Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-cfpbc Total loading time: 0 Render date: 2024-04-23T14:29:48.598Z Has data issue: false hasContentIssue false

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
Type
Chapter
Information
Life in Extreme Environments
Insights in Biological Capability
, pp. 149 - 296
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

References

Ackerly, D.D., Loarie, S.R., Cornwell, W.K., et al. (2010). The geography of climate change: implications for conservation biogeography. Diversity and Distribution, 16, 476487; https://doi.org/10.1111/j.1472-4642.2010.00654.xGoogle Scholar
Aitken, S.N., Bemmels, J.B. (2015). Time to get moving: assisted gene flow of forest trees. Evolutionary Applications, 9, 271290.CrossRefGoogle ScholarPubMed
Aitken, S.N., Yeaman, S., Holliday, J.A., Wang, T., Curtis‐McLane, S. (2008). Adaptation, migration or extirpation: climate change outcomes for tree populations. Evolutionary Applications, 1(1), 95111; https://doi.org/10.1111/j.1752–4571.2007.00013.CrossRefGoogle Scholar
Albertson, R.C., Cresko, W., Detrich, H.W., Postlethwait, J.H. (2009). Evolutionary mutant models for human disease. Trends in Genetics, 25, 7481.CrossRefGoogle ScholarPubMed
Altman, A., Hasegawa, P.M. (2012). Introduction to plant biotechnology 2011: basic aspects and agricultural implications. In: Altman, A, Hasegawa, P.M. (eds) Plant Biotechnology and Agriculture: Prospects for the 21st Century. Elsevier and Academic Press, San Diego, pp. 1586.Google Scholar
Andrady, A., Aucamp, P.J., Austin, A.T., Bais, A.F., Ballaré, C.L. (2016). Environmental effects of ozone depletion and its interactions with climate change: progress report, 2015. Photochemical & Photobiological Sciences, 15, 141174.Google Scholar
Anonymous (2009). Convention on Biological Diversity. Website. www.cbd.int/Google Scholar
Antony, C.P., Cockell, C.S., Shouche, Y.S. (2012). Life in (and on) the rocks. Journal of Biosciences, 37, 311.CrossRefGoogle ScholarPubMed
Augé, R.M. (2001). Water relations, drought and vesicular-arbuscular mycorrhizal symbiosis. Mychorriza, 11, 342.Google Scholar
Battaglia, M., Covarrubias, A.A. (2013). Late embryogenesis abundant (LEA) proteins in legumes. Frontiers in Plant Science, 4; https://doi.org/10.3389/fpls.2013.00190. eCollection 2013.Google Scholar
Beckett, M., Loreto, F., Velikova, V., et al. (2012). Photosynthetic limitations and volatile and non-volatile isoprenoids in the poikilochlorophyllous resurrection plant Xerophyta humilis during dehydration and rehydration. Plant Cell Environment, 35, 20612074.CrossRefGoogle ScholarPubMed
Bell, G., Collins, S. (2008). Adaptation, extinction and global change. Evolutionary Applications, 1(1), 316; https://doi.org/10.1111/j.1752-4571.2007.00011.x.Google Scholar
Bellard, C., Bertelsmeier, C., Leadley, P., Thuiller, W., Courchamp, F. (2012). Impacts of climate change on the future of biodiversity. Ecology Letters, 15, 365377; https://doi.org/10.1111/j.1461-0248.2011.01736.xGoogle Scholar
Bjorkman, A.D., Vellend, M., Frei, E.R., Henry, G.H.R. (2016). Climate adaptation is not enough: warming does not facilitate success of southern tundra plant populations in the high Arctic. Global Change Biology, 23, 15401551; https://doi.org/10.1111/gcb.13417.CrossRefGoogle Scholar
Blum, A. (2017). Osmotic adjustment is a prime drought stress adaptive engine in support of plant production. Plant Cell Environment, 40(1), 4–10; https://doi.org/10.1111/pce.12800.Google Scholar
Bossdorf, O., Richards, C.L., Pigliucci, M. (2008). Epigenetics for ecologists. Ecology Letters, 11, 106115.Google Scholar
Bressan, R.A., Reddy, M.P., Chung, S.O., et al. (2012). Stress-adapted extremophiles provide energy without interference with food production. Food Security, 3, 93105.CrossRefGoogle Scholar
Bush, M.B., Silman, M.R., Urrego, D.H. (2004). 48,000 years of climate and forest change in a biodiversity hot spot. Science, 303, 827829.Google Scholar
Callaghan, T.V., Bjorn, L.O., Chernov, Y., et al. (2004). Biodiversity, distributions and adaptations of arctic species in the context of environmental change. Ambio, 33, 404417.Google Scholar
CAREX (2011). CAREX Roadmap for Research on Life in Extreme Environments. CAREX Publication no. 9, pp. 140. Available at: www.carex-eu.org/Google Scholar
Carlson, C.J., Burgio, K.R., Dougherty, E.R., et al. (2017). Parasite biodiversity faces extinction and redistribution in a changing climate. Science Advances, 3(9), e1602422.Google Scholar
Cavicchioli, R., Siddiqui, K.S., Andrews, D., Sowers, K.R. (2002). Low-temperature extremophiles and their applications. Current Opinion in Biotechnology, 13, 253261.CrossRefGoogle ScholarPubMed
Chapelle, G., Peck, L.S. (1999). Polar gigantism dictated by oxygen availability. Nature, 399, 114115.CrossRefGoogle Scholar
Chaves, M.M., Maroco, J.P., Pereira, J.S. (2003). Understanding plant responses to drought – from genes to the whole plant. Functional Plant Biology, 30, 239264.Google Scholar
Chen, I.-C., Hill, J.K., Ohlemüller, R., Roy, D.B., Thomas, C.D. (2011). Rapid range shifts of species associated with high levels of climate warming. Science, 333, 10241026.CrossRefGoogle ScholarPubMed
Cheung, W.W.L., Lam, V.W.Y., Sarmiento, J.L., et al. (2009). Projecting global marine biodiversity impacts under climate change scenarios. Fish and Fisheries, 10, 235251.Google Scholar
Chevin, L.-M., Gallet, R., Gomulkiewicz, R., Holt, R.D., Fellous, S. (2013). Phenotypic plasticity in evolutionary rescue experiments. Phylosophical Transactions of the Royal Society B, 368, 20120089; https://doi.org/10.1098/rstb.2012.0089.Google Scholar
Chinnusamy, W., Zhu, J.-K. (2009). Epigenetic regulation of stress responses in plants. Current Opinions in Plant Biology, 12, 133139.CrossRefGoogle ScholarPubMed
Chown, S.L., Brooks, C.M., Terauds, A., et al. (2017). Antarctica and the strategic plan for biodiversity. PLoS Biology, 15(3), e2001656. https://doi.org/10.1371/journal.pbio.2001656.Google Scholar
Christmas, M.J., Breed, M.F., Lowe, A.J. (2016). Constraints to and conservation implications for climate change adaptation in plants. Conservation Genetics, 17, 305320.CrossRefGoogle 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., Husmann, G., Thorne, M.A.S., et al. (2013). Hypoxia impacts large adults first: consequences in a warming world. Global Change Biology, 19, 22512263.CrossRefGoogle Scholar
Clark, M.S., Thorne, M.A.S., King, M., et al. (2018). Life in the intertidal: cellular responses, methylation and epigenetics. Functional Ecology, 32, 19821994.CrossRefGoogle Scholar
Clarke, A., Griffith, H.J., Linse, K., Crame, J.A. (2007). How well do we know the Antarctic marine fauna? A preliminary study of macroecological and biogeographic patterns in Southern Ocean gastropod and bivalve molluscs. Diversity and Distributions, 13(5), 620632; https://doi.org/10.1111/j.1472-4642.2007.00380.xGoogle Scholar
Cocca, E., Ratnayake-Lecamwasam, M., Parker, S.K., et al. (1995). Genomic remnants of α-globin genes in the hemoglobinless Antarctic icefishes. Proceedings of the National Academy of Sciences of the USA, 92, 18171821.Google Scholar
Coker, J.A. (2016). Extremophiles and biotechnology: current uses and prospects. F1000 Research, Faculty Reviews-396; https://doi.org/10.12688/f1000research.7432.1. eCollection 2016.Google Scholar
Colwell, R.K., Brehm, G.,Cardelús, C.L., Gilman, A.C., Longino, J.C. (2008). Global warming, elevational range shifts, and lowland biotic attrition in the wet tropics. Science, 322, 258261.CrossRefGoogle ScholarPubMed
