogers et al. (eds) Antarctic Ecosystems: An Extreme Environment in a Changing World. Wiley Interscience, pp. 255290.Google Scholar
Barnes, D.K.A., Fuentes, V., Clarke, A., Schloss, I.R., Wallace, M.I. (2006). Spatial and temporal variation in shallow seawater temperatures around Antarctica. Deep-Sea Research II, 53, 853858.Google Scholar
Beers, J.M., Jayasundara, M. (2014). Antarctic notothenioid fish: what are the future consequences of ‘losses’ and ‘gains’ acquired during long-term evolution at cold and stable temperatures? Journal of Experimental Biology, 218, 18341845.Google Scholar
Bergmann, C. (1847). Über die verhältnisse der wärmeökonomie der thiere zu ihrer grösse. Göttinger Studien, 3, 595708.Google Scholar
Bilyk, K.T., DeVries, A.L. (2011). Heat tolerance and its plasticity in Antarctic fishes. Comparative Biochemistry and Physiology, 158, 382390.CrossRefGoogle ScholarPubMed
Blackburn, T.M., Gaston, K.J., Loder, N. (1999). Geographic gradients in body size: a clarification of Bergmannʼs Rule. Diversity and Distributions, 5, 165174.Google Scholar
Brown, K.M., Fraser, K.P.P., Barnes, D.K.A., Peck, L.S. (2004). Ice scour frequency dictates Antarctic shallow-water community structure. Oecologia, 141, 121129.Google Scholar
Butler, P.G., Wanamaker, A.D., Scourse, J.D., Richardson, C.A., Reynolds, D.J. (2013). Variability of marine climate on the North Icelandic Shelf in a 1357-year proxy archive based on growth increments in the bivalve Arctica islandica. Palaeogeography, Palaeoclimatology, Palaeoecology, 373, 141151.Google Scholar
Byrne, M. (2011). Impact of ocean warming and ocean acidification on marine invertebrate life-history stages: vulnerabilities and potential for persistence in a changing ocean. Oceanography and Marine Biology: An Annual Review, 49, 142.Google Scholar
Caldeira, K., Wickett, M.E. (2003). Anthropogenic carbon and ocean pH. Nature, 425, 365.Google Scholar
Cerrone, D., Fusco, G., Simmonds, I., Aulicino, G., Budillon, G. (2017). Dominant covarying climate signals in the Southern Ocean and Antarctic sea ice influence during the last three decades. Journal of Climate, 30, 30553072.Google Scholar
Chapelle, G., Peck, L.S. (1999). Polar gigantism dictated by oxygen availability. Nature, 399, 144145.Google Scholar
Chapelle, G., Peck, L.S. (2004). Amphipod crustacean size spectra: new insights in the relationship between size and oxygen. Oikos, 106, 167175.Google Scholar
Chen, L., DeVries, A.L., Cheng, C.-H.C. (1997). Evolution of antifreeze glycoprotein gene from a trypsinogen gene in Antarctic notothenioid fish. Proceedings of the National Academy of Sciences of the USA, 94, 38113816.Google Scholar
Cheng, C.-H.C., Detrich III, H.W. (2012). Molecular ecophysiology of Antarctic notothenioid fishes. In: Rogers, A et al. (eds). Antarctic Ecosystems: An Extreme Environment in a Changing World. Wiley Interscience, pp. 357378.Google Scholar
Chivers, D.P., McCormik, M.I., Nilsson, G.E., et al. (2014). Impaired learning of predators and lower prey survival under elevated CO2: a consequence of neurotransmitter interference. Global Change Biology, 20, 515522.Google Scholar
Clark, M.S., Peck, L.S. (2009). HSP70 heat shock proteins and environmental stress in Antarctic marine organisms: a mini-review. Marine Genomics, 2, 1118.Google Scholar
Clark, M.S., DuPont, S., Rosetti, H., et al. (2007). Delayed arm regeneration in the Antarctic brittle star (Ophionotus victoriae). Aquatic Biology, 1, 4553.Google Scholar
Clark, M.S., Fraser, K.P.P., Peck, L.S. (2008a). Antarctic marine molluscs do have an HSP70 heat shock response. Cell Stress and Chaperones, 13, 3949.Google Scholar
Clark, M.S., Fraser, K.P.P., Burns, G., Peck, L.S. (2008b). The HSP70 heat shock response in the Antarctic fish Harpagifer antarcticus. Polar Biology, 31, 171180.Google Scholar
Clark, M.S, Fraser, K.P.P.F., Peck, L.S. (2008c). Antarctic marine molluscs do have an HSP70 heat shock response. Cell Stress and Chaperones, 13, 3949.CrossRefGoogle ScholarPubMed