Considine, M.J., Considine, J.A. (2016). On the language and physiology of dormancy and quiescence in plants. Journal of Experimental Botany, 67, 31893203; https://doi.org/10.1093/jxb/erw138.Google Scholar
Dalmaso, G.Z.L., Ferreira, D., Vermelho, A.B. (2015). Marine extremophiles: A source of hydrolases for biotechnological applications. Marine Drugs, 13, 1925 –1965.Google Scholar
Danovaro, R., Corinaldesi, C., Dell’Anno, A., Rastelli, E. (2017). Potential impact of global climate change on benthic deep-sea microbes. FEMS Microbiology Letters, 364(23); https://doi.org/10.1093/femsle/fnx214..Google Scholar
Das, K., Roychoudhury, A. (2014). Reactive oxygen species (ROS) and response of antioxidants as ROS-scavengers during environmental stress in plants. Frontiers in Environmental Science; https://doi.org 10.3389/fenvs.2014.00053.Google Scholar
De Micco, V., Aronne, G. (2012). Morpho-anatomical traits for plant adaptation to drought. In: R. Aroca (ed.) Plant Responses to Drought Stress. Springer, Berlin/Heidelberg, pp. 3762; https://doi.org/10.1007/978-3-642-32653-0_1.CrossRefGoogle Scholar
de Pascale, D., De Santi, C., Fu, J., Landfald, B. (2012). The microbial diversity of Polar environments is a fertile ground for bioprospecting. Marine Genomics, 8, 1522.CrossRefGoogle ScholarPubMed
De Vries, A.L., Cheng, C.-H.C. (2005). Antifreeze proteins and organismal freezing avoidance in polar fishes. In: Farrell, A.P., Steffensen, J.F. (eds) The Physiology of Polar Fishes, Vol. 22. Elsevier Academic Press, San Diego, pp. 155201.Google Scholar
Dhamankar, H., Prather, K.L.J. (2011). Microbial chemical factories: recent advances in pathway engineering for synthesis of value added chemicals. Current Opinion in Structural Biology, 21, 488494.Google Scholar
Di Fraia, R., Wilquet, V., Ciardiello, M.A., et al. (2000). NADP+-dependent glutamate dehydrogenase in the Antarctic psychrotolerant bacterium Psychrobacter sp. TAD1. European Journal of Biochemistry, 267, 121131.Google Scholar
di Prisco, G., Cocca, E., Parker, S., Detrich, H. (2002). Tracking the evolutionary loss of hemoglobin expression by the white-blooded Antarctic icefishes. Gene, 295, 185191.Google Scholar
di Prisco, G., Convey, P., Gutt, J., Cowan, D., Conlan, K., Verde, C. (2012a). Understanding and protecting the world’s biodiversity: The role and legacy of the SCAR programme ‘Evolution and Biodiversity in the Antarctic’. Marine Genomics, 8, 38.Google Scholar
di Prisco, G., Giordano, D., Russo, R., Verde, C. (2012b). The challenges of low temperature in the evolution of bacteria. In: di Prisco, G, Verde, C (eds) Pole to Pole, Adaptation and Evolution in Marine Environments, Vol. 1. A book series on the scientific achievements of environmental research during the International Polar Year (IPY). Springer, pp. 183195.Google Scholar
Dirnböck, T., Essl, F., Rabitsch, W. (2011). Disproportional risk for habitat loss of high-altitude endemic species under climate change. Global Change Biology, 17, 990996.Google Scholar
Dobrowski, S.Z., Abatzoglou, J., Swanson, A.K., et al. (2013). The climate velocity of the contiguous United States during the 20th century. Global Change Biology, 19, 241251; https://doi.org/10.1111/gcb.12026Google Scholar
Duman, J.G., Olsen, T.M. (1993). Thermal hysteresis protein activity in bacteria, fungi, and phylogenetically diverse plants. Cryobiology, 30, 322328.Google Scholar
Dunton, K. (1992). Arctic biogeography – The paradox of the marine benthic fauna and flora. Trends in Ecology & Evolution, 7, 183189.CrossRefGoogle ScholarPubMed
Duplantis, B.N., Osusky, M., Schmerk, C.L., et al. (2010). Essential genes from Arctic bacteria used to construct stable, temperature-sensitive bacterial vaccines. Proceedings of the National Academy of Sciences of the USA, 107, 1345613460.Google Scholar
Easterling, D.R., Meehl, G.A., Parmesan, C., et al. (2000). Climate extremes: observations, modeling, and impacts. Science, 289, 20682074.Google Scholar
Ekblom, R., Galindo, J. (2011). Applications of next generation sequencing in molecular biology of non-model organisms. Heredity, 107, 115.CrossRefGoogle ScholarPubMed
Fox-Powell, M.G., Hallsworth, J.E., Cousins, C.R., Cockell, C.S. (2016). Ionic strength is a barrier to the habitability of Mars. Astrobiology, 16, 427442. https://doi.org/10.1089/ast.2015.1432.Google Scholar
Gomes, J., Steiner, W. (2004). The biocatalytic potential of extremophiles and extremozymes. Food Technology and Biotechnology, 42, 223235.Google Scholar
Grace, J., Berninger, F., Nagy, L. (2002). Impacts of climate change on the tree line. Annals of Botany, 90, 537544.Google Scholar
Graham, J.E., Clark, M.E., Nadler, D.C., et al. (2011). Identification and characterization of a multidomain hyperthermophilic cellulase from an archaeal enrichment. Nature Communications, 2, 375; https://doi.org/10.1038/ncomms1373.Google Scholar
Gribaldo, S., Brochier-Armanet, C. (2006). The origin and evolution of Archaea: a state of the art. Philosophical Transactions of the Royal Society of London B Biological Sciences, 361, 10071022.CrossRefGoogle ScholarPubMed
Gunderson, L.H. (2000). Ecological resilience – in theory and application. Annual Review of Ecology and Systematics, 31, 425439.Google Scholar
Gutt, J., Isla, E., Bertler, A.N., et al. (2018). Cross-disciplinarity in the advance of Antarctic ecosystem research. Marine Genomics, 37, 1–18; https://doi.org/10.1016/j.margen.2017.09.006.Google Scholar
Hamilton, J.A., Miller, J.M. (2015). Adaptive introgression as a resource for management and genetic conservation in a changing climate. Conservation Biology, 30, 3341; https://doi.org/10.1111/cobi.12574.Google Scholar
Hampe, A. (2011). Plants on the move: the role of seed dispersal and initial population establishment for climate-driven range expansion. Acta Oecologica, 37, 666673.CrossRefGoogle Scholar
Hansen, J., Sato, M., Ruedy, R., Lo, K., Lea, D.W., Medina-Elizade, M. (2006). Global temperature change. Proceedings of the National Academy of Science of the USA, 103, 1428814293.Google Scholar
Harrison, J.P., Gheeraert, N., Tsigelnitskiy, D., Cockell, C.S. (2013). The limits for life under multiple extremes. Trends in Microbiology, 21, 204212. http://dx.doi.org/10.1016/j.tim2013.01.006.CrossRefGoogle ScholarPubMed
Hasselmann, K., Latif, M., Hooss, G. et al. (2003). The challenge of long-term climate change. Science, 302, 1923–1925.Google Scholar
Hewitt, G. (2000). The genetic legacy of the Quaternary ice ages. Nature, 405, 907913.Google Scholar
Hoffman, J.I., Clarke, A., Linse, K., Peck, L.S. (2011). Effects of brooding and broadcasting reproductive modes on the population genetic structure of two Antarctic gastropod molluscs. Marine Biology, 158, 287296.CrossRefGoogle Scholar
Hooper, D.U., Adair, E.C., Cardinale, B.J., et al. (2012). A global synthesis reveals biodiversity loss as a major driver of ecosystem change. Nature, 486, 105109.Google Scholar
Hughes, I.I. (2000). Biological consequences of global warming: is the signal already apparent?Trends in Ecology and Evolution, 15(2), 5661.Google Scholar
IPCC (2014). Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the First Assessment Report of the Intergovernmental Panel on Climate Change. Edited by Core Writing Team,Pachauri, R.K.,Meyer, L.A.. IPCC, Geneva, Switzerland.Google Scholar
Johannessen, O.M., Bengtsson, L., Miles, M.W., et al. (2004). Arctic climate change: observed and modelled temperature and sea-ice variability. Tellus, 56A, 328341.Google Scholar