Clark, M.S., Fraser, K.P.P., Peck, L.S. (2008d). Lack of an HSP70 heat shock response in two Antarctic marine invertebrates. Polar Biology, 31, 10591065.Google Scholar
Clark, M.S., Husmann, G., Thorne, M.A.S., et al. (2013). Hypoxia impacts large adults first: consequences in a warming world. Global Change Biology, 19, 22512263.Google Scholar
Clark, M.S., Sommer, U., Kaur, J., et al. (2017). Biodiversity in marine invertebrate responses to acute warming revealed by a comparative multi-omics approach. Global Change Biology, 23, 318330.Google Scholar
Clark, M.S., Thorne, M.A., Burns, G., Peck, L.S. (2016). Age-related thermal response: the cellular resilience of juveniles. Cell Stress and Chaperones, 21, 7585.CrossRefGoogle ScholarPubMed
Clarke, A. (1979). On living in cold water: K strategies in Antarctic benthos. Marine Biology, 55, 111119.Google Scholar
Clarke, A., Crame, J.A. (1992). The Southern Ocean benthic fauna and climate change – a historical perspective. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 338, 299309.Google Scholar
Clarke, A., Crame, J.A. (2010). Evolutionary dynamics at high latitudes: speciation and extinction in polar marine faunas. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 365, 36553666.Google Scholar
Clarke, A., Gaston, K.J. (2006). Climate, energy and diversity. Proceedings of the Royal Society of London Series B: Biological Sciences, 273, 22572266.Google Scholar
Clarke, A., Johnston, N. (1999). Scaling of metabolic rate and temperature in teleost fish. Journal of Animal Ecology, 68, 893905.Google Scholar
Clarke, A., Holmes, L.J., White, M.G. (1988). The annual cycle of temperature, chlorophyll and major nutrients at Signy Island, South Orkney Islands, 1969–1982. British Antarctic Survey Bulletin, 80, 6586.Google Scholar
Clarke, A., Meredith, M.P., Wallace, M.I., Brandon, M.I., Thomas, D.N. (2008). Seasonal and interannual variability in temperature, chlorophyll and macronutrients in northern Marguerite Bay, Antarctica. Deep-Sea Research II, 55, 1988–2006.Google Scholar
Clarke, A., Murphy, E.J.M., Meredith, M.P., et al. (2007). Climate change and the marine ecosystem of the western Antarctic Peninsula. Philosophical Transactions of the Royal Society of London B: Biological Sciences, 362, 149166.Google Scholar
Clusella-Trullas, S., Boardman, L., Faulkner, K.T., Peck, L.S., Chown, S.L. (2014). Effects of temperature on heat-shock responses and survival of two species of marine invertebrates from sub-Antarctic Marion Island. Antarctic Science, 26, 145152.Google Scholar
Collins, S. (2012). Marine microbiology: evolution on acid. Nature Geoscience, 5, 310311.Google Scholar
Cook, A.J., Vaughan, D.G. (2010). Overview of areal changes of the ice shelves on the Antarctic Peninsula over the past 50 years. The Cryosphere, 4, 7798.Google Scholar
Cook, A.J., Fox, A.J., Vaughan, D.G., Ferrigno, J.G. (2005). Retreating glacier fronts on the Antarctic Peninsula over the past half-century. Science, 308, 541544.Google Scholar
Cross, E.L., Peck, L.S., Harper, E.M. (2015). Ocean acidification does not impact shell growth or repair of the Antarctic brachiopod Liothyrella uva (Broderip, 1833). Journal of Experimental Marine Biology and Ecology, 462, 2935.Google Scholar
Cross, E.L., Peck, L.S., Lamare, M.D., Harper, E.M. (2016). No ocean acidification effects on shell growth and repair in the New Zealand brachiopod Calloria inconspicua (Sowerby, 1846). ICES Journal of Marine Science, 73, 920926.Google Scholar
Cross, E.L., Harper, E.M., Peck, L.S. (2017). A 120-year record of resilience to environmental change in brachiopods. Global Change Biology, 24, 22622271.Google Scholar
Daufresne, M., Lengfellnera, K., Sommera, U. (2009). Global warming benefits the small in aquatic ecosystems. Proceedings of the National Academy of Sciences of the USA, 106, 1278812793.Google Scholar
Davison, W., Franklin, C.E. (2002). The Antarctic nemertean Parborlasia corrugatus: an example of an extreme oxyconformer. Polar Biology, 25, 238240.Google Scholar
Dayton, P.K., Oliver, J.S. (1977). Antarctic soft-bottom benthos in oligotrophic and eutrophic environments. Science, 197, 5558.Google Scholar
Dayton, P.K., Robilliard, G.A. (1971). Implications of pollution to the McMurdo Sound benthos. Antarctic Journal of the United States, 6, 5356.Google Scholar
De Broyer, C. (1977). Analysis of the gigantism and dwarfness of Antarctic and Sub-Antarctic gammaridean Amphipoda. In:Llano, G. A. (ed.), Adaptations within Antarctic ecosystems. Proceedings of the Third SCAR Symposium on Antarctic Biology. Smithsonian Institution, Houston, pp. 327334.Google Scholar
De Broyer, C., Koubbi, P., Griffiths, H.J., et al. (eds.) (2014). Biogeographic Atlas of the Southern Ocean. Scientific Committee on Antarctic Research, Cambridge.Google Scholar
Denny, M., Dorgan, K.M., Evangelista, D., et al. (2011). Anchor ice and benthic disturbance in shallow Antarctic waters: interspecific variation in initiation and propagation of ice crystals. Biological Bulletin, 221(2), 155163.Google Scholar
Deutsch, C. A., Tewksbury, J.J., Huey, R.B., et al. (2008). Impacts of climate warming on terrestrial ectotherms across latitude. Proceedings of the National Academy of Sciences of the USA, 105, 66686672.Google Scholar
DeVries, A.L., Wohlschlag, D.E. (1969). Freezing resistance in some Antarctic fishes. Science, 163, 10731075.CrossRefGoogle ScholarPubMed
di Prisco, G., Verde, C. (2015). The Ross Sea and its rich life: research on molecular adaptive evolution of stenothermal and eurythermal Antarctic organisms and the Italian contribution. Hydrobiologia, 761, 335361.Google Scholar
Dömel, J.S., Convey, P., Leese, F. (2015). Genetic data support independent glacial refugia and open ocean barriers to dispersal for the Southern Ocean sea spider Austropallene Cornigera. Journal of Crustacean Biology, 35, 480490.Google Scholar
Doney, S.C., Fabry, V.J., Feely, R.A., Kleypas, J.A. (2009). Ocean acidification: the other CO2 problem. Annual Review of Marine Science, 1, 169192.CrossRefGoogle ScholarPubMed
Doyle, S.R., Momo, F.R., Brethes, J.C., Ferrera, G.A. (2012). Metabolic rate and food availability of the Antarctic amphipod Gondogeneia antarctica (Chevreux 1906): seasonal variation in allometric scaling and temperature dependence. Polar Biology, 25, 413424.Google Scholar
Duffy, S., Shackelton, L.A., Holmes, E.C. (2008). Rates of evolutionary change in viruses: patterns and determinants. Nature Reviews Genetics, 9, 267276.Google Scholar
Dupont, S., Havenhand, J., Thorndyke, W., Peck, L.S., Thorndyke, M. (2008). CO2-driven ocean acidification radically affects larval survival and development in the brittlestar Ophiothrix fragilis. Marine Ecology Progress Series, 373, 285294.Google Scholar
Eastman, J.T., DeVries, A.L. (1986). Antarctic fishes. Scientific American, 254, 106114.Google Scholar
Enzor, L.A., Hunter, E.M., Place, S.P. (2017). The effects of elevated temperature and ocean acidification on the metabolic pathways of notothenioid fish. Conservation Physiology, 5, cox019.Google Scholar
Fabry, V.J., McClintock, J.B., Mathis, J.T., Grebmeier, J.M. (2009). Ocean acidification at high latitudes: the bellweather. Oceanography, 22, 160171.Google Scholar
Faulkner, K., Clusella-Trullas, S., Peck, L.S., Chown, S. (2014). Lack of coherence in the warming responses of marine crustaceans. Functional Ecology, 2, 895903.Google Scholar
Feder, M.E., Hofmann, G.E. (1999). Heat shock proteins, molecular chaperones and their stress response: evolutionary and ecological physiology. Annual Review of Physiology, 61, 243282.Google Scholar
Feely, R.A., Doney, S.C., Cooley, S.R. (2009). Ocean acidification: present conditions and future changes in a high-CO2 world. Oceanography, 22, 3647.Google Scholar
Feely, R.A., Sabine, C.L., Byrne, R.H., et al. (2012). Decadal changes in the aragonite and calcite saturation state of the Pacific Ocean. Global Biogeochemical Cycles, 26, 115.Google Scholar