Johnston, I.A., Fernandez, D.A., Calvo, J., et al. (2003). Reduction in muscle fibre number during the adaptive radiation of notothenioid fishes: a phylogenetic perspective. Journal of Experimental Biology, 206, 25952609.Google Scholar
Joseph, B., Ramteke, P.W., Thomas, G. (2008). Cold active microbial lipases: Some hot issues and recent developments. Biotechnology Advances, 26, 457470.Google Scholar
Kapsenberg, L., Kelley, A.L., Shaw, E.C., et al. (2015). Near-shore Antarctic pH variability has implications for the design of ocean acidification experiments. Scientific Reports, 5, 10497; https://doi.org/10.1038/srep10497.Google Scholar
Klanderud, K., Birks, H.J.B. (2003). Recent increases in species richness and shifts in altitudinal distributions of Norwegian mountain plants. Holocene, 13, 16.Google Scholar
Kremer, A., Ronce, O., Robledo-Arnuncio, J.J., et al. (2012). Long-distance gene flow and adaptation of forest trees to rapid climate change. Ecology Letters, 15, 378392.Google Scholar
Kristensen, T.N., Ketola, T., Kronholm, I. (2018). Adaptation to environmental stress at different timescales. Annals of the New York Academy of Sciences. https://doi.org/10.1111/nyas.13974. [Epub ahead of print].Google Scholar
Lawyer, F.C., Stoffel, S., Saiki, R.K., et al. (1993). High-level expression, purification, and enzymatic characterization of full-length Thermus aquaticus DNA polymerase and a truncated form deficient in 5’ to 3’ exonuclease activity. Genome Research, 2, 275287.Google Scholar
Leary, D. (2008). Bioprospecting in the Arctic. United Nations University Institute of Advanced Study Report (UNU-IAS 2008 report), Nishi-ku, Yokohama, Japan, pp. 145.Google Scholar
Li, H., Wei, J.C. (2016). Functional analysis of thioredoxin from the desert lichen-forming fungus, Endocarpon pusillum Hedwig, reveals its role in stress tolerance. Scientific Reports, 6, Article number 27184; https://doi.org/10.1038/srep27184.Google Scholar
Li, S.-J., Hua, Z.-S., Huang, L.N., et al. (2014). Microbial communities evolve faster in extreme environments. Scientific Reports, 4, Article number 6205.Google Scholar
Liszka, M.J., Clark, M.E., Schneider, E., Clark, D.S. (2012). Nature versus nurture: developing enzymes that function under extreme conditions. Annual Review of Chemical and Biomolecular Engineering, 3, 77102.CrossRefGoogle ScholarPubMed
López-Maury, L., Marguerat, S., Bähler, J. (2008). Tuning gene expression to changing environments: from rapid responses to evolutionary adaptation. Nature Reviews Genetics, 9, 583593; https://doi.org/10.1038/nrg2398.CrossRefGoogle ScholarPubMed
Lunine, J.L. (2006). Physical conditions on the early Earth. Philosophical Transactions of the Royal Society of London B Biological Sciences, 361, 17211731.CrossRefGoogle ScholarPubMed
Maher, B. (2009). Evolution: biology’s next top model? Nature, 458, 695698.Google Scholar
Maksym, T. (2018). Arctic and antarctic sea ice change: contrasts, commonalities, and causes. Annual Review of Marine Science, 11; https://doi.org/10.1146/annurev-marine-010816-060610.Google Scholar
Margesin, R., Schinner, F. (2001). Potential of halotolerant and halophilic microorganisms for biotechnology. Extremophiles, 5, 7383.Google Scholar
Margesin, R., Neuner, G., Storey, B. (2007). Cold-loving microbes, plants, and animals-fundamentals and applied aspects. Naturwissenschaften, 94, 7799.Google Scholar
Marques, C.R. (2018). Extremophilic Microfactories: applications in Metal and Radionuclide Bioremediation. Frontiers in Microbiology, 9, 1191.Google Scholar
Martin, T.G., Watson, J.E.M. (2016). Intact ecosystems provide best defence against climate change. Nature Climate Change, 6, 122124.Google Scholar
Mecklenburg, C.W., Møller, P.R., Steinke, D. (2011). Biodiversity of Arctic marine fishes: taxonomy and zoogeography. Marine Biodiversity, 41, 109140.Google Scholar
Meidlinger, K., Tyler, P.A., Peck, L.S. (1998). Reproductive patterns in the Antarctic brachiopod Liothyrella uva. Marine Biology, 132, 153162.Google Scholar
Meredith, M.P., King, J.C. (2005). Climate change in the ocean to the west of the Antarctic Peninsula during the second half of the 20th century. Geophysics Research Letters, 32, L19604.Google Scholar
Merilä, J., Hendry, A.P. (2014). Climate change, adaptation, and phenotypic plasticity: the problem and the evidence. Evolutionary Applications, 7, 114; https://doi.org/10.1111/eva.12137.Google Scholar
Millennium Ecosystem Assessment Board (2005). Dryland systems. In: U. Safriel, Z. Adeel (Lead Authors) Ecosystems and Human Well-being: Current State and Trends. Island Press, Washington, DC, pp. 625656.Google Scholar
Moline, M.A., Karnovsky, N.J., Brown, Z., et al. (2008). High latitude changes in ice dynamics and their impact on polar marine ecosystems. Annals of the New York Academy of Sciences, 1134, 267319.Google Scholar
Moore, S.E., Reeves, R.R. (2018). Tracking arctic marine mammal resilience in an era of rapid ecosystem alteration. PLoS Biology, 16(10); doi:10.1371/journal.pbio.2006708.Google Scholar
Naudts, K., Chen, Y., McGrath, M.J., et al. (2016). Europe’s forest management did not mitigate climate warming. Science, 351, 597.Google Scholar
Notz, D., Stroeve, J. (2016). Observed Arctic sea-ice loss directly follows anthropogenic CO2 emission. Science, 354, 747750.Google Scholar
Parmesan, C. (2006). Ecological and evolutionary responses to recent climate change. Annual Review of Ecology Evolution and Systematics, 37, 637669.Google Scholar
Pearson, R.G. (2006). Climate change and the migration capacity of species. Trends in Ecology and Evolution, 21, 111113.Google Scholar
Peck, L.S. (2011). Organisms and responses to environmental change. Marine Genomics, 4, 237243.Google Scholar
Peck, L.S. (2018). Antarctic marine biodiversity: adaptations, environments and responses to change. Oceanography and Marine Biology Annual Review, 56, 105236.Google Scholar
Peck, L.S., Clark, M.S., Morley, S.A., Massey, A., Rossetti, H. (2009). Animal temperature limits and ecological relevance: effects of size, activity and rates of change. Functional Ecology, 23, 248253.Google Scholar
Peck, L.S., Barnes, D.K.A., Cook, A.J., Fleming, A.H., Clarke, A. (2010a). Negative feedback in the cold: ice retreat produces new carbon sinks in Antarctica. Global Change Biology, 16, 26142623; https://doi.org/10.1111/j.1365-2486.2009.02071.Google Scholar
Peck, L.S., Morley, S.A., Clark, M.S. (2010b). 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
Pedersen, M.W., Ruter, A., Schweger, C., et al. (2016). Postglacial viability and colonization in North America’s ice-free corridor. Nature, 537, 4549.Google Scholar
Pellissier, L., Bronken Eidesen, P., Ehrich, D., et al. (2016). Past climate-driven range shifts and population genetic diversity in arctic plants. Journal of Biogeography, 43, 461470.Google Scholar
Petersen, J.M., Dubilier, N. (2009). Methanotrophic symbioses in marine invertebrates. Environmental Microbiology Reports, 1(5), 319335; https://doi.org/10.1111/j.1758-2229.2009.00081.x.Google Scholar
Petit, R.J., Hampe, A. (2006). Some evolutionary consequences of being a tree. Annual Reviews of Ecology, Evolution and Systematics, 37, 187214.Google Scholar
Pörtner, H.-O., Farrell, A.P. (2008). Physiology and climate change. Science, 322, 690692.Google Scholar
Pörtner, H.-O., Knust, R. (2007). Climate change affects marine fishes through the oxygen limitation of thermal tolerance. Science, 315, 9597.Google Scholar