Fraser, L.H., Greenall, A., Carlyle, C., Turkington, R., Friedman, C.R. (2009). Adaptive phenotypic plasticity of Pseudoroegneria spicata: response of stomatal density, leaf area and biomass to changes in water supply and increased temperature. Annals of Botany, 103(5), 769775.Google Scholar
Gatti, S. (2002). The role of sponges in high-antarctic carbon and silicon cycling: a modelling approach. Berichte zur Polarforschung/Reports on Polar Research, 434, 1124.Google Scholar
González, K., Gaitán-Espitia, J., Font, A., Cárdenas, C.A., González-Aravena, M. (2016). Expression pattern of heat shock proteins during acute thermal stress in the Antarctic sea urchin, Sterechinus neumayeri. Revista Chilena de Historia Natural, 89, 2.Google Scholar
Grange, L.J., Smith, C.R. (2013). Megafaunal communities in rapidly warming fjords along the West Antarctic Peninsula: hotspots of abundance and beta diversity. PLoS ONE, 8, e77917.Google Scholar
Griffiths, H.J., Danis, B., Clarke, A. (2011). Quantifying Antarctic marine biodiversity: the SCAR-MarBIN data portal. Deep Sea Research. II: Topical Studies in Oceanography, 58, 1829.Google Scholar
Gross, M. (2004). Emergency services: a bird’s eye perspective on the many different functions of stress proteins. Current Protein & Peptide Science, 5, 213223.Google Scholar
Gutt, J. (2001). On the direct impact of ice on marine benthic communities, a review. Polar Biology, 24, 553564.Google Scholar
Gutt, J., Bertler, N., Bracegirdle, T.J., et al. (2015). The Southern Ocean ecosystem under multiple climate stresses: an integrated circumpolar assessment. Global Change Biology, 21, 14341453.Google Scholar
Hardewig, I., Peck, L.S., Pörtner, H.O. (1999). Thermal sensitivity of mitochondrial function in the Antarctic Notothenioid Lepidonotothen nudifrons. Comparative Biochemistry and Physiology, 124A, 179189.Google Scholar
Harper, E.M., Peck, L.S. (2003). Predatory behaviour and metabolic costs in the Antarctic muricid gastropod Trophon longstaffi. Polar Biology, 26, 208217.Google Scholar
Harper, E.M., Peck, L.S. (2016). Latitudinal and depth gradients in predation pressure. Global Ecology and Biogeography, 25, 670678.Google Scholar
Harris, P.T., MacMillan-Lawler, M., Rupp, J., Baker, E.K. (2014). Geomorphology of the oceans. Marine Geology, 352, 424.Google Scholar
Hartman, O. (1964). Polychaeta errantia of Antarctica. Antarctic Research Series 3. American Geophysical Union, Washington, DC.CrossRefGoogle Scholar
Ho, C.-K., Pennings, S.C., Carefoot, T.H. (2010). Is diet quality an overlooked mechanism for Bergmannʼs rule? American Naturalist, 175, 269276.Google Scholar
Hoegh-Guldberg, O. (1999). Climate change, coral bleaching and the future of the world’s coral reefs. Marine and Freshwater Research, 50, 839866.Google Scholar
Hofmann, G.E., Buckley, B.A., Airaksinen, S., Keen, J.E., Somero, G.N. (2000). Heat-shock protein expression is absent in the Antarctic fish Trematomus bernacchii (Family Nototheniidae). Journal of Experimental Biology, 203, 23312339.Google Scholar
Horner, R.A. (2018). Sea Ice Biota. CRC Press, Boca Raton, FL.Google Scholar
Huth, T.P., Place, S.P. (2016). Transcriptome wide analyses reveal a sustained cellular stress response in the gill tissue of Trematomus bernacchii after acclimation to multiple stressors. BMC Genomics, 17, 127.Google Scholar
IPCC; Field, C. B., Barros, V. R., Dokken, D. J., et al. (eds.) (2014). Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK and New York.Google Scholar
Jakobsson, M., Nilsson, J., Anderson, L., et al. (2016). Evidence for an ice shelf covering the central Arctic Ocean during the penultimate glaciations. Nature Communications, 10.1038/ncomms10365.Google Scholar
Janecki, T., Kidawa, A., Potocka, M. (2010). The effects of temperature and salinity on vital biological functions of the Antarctic crustacean Serolis polita. Polar Biology, 33, 10131020.Google Scholar