Reaser, J.K., Pomerance, R., Thomas, P.O. (2000). Coral bleaching and global climate change: scientific findings and policy recommendations. Conservation Biology, 14, 15001511.Google Scholar
Rintoul, S.R., Chown, S.L., DeConto, R.M., et al. (2018). Choosing the future of Antarctica. Nature, 558, 233241. https://doi.org/10.1038/s41586-018-0173-4.Google Scholar
Rogers, A.D. (2007). Evolution and biodiversity of Antarctic organisms: a molecular perspective. Philosophical Transactions of the Royal Society of London B Biological Sciences, 362, 21912214.Google Scholar
Rothschild, L.J., Mancinelli, R.L. (2001). Life in extreme environments. Nature, 409, 10921101.Google Scholar
Runting, R.K., Bryan, B.A., Dee, L.E., et al. (2017). Incorporating climate change into ecosystem service assessments and decisions: a review. Global Change Biology, 23, 2841; https://doi.org/10.1111/gcb.13457.Google Scholar
Seufferheld, M.J., Alvarez, H.M., Farias, M.E. (2008). Role of polyphosphates in microbial adaptation to extreme environments. Applied and Environmental Microbiology, 74, 58675874; https://doi.org/10.1128/AEM.00501-08.Google Scholar
Siddiqui, K.S., Cavicchioli, R. (2006). Cold-adapted enzymes. Annual Review of Biochemistry, 75, 403433.Google Scholar
Smetacek, V., Nicol, S. (2005). Polar ocean ecosystems in a changing world. Nature, 437, 362368.Google Scholar
Smith, W.O., Jr, Ainley, D.G., Arrigo, K.R., Dinniman, M.S. (2014). The oceanography and ecology of the Ross Sea Annual Review of Marine Science, 6, 469487.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. (2012). The physiology of global change: linking patterns to mechanisms. Annual Review of Marine Science, 4, 3961.Google Scholar
Taylor, B.L., Chivers, S.J., Larese, J., Perrin, W.F. (2007). Generation length and percent mature estimates for IUCN assessments of cetaceans. Administrative Report LJ-07–01, Southwest Fisheries Science Center, 8604 La Jolla Shores Blvd., La Jolla, CA 92038, USA.Google Scholar
Thomas, C.D., Cameron, A., Green, R.E., et al. (2004). Extinction risk from climate change. Nature, 427, 145148.Google Scholar
Thuiller, W. (2003). BIOMOD: optimising predictions of species distributions and projecting potential future shifts under global change. Global Change Biology, 9, 13531362.Google Scholar
Thuiller, W. (2004). Patterns and uncertainties of species’ range shifts under climate change. Global Change Biology, 10, 20202027.Google Scholar
Thuiller, W., Lavorel, S., Arau, M.B., Sykes, M.T., Prentice, I.C. (2005). Climate change threats to plant diversity in Europe. Proceedings of the National Academy of Sciences of the USA, 102, 82458250.Google Scholar
Travis, J.M.J. (2003). Climate change and habitat destruction: a deadly anthropogenic cocktail. Proceedings of the Royal Society B – Biological Science, 270, 467473.Google Scholar
Verde, C., di Prisco, G., Giordano, D., Russo, R., Anderson, D., Cowan, D. (2012). Antarctic psychrophiles: models for understanding the molecular basis of survival at low temperature and responses to climate change. Biodiversity, 13, 349356.Google Scholar
Verde, C., Giordano, D., Bellas, C.M., di Prisco, G., Anesio, A.M. (2016). Polar marine microorganisms and climate change. Advances in Microbial Physiology, 69, 187215.Google Scholar
Volis, S., Ormanbenkova, D., Shulgina, I. (2016). Role of selection and gene flow in population differentiation at the edge vs. interior of the species range differing in climatic conditions. Molecular Ecology, 25, 14491464.Google Scholar
Walker, M.D., Wahren, C.H., Hollister, R.D., et al. (2006). Plant community responses to experimental warming across the tundra biome. Proceedings of the National Academy of Sciences of the USA, 103, 13421346.Google Scholar
Walther, G.-R. (2010). Community and ecosystem responses to recent climate change. Philosophical Transactions of the Royal Society Biological Sciences B, 365, 20192024; https://doi.org/10.1098/rstb.2010.0021.Google Scholar
Walther, G.-R., Post, E., Convey, P., et al. (2002). Ecological responses to recent climate change. Nature, 416, 389395.Google Scholar
Weber, W., Fussenegger, M. (2012). Emerging biomedical applications of synthetic biology. Nature Reviews Genetics, 13, 2135.Google Scholar
Wenzel, L., Gilbert, N., Goldsworthy, L., et al. (2016). Polar opposites? Marine conservation tools and experiences in the changing Arctic and Antarctic: Marine Conservation Tools and Experiences in the Arctic and Antarctic. Aquatic Conservation Marine and Freshwater Ecosystems, 26(S2), 6184; https://doi.org/10.1002/aqc.2649.Google Scholar
Wilkins, D., Yau, S., Williams, T.J., et al. (2013). Key microbial drivers in Antarctic aquatic environments. FEMS Microbiology Reviews, 37(3), 303–303; https://doi.org/10.1111/1574-6976.12007.Google Scholar
Zhu, K., Woodall, C., Clark, J.S. (2012). Failure to migrate: lack of tree range expansion in response to climate change. Global Change Biology, 18, 10421052.Google Scholar
Zona, D., Gioli, B., Commane, R., et al. (2016). Cold season emissions dominate the Arctic tundra methane budget. Proceedings of the National Academy of Sciences of the USA, 113, 4045; https://doi.org/10.1073/pnas.1516017113.Google Scholar

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.Google Scholar
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.Google Scholar
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.CrossRefGoogle 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.Google Scholar
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.Google Scholar
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.Google Scholar
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.Google Scholar
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.Google 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.Google Scholar
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.Google 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.Google 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.Google 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.Google 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.Google 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.Google 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

References

Arrigo, K.R., van Dijken, G.L., Bushinsky, S. (2008). Primary production in the Southern Ocean, 1997–2006. Journal of Geophysical Research Oceans, 113, C08004; doi:10.1029/2007JC004551Google Scholar
Barbraud, C., Weimerskirch, H. (2001). Emperor penguins and climate change. Nature, 411, 183186.Google Scholar
Barnes, D.K.A. (1995). Sublittoral epifaunal communities at Signy Island, Antarctica. I. The ice-foot zone. Marine Biology, 121, 565572.Google Scholar
Bornemann, H., Held, C, Nachtsheim, D., et al. (2016). Seal research at the Drescher Inlet (SEADI). In: M. Schröder (ed.) The Expedition PS96 of the Research Vessel POLARSTERN to the southern Weddell Sea in 2015/16. Reports on Polar and Marine Research. Alfred-Wegener-Institute, Bremerhaven, pp. 116–29; doi:10.2312/BzPM_0700_2016Google Scholar
Bruchhausen, P.M., Raymond, J.A., Jacobs, S.S., et al. (1979). Fish, crustaceans, and the sea floor under the Ross Ice Shelf. Science, 203, 449451.Google Scholar
Cape, M.R., Vernet, M., Kahru, M., Spreen, G.M. (2014). Polynya dynamics drive primary production in the Larsen A and B embayments following ice shelf collapse. Journal of Geophysical Research: Oceans, 119, 572594; doi:10.1002/2013JC009441.Google Scholar