Kleypas, J.A., Feely, R.A., Fabry, V.J., et al. (2006). Impact of ocean acidification on coral reefs and other marine calcifiers: a guide for future research. Report of a workshop held 18–20 April 2005, St. Petersburg, FL, sponsored by NSF, NOAA, and the US Geological Survey.Google Scholar
Kroeker, K.J., Kordas, R.L., Crim, R., et al. (2013). Impacts of ocean acidification on marine organisms: quantifying sensitivities and interaction with warming. Global Change Biology, 19, 18841896.Google Scholar
Lohbeck, K.T., Riebesell, U., Collins, S., Reusch, T.B.H. (2013). Functional genetic divergence in high CO2 adapted Emiliania huxleyi populations. Evolution, 67, 18921900.Google Scholar
McNeil, B.I., Matear, R.J. (2008). Southern Ocean acidification: a tipping point at 450-ppm atmospheric CO2. Proceedings of the National Academy of Sciences of the USA, 105, 1886018864.Google Scholar
Montes-Hugo, M., Doney, S.C., Ducklow, H.W., et al. (2009). Recent changes in phytoplankton communities associated with rapid regional climate change along the Western Antarctic Peninsula. Science, 323, 14701473.Google Scholar
Montgomery, J., Clements, K. (2000). Disaptation and recovery in the evolution of Antarctic fishes. Trends in Ecology and Evolution, 15, 267271.Google Scholar
Moran, A.L., Woods, H.A. (2012). Why might they be giants? Towards an understanding of polar gigantism. Journal of Experimental Biology, 215, 19952002.Google Scholar
Morley, S.A., Hirse, T., Thorne, M.A.S., Pörtner, H.O., Peck, L.S. (2012). Physiological plasticity, long term resistance or acclimation to temperature, in the Antarctic bivalve, Laternula elliptica.Comparative Biochemistry and Physiology, 162A, 1621.Google Scholar
Morley, S.A., Suckling, C.S., Clark, M.S., Cross, E.L., Peck, L.S. (2016). Long-term effects of altered pH and temperature on the feeding energetics of the Antarctic sea urchin, Sterechinus neumayeri. Biodiversity, 17, 3445.Google Scholar
Nghiem, S.V., Rigor, I.G., Clemente-Colón, P., Neumann, G., Li, P.P. (2016). Geophysical constraints on the Antarctic sea ice cover. Remote Sensing of Environment, 181, 281292.Google Scholar
Orr, J.C., Fabry, V.J., Aumont, O., et al. (2005). Anthropogenic ocean acidification over the twenty-first century and its impact on calcifying organisms. Nature, 437, 681686.Google Scholar
Peck, L.S. (1989). Temperature and basal metabolism in two Antarctic marine herbivores. Journal of Experimental Biology, 127, 112.Google Scholar
Peck, L.S. (2002). Ecophysiology of Antarctic marine ectotherms: limits to life. Polar Biology, 25, 3140.Google Scholar
Peck, L.S. (2008). Brachiopods and climate change. Earth and Environmental Science Transactions of The Royal Society of Edinburgh, 98, 451456.Google Scholar
Peck, L.S. (2011). Organisms and responses to environmental change. Marine Genomics, 4, 237243.Google Scholar
Peck, L.S. (2015). DeVries: the Art of not freezing fish. Classics series. Journal of Experimental Biology, 218, 21462147.Google Scholar
Peck, L.S. (2016). A cold limit to adaptation in the sea. Trends in Ecology and Evolution, 31, 1326.CrossRefGoogle ScholarPubMed
Peck, L.S. (2018). Antarctic marine biodiversity: adaptations, environments and responses to change. Oceanography and Marine Biology: An Annual Review, 56, 105236.Google Scholar
Peck, L.S., Brockington, S., Brey, T. (1997). Growth and metabolism in the antarctic brachiopod Liothyrella uva. Philosophical Transactions: Biological Sciences, 352(1355), 851858.Google Scholar
Peck, L.S., Chapelle, G. (1999). Amphipod gigantism dictated by oxygen availability? Ecology Letters, 2, 401403.Google Scholar
Peck, L.S., Chapelle, G. (2003). Reduced oxygen at high altitude limits maximum size. Proceedings of the Royal Society of London Series B: Biological Sciences, 270, S166S167.Google Scholar