CAREX (2011). CAREX Roadmap for Research on Life in Extreme Environments. CAREX publication no. 9. Carex Project Office, Strasbourg.Google Scholar
Clarke, A., North, A.W. (1991). Is the growth of polar fish limited by temperature? In: Di Prisco, G, Maresca, B, Tota, B (eds) Biology of Antarctic Fish. Springer-Verlag, Berlin, pp. 5469; doi:10.1007/978-3-642-76217-8Google Scholar
Clarke, A., Griffiths, H., Barnes, D.K.A., Meredith, M.P., Grant, S.M. (2009). Spatial variation in seabed temperatures in the Southern Ocean: implications for benthic ecology and biogeography. Journal of Geophysical Research, 114, G03003; doi:10.1029/2008JG000886Google Scholar
Daly, M., Rack, F., Zook, R. (2013). Edwardsiella andrillae, a new species of sea anemone from Antarctic ice. PLoS ONE, 8(12), e83476.Google Scholar
Dayton, P.K. (1979). Observations of growth, dispersal and population dynamics of some sponges in Mc Murdo Sound, Antarctica. Colloques internationaux du C.N.R.S. Biologie des Spongaires, 291, 271282.Google Scholar
Dayton, P., Jarrell, S., Kim, S., et al. (2016). Surprising episodic recruitment and growth of Antarctic sponges: implications for ecological resilience. Journal of Experimental Marine Biology and Ecology, 482, 3855.Google Scholar
De Broyer, C., Koubbi, P., Griffiths, H.J., et al. (2014). Biogeographic Atlas of the Southern Ocean. SCAR, Cambridge.Google Scholar
di Prisco, G., Pisano, E., Clark, A. (1998). Fishes of Antarctica. A Biological Overview. Springer, Milan.Google Scholar
Flores, H., Atkinson, A., Kawaguchi, S., et al. (2012). Impact of climate change on Antarctic krill. Marine Ecology Progress Series, 458, 119.Google Scholar
Gerdes, D. (1992). Quantitative investigations on macrobenthos communities of the southeastern Weddell Sea shelf based on multibox corer samples. Polar Biology, 12, 291301.Google Scholar
Gieskes, W.W.C., Veth, C., Woehrmann, A., Graefe, M. (1987). Secchi disc visibility world record shattered. EOS, 68, 123.Google Scholar
Griffiths, H.J. (2010). Antarctic marine biodiversity – what do we know about the distribution of life in the Southern Ocean? PLoS ONE, 5(8), e11683; doi:10.1371/journal.pone.0011683Google Scholar
Gutt, J., Gerdes, D., Klages, M. (1992). Seasonality and spatial variability in the reproduction of two Antarctic holothurians (Echinodermata). Polar Biology, 11, 533544.Google Scholar
Gutt, J., Ekau, W. (1996). Habitat partitioning of dominant high Antarctic demersal fish in the Weddell Sea and Lazarev Sea. Journal of Experimental Marine Biology and Ecology, 206, 2537.Google Scholar
Gutt, J. (2000). Some ‘driving forces’ structuring communities of the sublittoral antarctic macrobenthos. Antarctic Science, 12, 297313.Google Scholar
Gutt, J. (2002). The Antarctic iceshelf: an extreme habitat for notothenioid fish. Polar Biology, 25, 320322.Google Scholar
Gutt, J., Piepenburg, D. (2003). Scale-dependent impact on diversity of Antarctic benthos caused by grounding of icebergs. Marine Ecology Progress Series, 253, 7783.Google Scholar
Gutt, J., Cape, M., Dimmler, W., et al. (2013). Shifts in Antarctic megabenthic structure after ice-shelf disintegration in the Larsen area east of the Antarctic Peninsula. Polar Biology, 36, 895906.Google Scholar
Jacob, U., Thierry, A., Brose, U., et al. (2011). The role of body size in complex food webs: a cold case. In Belgrano, A, Reiss, J (eds) Advances in Ecological Research, Vol. 45. Elsevier, Amsterdam, the Netherlands, pp. 181223.Google Scholar
Pardo, P.C., Pérez, F.F., Velo, A., Gilcoto, M. (2012). Water masses distribution in the Southern Ocean: improvement of an extended OMP (eOMP) analysis. Progress in Oceanography, 103, 92105; doi.org/10.1016/j.pocean.2012.06.002Google Scholar
Peck, L.S. (2005). Prospect for survival in the Southern Ocean: vulnerability of benthic species to temperature change. Antarctic Science, 17(4), 497507; doi.10.1017/S0954102005002920Google Scholar
Piepenburg, D., Schmid, M.K., Gerdes, D. (2002). The benthos off King George Island (South Shetland Islands, Antarctica): further evidence for a lack of a latitudinal biomass cline in the Southern Ocean. Polar Biology, 25, 146158.Google Scholar
Smith, C.R., De Leo, F.C., Bernardino, A.F., Sweetman, A.K., Martinez Arbizu, P. (2008). Abyssal food limitation, ecosystem structure and climate change. Trends in Ecology and Evolution, 23(9), 518529; doi:10.1016/j.tree.2008.05.002Google Scholar
Smith, C.R., Grange, L.J., Honig, D.L., et al. (2012). A large population of king crabs in Palmer Deep on the west Antarctic Peninsula shelf and potential invasive impacts. Proceedings of the Royal Society B Biological Sciences, 279, 1077–1026; doi:10.1098/rspb.2011.1496Google Scholar
Somero, G.N. (1991). Biochemical mechanisms of cold adaptation and stenothermality in Antarctic fish. In: Di Prisco, G, Maresca, B, Tota, B (eds) Biology of Antarctic Fish. Springer-Verlag, Berlin, pp. 232247; doi:10.1007/978-3-642-76217-8Google Scholar
Thomas, D.N., Dieckmann, G.S. (2002). Antarctic sea-ice – a habitat for extremophiles. Science, 295, 641644.Google Scholar
Thomas, D.N., Dieckmann, G.S. (2010). Sea Ice. Wiley Blackwell Publishing, Oxford.Google Scholar

References

Ausec, L., van Elsas, J.D., Mandic-Mulec, I. (2011). Two- and three-domain bacterial laccase-like genes are present in drained peat soils. Soil Biology and Biochemistry. Soil Biology and Biochemistry, 43(5), 975983.Google Scholar
Bailey, V.L., Smith, J.L., Bolton, H. (2002). Fungal-to-bacterial ratios in soils investigated for enhanced C sequestration. Soil Biology and Biochemistry, 34(7), 9971007.Google Scholar
Bakermans, C., Tsapin, A.I., Souza-Egipsy, V., Gilichinsky, D.A., Nealson, K.H. (2003). Reproduction and metabolism at -10°C of bacteria isolated from Siberian permafrost. Environmental Microbiology, 5(4), 321326.Google Scholar
Baldrian, P., Valášková, V. (2008). Degradation of cellulose by basidiomycetous fungi. FEMS Microbiology Reviews, 32(3), 501521.Google Scholar
Barka, E.A., Vatsa, P., Sanchez, L., et al. (2016). Taxonomy, physiology, and natural products of actinobacteria. Microbiology and Molecular Biology Reviews, 80(1), 143.Google Scholar
Bellemain, E., Davey, M.L., Kauserud, H., et al. (2013). Fungal palaeodiversity revealed using high-throughput metabarcoding of ancient DNA from arctic permafrost. Environmental Microbiology, 15(4), 11761189.Google Scholar
Bugg, T.D.H., Ahmad, M., Hardiman, E.M., Rahmanpour, R. (2011). Pathways for degradation of lignin in bacteria and fungi. Natural Product Reports, 28(12), 18831896.Google Scholar
Čapek, P., Diáková, K., Dickopp, J.-E., et al. (2015). The effect of warming on the vulnerability of subducted organic carbon in arctic soils. Soil Biology and Biochemistry, 90, 1929.Google Scholar
Coolen, M.J.L., Orsi, W.D. (2015). The transcriptional response of microbial communities in thawing Alaskan permafrost soils. Frontiers in Microbiology, 6, 197. doi:10.3389/fmicb.2015.00197Google Scholar
Costa, O.Y.A., Raaijmakers, J.M., Kuramae, E.E. (2018). Microbial extracellular polymeric substances: Ecological function and impact on soil aggregation. Frontiers in Microbiology, 9, 1636. doi:10.3389/fmicb.2018.01636Google Scholar
Czapski, T.R., Trun, N. (2014). Expression of csp genes in E. Coli K-12 in defined rich and defined minimal media during normal growth, and after cold-shock. Gene, 547(1), 9197. doi:10.1016/j.gene.2014.06.033Google Scholar
Dao, T.T., Gentsch, N., Mikutta, R., et al. (2018). Fate of carbohydrates and lignin in north-east Siberian permafrost soils. Soil Biology and Biochemistry, 116, 311322.Google Scholar
Davidson, E.A., Janssens, I.A. (2006). Temperature sensitivity of soil carbon decomposition and feedbacks to climate change. Nature, 440(7081), 165173.Google Scholar
De Boer, W., Folman, L.B., Summerbell, R.C., Boddy, L. (2005). Living in a fungal world: Impact of fungi on soil bacterial niche development. FEMS Microbiology Reviews, 29(4), 795811.Google Scholar
DeAngelis, K.M., Allgaier, M., Chavarria, Y., et al. (2011). Characterization of trapped lignin-degrading microbes in tropical forest soil. PLoS ONE, 6(4), e19306.Google Scholar
Dungait, J.A.J., Hopkins, D.W., Gregory, A.S., Whitmore, A.P. (2012). Soil organic matter turnover is governed by accessibility not recalcitrance. Global Change Biology, 18(6), 17811796.Google Scholar
Eilers, K.G., Debenport, S., Anderson, S., Fierer, N. (2012). Digging deeper to find unique microbial communities: The strong effect of depth on the structure of bacterial and archaeal communities in soil. Soil Biology and Biochemistry, 50, 5865.Google Scholar
Fierer, N., Allen, A.S., Schimel, J.P., Holden, P.A. (2003a). Controls on microbial CO2 production: A comparison of surface and subsurface soil horizons. Global Change Biology, 9(9), 13221332.Google Scholar
Fierer, N., Schimel, J.P., Holden, P.A. (2003b). Variations in microbial community composition through two soil depth profiles. Soil Biology and Biochemistry, 35(1), 167176.Google Scholar
Finore, I., Di Donato, P., Mastascusa, V., Nicolaus, B., Poli, A. (2014). Fermentation technologies for the optimization of marine microbial exopolysaccharide production. Marine Drugs, 12(5), 30053024.Google Scholar
Frey, B., Rime, T., Phillips, M., et al. (2016). Microbial diversity in European alpine permafrost and active layers. FEMS Microbiology Ecology, 92(3).Google Scholar
Gentsch, N., Mikutta, R., Alves, R.J.E., et al. (2015). Storage and transformation of organic matter fractions in cryoturbated permafrost soils across the Siberian Arctic. Biogeosciences, 12(14), 45254542.Google Scholar
Gentsch, N., Wild, B., Mikutta, R., et al. (2018). Temperature response of permafrost soil carbon is attenuated by mineral protection. Global Change Biology, 24, 34013415. doi:10.1111/gcb.14316Google Scholar
Gilichinsky, D., Rivkina, E., Bakermans, C., et al. (2005). Biodiversity of cryopegs in permafrost. FEMS Microbiology Ecology, 53, 117128. doi:10.1016/j.femsec.2005.02.003Google Scholar
Gittel, A., Bárta, J., Kohoutová¡, I., et al. (2014a). Site- and horizon-specific patterns of microbial community structure and enzyme activities in permafrost-affected soils of Greenland. Frontiers in Microbiology, 5. doi:10.3389/fmicb.2014.00541Google Scholar
Gittel, A., Bárta, J., Kohoutová, I., et al. (2014b). Distinct microbial communities associated with buried soils in the Siberian tundra. ISME Journal, 8(4), 841853. doi:10.1038/ismej.2013.219Google Scholar
Orwin, K.H., Kirschbaum, M.U.F., St John, M.G., Dickie, I.A. (2011). Organic nutrient uptake by mycorrhizal fungi enhances ecosystem carbon storage: a model‐based assessment. Ecology Letters, 14, 493502. doi:10.1111/j.1461-0248.2011.01611.xGoogle Scholar
Harden, J.W., Koven, C.D., Ping, C.L., et al. (2012). Field information links permafrost carbon to physical vulnerabilities of thawing. Geophysical Research Letters, 39(15). doi:10.1029/2012GL051958Google Scholar
Hartmann, M., Lee, S., Hallam, S.J., Mohn, W.W. (2009). Bacterial, archaeal and eukaryal community structures throughout soil horizons of harvested and naturally disturbed forest stands. Environmental Microbiology, 11(12), 30453062. doi:10.1111/j.1462-2920.2009.02008.xGoogle Scholar
Hibbett, D.S., Ohman, A., Glotzer, D., et al. (2011). Progress in molecular and morphological taxon discovery in Fungi and options for formal classification of environmental sequences. Fungal Biology Reviews, 25(1), 3847. doi:10.1016/j.fbr.2011.01.001Google Scholar
Hobbie, J.E., Hobbie, E.A. (2013). Microbes in nature are limited by carbon and energy: The starving-survival lifestyle in soil and consequences for estimating microbial rates. Frontiers in Microbiology, 4. doi:10.3389/fmicb.2013.00324Google Scholar
Hodge, A., Campbell, C.D., Fitter, A.H. (2001). An arbuscular mycorrhizal fungus accelerates decomposition and acquires nitrogen directly from organic material. Nature, 413(6853), 297299.Google Scholar
Hoshino, T., Xiao, N., Tkachenko, O.B. (2009). Cold adaptation in the phytopathogenic fungi causing snow molds. Mycoscience, 50(1), 2638.Google Scholar
Hu, W., Zhang, Q., Li, D., et al. (2014). Diversity and community structure of fungi through a permafrost core profile from the Qinghai-Tibet Plateau of China. Journal of Basic Microbiology, 54(12), 13311341.Google Scholar
Hugelius, G., Kuhry, P., Tarnocai, C. (2016). Ideas and perspectives: Holocene thermokarst sediments of the Yedoma permafrost region do not increase the northern peatland carbon pool. Biogeosciences, 13(7), 20032010.Google Scholar
IPCC (2014). AR5 Climate Change 2014: Impacts, Adaptation, and Vulnerability. www.ipcc.ch/report/ar5/wg2/Google Scholar
Iversen, C.M., Sloan, V.L., Sullivan, P.F., et al. (2015). The unseen iceberg: Plant roots in arctic tundra. New Phytologist, 205(1), 3458.Google Scholar
Jansson, J.K., Taş, N. (2014). The microbial ecology of permafrost. Nature Reviews Microbiology, 12(6), 414425.Google Scholar
Jobbágy, E.G., Jackson, R.B. (2000). The vertical distribution of soil organic carbon and its relation to climate and vegetation. Ecological Applications, 10(2), 423436.Google Scholar
Johnson, S.S., Hebsgaard, M.B., Christensen, T.R., et al. (2007). Ancient bacteria show evidence of DNA repair. Proceedings of the National Academy of Sciences of the United States of America, 104(36), 1440114405.Google Scholar
Kaiser, C., Meyer, H., Biasi, C., et al. (2007). Conservation of soil organic matter through cryoturbation in arctic soils in Siberia. Journal of Geophysical Research: Biogeosciences, 112(2).Google Scholar
Kochkina, G., Ivanushkina, N., Ozerskaya, S., et al. (2012). Ancient fungi in Antarctic permafrost environments. FEMS Microbiology Ecology, 82(2), 501509.Google Scholar
Koven, C.D., Ringeval, B., Friedlingstein, P., et al. (2011). Permafrost carbon-climate feedbacks accelerate global warming. Proceedings of the National Academy of Sciences of the United States of America, 108(36), 1476914774.Google Scholar
Lawrence, D.M., Koven, C.D., Swenson, S.C., Riley, W.J., Slater, A.G. (2015). Permafrost thaw and resulting soil moisture changes regulate projected high-latitude CO2 and CH4 emissions. Environmental Research Letters, 10(9). doi:10.1088/1748-9326/10/9/094011Google Scholar
Le Roes-Hill, M., Khan, N., Burton, S.G. (2011). Actinobacterial peroxidases: An unexplored resource for biocatalysis. Applied Biochemistry and Biotechnology, 164(5), 681713.Google Scholar
Lipson, D.A., Haggerty, J.M., Srinivas, A., et al. (2013). metagenomic insights into anaerobic metabolism along an arctic peat soil profile. PLoS ONE, 8(5). doi:10.1371/journal.pone.0064659Google Scholar
Lydolph, M.C., Jacobsen, J., Arctander, P., et al. (2005). Beringian paleoecology inferred from permafrost-preserved fungal DNA. Applied and Environmental Microbiology, 71(2), 10121017.Google Scholar
Lykidis, A., Mavromatis, K., Ivanova, N., et al. (2007). Genome sequence and analysis of the soil cellulolytic actinomycete Thermobifida fusca YX. Journal of Bacteriology, 189(6), 24772486.Google Scholar