Peck, L.S., Conway, L.Z. (2000). The myth of metabolic cold adaptation: oxygen consumption in stenothermal Antarctic bivalves. In:Harper, E. M.,Taylor, J. D., Crame, J. A. (eds) The Evolutionary Biology of the Bivalvia. Geological Society, London, Special Publications, Vol. 177.Geological Society, London, pp. 441445.Google Scholar
Peck, L.S., Brockington, S., VanHove, S., Beghyn, M. (1999). Community recovery following catastrophic iceberg impacts in Antarctica. Marine Ecology Progress Series, 186, 18.Google Scholar
Peck, L.S., Clark, M.S., Morley, S.A., Massey, A., Rossetti, H. (2009b). Animal temperature limits and ecological relevance: effects of size, activity and rates of change. Functional Ecology, 23, 248253.Google Scholar
Peck, L. S., Convey, P., Barnes, D. K. A. (2006). Environmental constraints on life histories in Antarctic ecosystems: tempos, timings and predictability. Biological Reviews, 81, 75109.Google Scholar
Peck, L.S., Massey, A., Thorne, M.A.S., Clark, M.S. (2009a). Lack of acclimation in Ophionotus victoriae: brittle stars are not fish. Polar Biology, 32, 399402.CrossRefGoogle Scholar
Peck, L.S., Morley, S.A., Clark, M.S. (2010). Poor acclimation capacities in Antarctic marine ectotherms. Marine Biology, 157, 20512059.Google Scholar
Peck, L.S., Morley, S.A., Richard, J., Clark, M.S. (2014). Acclimation and thermal tolerance in Antarctic marine ectotherms. Journal of Experimental Biology, 217, 1622.Google Scholar
Peck, L.S., Pörtner, H.O., Hardewig, I. (2002).Metabolic demand, oxygen supply and critical temperatures in the Antarctic bivalve Laternula elliptica. Physiological Biochemical Zoology, 75, 123133.Google Scholar
Peck, L.S., Souster, T., Clark, M.S. (2013). Juveniles are more resistant to warming than adults in 4 species of Antarctic marine invertebrates. PLoS ONE, 8, e66033.Google Scholar
Peck, L.S., Webb, K.E., Bailey, D. (2004). Extreme sensitivity of biological function to temperature in Antarctic marine species. Functional Ecology, 18, 625630.CrossRefGoogle Scholar
Peck, L.S., Webb, K.E., Clark, M.S., Miller, A., Hill, T. (2008). Temperature limits to activity, feeding and metabolism in the Antarctic starfish Odontaster validus. Marine Ecology Progress Series, 381, 181189.Google Scholar
Peck, V.L., Oakes, R.L., Harper, E.M., Manno, C., Tarling, G.A. (2018). Pteropods counter mechanical damage and dissolution through extensive shell repair. Nature Communications, 9, 264.Google Scholar
Peel, M.C., Wyndham, R.C. (1999). Selection of clc, cba, and fcb chlorobenzoate-catabolic genotypes from groundwater and surface waters adjacent to the Hyde Park, Niagara Falls, chemical landfill. Applied and Environmental Microbiology, 65, 16271635.Google Scholar
Pespeni, M.H., Sanford, E., Gaylord, B., et al. (2013). Evolutionary change during experimental ocean acidification. PNAS, 110(17), 6937–6942.Google Scholar
Piepenburg, D., Archambault, P., Ambrose, W, et al. (2010). Towards a pan-Arctic inventory of the species diversity of the macro- and megabenthic fauna of the Arctic shelf seas. Marine Biodiversity, 41, 5170.CrossRefGoogle Scholar
Place, S.P., Hofmann, G.E. (2005). Constitutive expression of a stress inducible heat shock protein gene, hsp70, in phylogenetically distant Antarctic fish. Polar Biology, 28, 261267.CrossRefGoogle Scholar
Podrabsky, J.E., Somero, G.N. (2006). Inducible heat tolerance in Antarctic notothenioid fishes. Polar Biology, 30, 3943.Google Scholar
Pörtner, H.O., Peck, L.S., Hirse, T. (2006). Hyperoxia alleviates thermal stress in the Antarctic bivalve, Laternula elliptica: evidence for oxygen limited thermal tolerance? Polar Biology, 29, 688693.Google Scholar
Pörtner, H.O., Peck, L.S. Zielinski, S., Conway, L.Z. (1999). Temperature and metabolism in the highly stenothermal bivalve mollusc Limopsis marionensis from the Weddell Sea, Antarctica. Polar Biology, 22, 1730.Google Scholar
Richard, J., Morley, S.A., Peck, L.S. (2012). Estimating long-term survival temperatures at the assemblage level in the marine environment: towards macrophysiology. PLoS ONE, 7, e34655.Google Scholar