MacKelprang, R., Waldrop, M.P., Deangelis, K.M., et al. (2011). Metagenomic analysis of a permafrost microbial community reveals a rapid response to thaw. Nature, 480(7377), 368371. doi:10.1038/nature10576Google Scholar
MacKelprang, R., Burkert, A., Haw, M., et al. (2017). Microbial survival strategies in ancient permafrost: Insights from metagenomics. ISME Journal, 11(10), 23052318.Google Scholar
McCarthy, A.J. (1987). Lignocellulose-degrading actinomycetes. FEMS Microbiology Letters, 46(2), 145163.Google Scholar
McMahon, S.K., Wallenstein, M.D., Schimel, J.P. (2011). A cross-seasonal comparison of active and total bacterial community composition in Arctic tundra soil using bromodeoxyuridine labeling. Soil Biology and Biochemistry, 43(2), 287295.Google Scholar
Mondav, R., McCalley, C.K., Hodgkins, S.B., et al. (2017). Microbial network, phylogenetic diversity and community membership in the active layer across a permafrost thaw gradient. Environmental Microbiology, 19(8), 32013218.Google Scholar
Mueller, G.M., Schmit, J.P. (2007). Fungal biodiversity: What do we know? What can we predict? Biodiversity and Conservation, 16(1), 15.Google Scholar
Ozerskaya, S., Kochkina, G., Ivanushkina, N., Gilichinsky, D.A. (2009). Fungi in permafrost. In: Margesin, R. (ed.) Permafrost Soils. Soil Biology, vol 16. Springer, Berlin, Heidelberg.Google Scholar
Phadtare, S. (2004). Recent developments in bacterial cold-shock response. Current Issues in Molecular Biology, 6(2), 125136.Google Scholar
Ping, C.L., Jastrow, J.D., Jorgenson, M.T., Michaelson, G.J., Shur, Y.L. (2015). Permafrost soils and carbon cycling. Soil, 1(1), 147171.Google Scholar
Rivkina, E.M., Kraev, G.N., Krivushin, K.V., et al. (2006). Methane in permafrost of Northeastern Arctic. Earth’s Cryosphere, 10(3), 2341.Google Scholar
Robinson, C.H. (2001). Cold adaptation in Arctic and Antarctic fungi. New Phytologist, 151(2), 341353.Google Scholar
Šantrůčková, H., Kotas, P., Bárta, J., et al. (2018). Significance of dark CO2 fixation in arctic soils. Soil Biology and Biochemistry, 119. doi:10.1016/j.soilbio.2017.12.021Google Scholar
Schellenberger, S., Kolb, S., Drake, H.L. (2010). Metabolic responses of novel cellulolytic and saccharolytic agricultural soil Bacteria to oxygen. Environmental Microbiology, 12(4), 845861.Google Scholar
Schmidt, G. (2011). Climate change and climate modeling. Eos, Transactions American Geophysical Union, 92(23), 198199.Google Scholar
Schnecker, J., Wild, B., Hofhansl, F., et al. (2014). Effects of soil organic matter properties and microbial community composition on enzyme activities in cryoturbated arctic soils. PLoS ONE, 9(4).Google Scholar
Schuur, E.A.G., Abbott, B. (2011). Climate change: High risk of permafrost thaw. Nature, 480(7375), 3233.Google Scholar
Schuur, E.A.G., Bockheim, J., Canadell, J.G., et al. (2008). Vulnerability of permafrost carbon to climate change: Implications for the global carbon cycle. BioScience, 58(8), 701714.Google Scholar
Soil Survey Staff (2010). Keys to Soil Taxonomy, 11th ed. USDA-NRCS, Washington DC.Google Scholar
Soina, V.S., Vorobiova, E.A., Zvyagintsev, D.G., Gilichinsky, D.A. (1995). Preservation of cell structures in permafrost: A model for exobiology. Advances in Space Research, 15, 237242. doi:10.1016/S0273-1177(99)80090-8Google Scholar
Talbot, J.M., Treseder, K.K. (2010). Controls over mycorrhizal uptake of organic nitrogen. Pedobiologia, 53(3), 169179. doi:10.1016/j.pedobi.2009.12.001Google Scholar
Talbot, J.M., Allison, S.D., Treseder, K.K. (2008). Decomposers in disguise: Mycorrhizal fungi as regulators of soil C dynamics in ecosystems under global change. Functional Ecology, 22(6), 955963.Google Scholar
Tarnocai, C. (2018). The amount of organic carbon in various soil orders and Ecological Provinces in Canada. In: Harms, D., Korschens, M., Olson, G., et al. (eds) Soil Processes and the Carbon Cycle. CRC Press, Boca Raton, FL. doi:10.1201/9780203739273Google Scholar
Tarnocai, C. (2006). The effect of climate change on carbon in Canadian peatlands. Global and Planetary Change 53(4), 222232.Google Scholar
Tarnocai, C., Bockheim, J. (2011). Cryosolic soils of Canada: Genesis, distribution, and classification. Canadian Journal of Soil Science, 91(5), 749762.Google Scholar
Tibbett, M., Sanders, F.E., Cairney, J.W.G. (2002). Low-temperature-induced changes in trehalose, mannitol and arabitol associated with enhanced tolerance to freezing in ectomycorrhizal basidiomycetes (Hebeloma spp.). Mycorrhiza, 12(5), 249255.Google Scholar
Tisdall, J.M., Oades, J.M. (1982). Organic matter and water‐stable aggregates in soils. Journal of Soil Science, 33(2), 141163.Google Scholar
Toljander, J.F., Lindahl, D., Paul, L.R., Elfstrand, M., Finlay, R.D. (2007). Influence of arbuscular mycorrhizal mycelial exudates on soil bacterial growth and community structure. FEMS Microbiology Ecology, 61, 295304. doi:10.1111/j.1574-6941.2007.00337.xGoogle Scholar
Tveit, A., Schwacke, R., Svenning, M.M., Urich, T. (2013). Organic carbon transformations in high-Arctic peat soils: Key functions and microorganisms. ISME Journal, 7(2), 299311.Google Scholar
Tveit, A.T., Urich, T., Frenzel, P., Svenning, M.M. (2015). Metabolic and trophic interactions modulate methane production by Arctic peat microbiota in response to warming. Proceedings of the National Academy of Sciences of the United States of America, 112(19), E2507–2516.Google Scholar
Uchimura, H., Enjoji, H., Seki, T., et al. (2002). Nitrate reductase-formate dehydrogenase couple involved in the fungal denitrification by Fusarium oxysporum. Journal of Biochemistry, 131(4), 579586.Google Scholar
Van Der Heul, H.U., Bilyk, B.L., McDowall, K.J., Seipke, R.F., Van Wezel, G.P. (2018). Regulation of antibiotic production in Actinobacteria: New perspectives from the post-genomic era. Natural Product Reports, 35(6), 575604.Google Scholar
Weinstein, R.N., Montiel, P.O., Johnstone, K. (2007). Influence of growth temperature on lipid and soluble carbohydrate synthesis by fungi isolated from Fellfield Soil in the Maritime Antarctic. Mycologia, 92(2), 222229.Google Scholar
Wild, B., Schnecker, J., Bárta, J., et al. (2013). Nitrogen dynamics in Turbic Cryosols from Siberia and Greenland. Soil Biology and Biochemistry, 67. doi:10.1016/j.soilbio.2013.08.004Google Scholar
Wild, B., Schnecker, J., Alves, R.J.E., et al. (2014). Input of easily available organic C and N stimulates microbial decomposition of soil organic matter in arctic permafrost soil. Soil Biology and Biochemistry, 75, 143151. doi:10.1016/j.soilbio.2014.04.014Google Scholar
Wild, B., Schnecker, J., Knoltsch, A., et al. (2015). Microbial nitrogen dynamics in organic and mineral soil horizons along a latitudinal transect in western Siberia. Global Biogeochemical Cycles, 29(5), 567582. doi:10.1002/2015GB005084Google Scholar
Wild, B., Gentsch, N., Capek, P., et al. (2016). Plant-derived compounds stimulate the decomposition of organic matter in arctic permafrost soils. Scientific Reports, 6. doi:10.1038/srep25607Google Scholar
Wild, B., Alves, R.J.E., Bárta, J., et al. (2018). Amino acid production exceeds plant nitrogen demand in Siberian tundra. Environmental Research Letters, 13. doi:10.1088/1748-9326/aaa4faGoogle Scholar
Wilhelm, R.C., Niederberger, T.D., Greer, C., Whyte, L.G. (2011). Microbial diversity of active layer and permafrost in an acidic wetland from the Canadian High Arctic. Canadian Journal of Microbiology, 57(4), 303315.Google Scholar