Robertson, R.F., El-Haj, A.J., Clarke, A., Peck, L.S., Taylor, E.W. (2001). The effects of temperature on metabolic rate and protein synthesis following a meal in the isopod Glyptonotus antarcticus eights (1852). Polar Biology, 24, 677686.Google Scholar
Robinson, E., Davison, W. (2008). The Antarctic notothenioid fish Pagothenia borchgrevinki is thermally flexible: acclimation changes oxygen consumption. Polar Biology, 31, 317326.Google Scholar
Rodríguez-Romero, A., Jarrold, M.D., Massamba-N’Siala, G., Spicer, J.I., Calosi, P. (2016). Multi-generational responses of a marine polychaete to a rapid change in seawater pCO2. Evolutionary Applications, 9, 10821095.Google Scholar
Roggatz, C.C., Lorch, M., Hardege, J.D., Benoit, D.M. (2016). Ocean acidification affects marine chemical communication by changing structure and function of peptide signalling molecules. Global Change Biology, 22, 39143926.Google Scholar
Reed, A.J., Thatje, S. (2015). Long-term acclimation and potential scope for thermal resilience in Southern Ocean bivalves. Marine Biology, 162, 22172224.Google Scholar
Rohling, E.J., Foster, G.L., Grant, K.M., et al. (2014). Sea-level and deep-sea-temperature variability over the past 5.3 million years. Nature, 508, 477482.Google Scholar
Ruud, J.T. (1954). Vertebrates without erythrocytes and blood pigment. Nature, 173, 848850.Google Scholar
Schloss, I.R., Abele, D., Moreau, S., et al. (2012). Response of phytoplankton dynamics to 19-year (1991–2009) climate trends in Potter Cove (Antarctica). Journal of Marine Systems, 92, 5366.Google Scholar
Schofield, O., Ducklow, H.W., Martinson, D.G., et al. (2010). How do polar marine ecosystems respond to rapid climate change? Science, 328, 15201523.Google Scholar
Schram, J.B., McClintock, J.B., Amsler, C.D., Baker, B.J. (2015). Impacts of acute elevated seawater temperature on the feeding preferences of an Antarctic amphipod toward chemically deterrent macroalgae. Marine Biology, 162, 425433.Google Scholar
Segawa, T., Ushida, K., Narita, H., Kanda, H., Kohshima, S. (2010). Bacterial communities in two Antarctic ice cores analyzed by 16S rRNA gene sequencing analysis. Polar Science, 4, 215227.Google Scholar
Shin, S.C., Kim, S.J., Lee, J.K., et al. (2012). Transcriptomics and comparative analysis of three Antarctic notothenioid fishes. PLoS ONE, 16, e43762.Google Scholar
Somero, G.N. (2010). The physiology of climate change: how potentials for acclimatization and genetic adaptation will determine ‘winners’ and ‘losers’. Journal of Experimental Biology, 213, 912920.Google Scholar
Somero, G.N. (2015). Temporal patterning of thermal acclimation: from behaviour to membrane biophysics. Journal of Experimental Biology, 218, 167169.Google Scholar
Somero, G.N., De Vries, A.L. (1967). Temperature tolerance of some Antarctic fishes. Science, 156, 257–258.CrossRefGoogle Scholar
Stanwell-Smith, D.P., Peck, L.S. (1998). Temperature and embryonic development in relation to spawning and field occurrence of larvae of 3 Antarctic echinoderms. Biological Bulletin, 194, 4452.Google Scholar
Stillman, J.H., Paganini, A.W. (2015). Biochemical adaptation to ocean acidification. Journal of Experimental Biology, 218, 19461955.Google Scholar
Stroeve, J., Notz, D. (2018). Changing state of Arctic sea ice across all seasons. Environmental Research Letters, 13,103001.Google Scholar
Suckling, C.C., Clark, M.S., Richard, J., et al. (2015). Adult acclimation to combined temperature and pH stressors significantly enhances reproductive outcomes compared to short-term exposures. Journal of Animal Ecology, 84, 773–784.Google Scholar
Sunday, J.M., Calosi, P., Dupont, S., et al. (2014). Evolution in an acidifying ocean. Trends in Ecology and Evolution, 29, 117125.CrossRefGoogle Scholar
Sweetman, A.K., Thurber, A.R., Smith, C.R., et al. (2017). Major impacts of climate change on deep-sea benthic ecosystems. Elementa: Science of the Anthropocene, 5, 4.Google Scholar
Thorson, G. (1957). Bottom communities. In: Hedgpeth, J. W. (ed.) Treatise on Marine Ecology and Paleoecology. Geological Society of America, pp. 461534.Google Scholar
Todgham, A.E., Hoaglund, E.A., Hofmann, G.E. (2007). Is cold the new hot? Elevated ubiquitin-conjugated protein levels in tissues of Antarctic fish as evidence for cold-denaturation of proteins in vivo.Journal of Comparative Physiology B, 177, 857866.Google Scholar
Todgham, A.E., Crombie, T.A., Hofmann, G.E. (2017). The effect of temperature adaptation on the ubiquitin-proteasome pathway in notothenioid fishes. Journal of Experimental Biology, 220, 369378.Google Scholar
Tomanek, L. (2010). Variation in the heat shock response and its implication for predicting the effect of global climate change on species’ biogeographical distribution ranges and metabolic costs. Journal of Experimental Biology, 213, 971979.Google Scholar
US National Snow and Ice Data Centre (2018). Arctic sea ice extent arrives at its minimum. http://nsidc.org/arcticseaicenews/2018/09/Google Scholar
Verde, C., Giordano, D., di Prisco, G., Andersen, Ø. (2012). The hemoglobins of polar fish: evolutionary and physiological significance of multiplicity in Arctic fish. Biodiversity, 13, 228233.Google Scholar
Vermeij, G.J. (2016). Gigantism and its implications for the history of life. PLoS ONE, 11, e0146092.Google Scholar
Watson, J.D., Baker, T.A., Bell, S.P., et al. (2014). Molecular Biology of the Gene. Cold Spring Harbour Laboratory Press, New York.Google Scholar
Watson, S.-Southgate, A, Tyler, P, Peck, P.A., L.S. (2009). Early larval development of the Sydney rock oyster Saccostrea glomerata under near-future predictions of CO2-driven ocean acidification. Journal of Shellfish Research, 28, 431437.Google Scholar
Watson, S.-Peck, A, Tyler, L.S., P.A., et al. (2012). Marine invertebrate skeleton size varies with latitude, temperature, and carbonate saturation: implications for global change and ocean acidification. Global Change Biology, 18, 30263038.Google Scholar
Watson, S.-A., Peck, L.S., Morley, S.A., Munday, P.L. (2017). Latitudinal trends in shell production cost from the tropics to the poles. Scientific Reports, 3, e1701362.Google Scholar
Węsławski, J.M., Kendall, M.A, Włodarska-Kowalczuk, M., et al. (2011). Climate change effects on Arctic fjord and coastal macrobenthic diversity – observations and predictions. Marine Biodiversity, 41, 7185.Google Scholar
Whiteley, N.M., Taylor, E.W., El Haj, A.J. (1997). Seasonal and latitudinal adaptation to temperature in crustaceans. Journal of Thermal Biology, 22, 419427.Google Scholar
Wiedenmann, J., Cresswell, K.A., Mangel, M. (2009). Connecting recruitment of Antarctic krill and sea-ice. Limnology and Oceanography, 54, 799811.Google Scholar
Wiest, L.A., Buynevich, I.G., Grandstaff, D.E., et al. (2015). Trace fossil evidence suggests widespread dwarfism in response to the end-Cretaceous mass extinction: Braggs, Alabama and Brazos River, Texas. Palaeogeography, Palaeoclimatology, Palaeoecology, 417, 405411.Google Scholar
Young, J.S., Peck, L.S., Matheson, T. (2006a). The effects of temperature on walking in temperate and Antarctic crustaceans. Polar Biology, 29, 978987.Google Scholar
Young, J.S., Peck, L.S., Matheson, T. (2006b). The effects of temperature on peripheral neuronal function in eurythermal and stenothermal crustaceans. Journal of Experimental Biology, 209, 1976–1987.Google Scholar
Zachos, J., Pagani, M., Sloan, L., Thomas, E., Billups, K. (2001). Trends, rhythms, and aberrations in global climate 65 Ma to present. Science, 292, 686693.Google Scholar
Zachos, J.C., Dickens, G.R., Zeebe, R.E. (2008). An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics. Nature, 451, 279283.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

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

Available formats
×

Save book to Google Drive

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

Available formats
×