Xiao, D., Peng, S.P., Wang, E.Y. (2015). Fermentation enhancement of methanogenic archaea consortia from an Illinois basin coalbed via DOL emulsion nutrition. PLoS ONE, 10(4).Google Scholar
Yergeau, E., Hogues, H., Whyte, L.G., Greer, C.W. (2010). The functional potential of high Arctic permafrost revealed by metagenomic sequencing, qPCR and microarray analyses. ISME Journal, 4(9), 12061214.Google Scholar
Yong‐Liang, C., Ye, D., Jin‐Zhi, D., et al. (2017). Distinct microbial communities in the active and permafrost layers on the Tibetan Plateau. Molecular Ecology, 26, 66086620. doi:10.1111/mec.14396Google Scholar
Zhou, L., Xing, X., Peng, B., Fang, G. (2016). Extracellular polymeric substance (EPS) characteristics and comparison of suspended and attached activated sludge at low temperatures. Qinghua Daxue Xuebao/Journal of Tsinghua University, 56(9), 10091015.Google Scholar

References

Amsler, C.D., Fairhead, V.A. (2006). Defensive and sensory chemical ecology of brown algae. Advances in Botanical Research, 43, 191.Google Scholar
Amsler, C.D., Iken, K.B., McClintock, J.B., Baker, B.J. (2001). Secondary metabolites from Antarctic marine organisms and their ecological implications. In: McClintock, J.B., Baker, B.J. (eds) Marine Chemical Ecology. CRC Press, Boca Raton, FL, pp. 267300.Google Scholar
Amsler, C.D., Iken, K.B., McClintock, J.B., et al. (2005). Comprehensive evaluation of the palatability and chemical defences of subtidal macroalgae from the Antarctic Peninsula. Marine Ecology Progress Series, 294, 141159.Google Scholar
Amsler, C.D., McClintock, J.B., Baker, B.J. (2008). Macroalgal chemical defenses in polar marine communities. In: Amsler, C. D. (ed.) Algal Chemical Ecology. Springer-Verlag,Berlin, pp. 91103.Google Scholar
Amsler, C.D., Iken, K., McClintock, J.B., Baker, B.J. (2009). Defenses of polar macroalgae against herbivores and biofoulers. Botanica Marina, 52, 535545.Google Scholar
Amsler, C.D., McClintock, J.B., Baker, B.J. (2014). Chemical mediation of mutualistic interactions between macroalgae and mesograzers structure unique coastal communities along the western Antarctic Peninsula. Journal of Phycology, 50, 110.Google Scholar
Angulo-Preckler, C., Cid, C., Oliva, F., Avila, C. (2015). Antifouling activity in some benthic Antarctic invertebrates by ‘in situ’ experiments at Deception Island, Antarctica. Marine Environmental Research, 105, 3038.Google Scholar
Angulo-Preckler, C., San Miguel, O., Garcia-Aljaro, C., Avila, C. (2018). Antibacterial defenses and palatability of shallow-water Antarctic sponges. Hydrobiologia, 806, 123128.Google Scholar
Ankisetty, S., Nandiraju, S., Win, H., et al. (2004). Chemical investigation of predator-deterred macroalgae from the Antarctic Peninsula. Journal of Natural Products, 67, 12951302.Google Scholar
Antonov, A.S., Avilov, S.A., Kalinovsky, A.I., et al. (2008). Triterpene glycosides from Antarctic sea cucumbers. 1. Structure of Liouvillosides A1, A2, A3, B1, and B2 from the sea cucumber Staurocucumis liouvillei: new procedure for separation of highly polar glycoside fractions and taxonomic revision. Journal of Natural Products, 71, 16771685.Google Scholar
Antonov, A.S., Avilov, S.A., Kalinovsky, A.I., et al. (2009). Triterpene glycosides from Antarctic sea cucumbers. 2. Structure of Achlioniceosides A (1), A (2), and A (3) from the sea cucumber Achlionice violaecuspidata (=Rhipidothuria racowitzai). Journal of Natural Products, 72, 3338.Google Scholar
Antonov, A.S., Avilov, S.A., Kalinovsky, A.I., et al. (2011). Triterpene glycosides from Antarctic sea cucumbers III. Structures of liouvillosides A (4) and A (5), two minor disulphated tetraosides containing 3-O-methylquinovose as terminal monosaccharide units from the sea cucumber Staurocucumis liouvillei (Vaney). Natural Product Research, 25, 13241333.Google Scholar
Appleton, D.R., Chuen, C.S., Berridge, M.V., Webb, V.L., Copp, B.R. (2009). Rossinones, A., B, biologically active meroterpenoids from the Antarctic Ascidian, Aplidium species. Journal of Organic Chemistry, 74, 91959198.Google Scholar
Argandona, V.H., Rovirosa, J., San-Martin, A., et al. (2002). Antifeedant effect of marine halogenated monoterpenes. Journal of Agricultural and Food Chemistry, 50, 70297033.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
Aumack, C.F., Amsler, C.D., McClintock, J.B., Baker, B.J. (2010). Chemically mediated resistance to meso-herbivory in finely branched macroalgae along the western Antarctic Peninsula. European Journal of Phycology, 45, 1926.Google Scholar
Avila, C. (2016a). Ecological and pharmacological activities of Antarctic marine natural products. Planta Medica, 82, 767774.Google Scholar
Avila, C. (2016b). Biological and chemical diversity in Antarctica: from new species to new natural products. Biodivers, 17, 511.Google Scholar
Avila, C., Iken, K.B., Fontana, A., Gimino, G. (2000). Chemical ecology of the Antarctic nudibranch Bathydoris hodgsoni Eliot, 1907: defensive role and origin of its natural products. Journal of Experimental Marine Biology and Ecology, 252, 2744.Google Scholar
Avila, C., Taboada, S., Núñez-Pons, L. (2008). Antarctic marine chemical ecology: what is next? Marine Ecology, 29, 171.Google Scholar
Avila, C., Núñez-Pons, L., Moles, J. (2018). From the tropics to the poles: chemical defensive strategies in sea slugs (Mollusca: Heterobranchia). In: Puglisi, M. P., Becerro, M. A. (eds) Chemical Ecology: The Ecological Impacts of Marine Natural Products. CRC Press, Boca Raton, FL, pp. 71163.Google Scholar
Barnes, D.K.A., Griffiths, H.J., Kaiser, S. (2009). Geographic range shift responses to climate change by Antarctic benthos: where we should look. Marine Ecology Progress Series, 393, 1326.Google Scholar
Barnes, D.K.A., Fenton, M., Cordingley, A. (2014). Climate-linked iceberg activity massively reduces spatial competition in Antarctic shallow waters. Current Biology, 24, 553554.Google Scholar
Best, B.A., Winston, J.E. (1984). Skeletal strength of encrusting cheilostome bryozoans. Biological Bulletin, 167, 390409.Google Scholar
Blunt, J.W., Copp, B.R., Hu, W.P., et al. (2007). Marine natural products. Natural Product Report, 24, 3186.Google Scholar
Blunt, J.W., Carroll, A.R., Copp, B.R., Keyzers, R.A., Davis, R.A. (2018). Marine natural products. Natural Product Report, 35, 853.Google Scholar
Bryan, P., McClintock, J., Slattery, M., Rittschof, D. (2003). A comparative study of the non-acidic chemically mediated antifoulant properties of three sympatric species of ascidians associated with seagrass habitats. Biofouling, 19, 235245.Google Scholar
Bucolo, P., Amsler, C.D., McClintock, J.B., Baker, B.J. (2011). Palatability of the Antarctic rhodophyte Palmaria decipiens (Reinsch) RW Ricker and its endo/epiphyte Elachista antarctica Skottsberg to sympatric amphipods. Journal of Experimental Marine Biology and Ecology, 396, 202206.Google Scholar
Campbell, A.H., Harder, T., Nielsen, S., Kjelleberg, S., Steinberg, P.D. (2011). Climate change and disease: bleaching of a chemically defended seaweed. Global Change Biology, 17, 29582970.Google Scholar
Carbone, M., Nuñez-Pons, L., Castelluccio, F., Avila, C., Gavagnin, M. (2009). Illudalene sesquiterpenoids of the alcyopterosin series from the Antarctic marine soft-coral Alcyonium grandis. Journal of Natural Products, 72, 13571360.Google Scholar
Carbone, M., Núñez-Pons, L., Paone, M., et al. (2012). Rossinone-related meroterpenes from the Antarctic ascidian Aplidium fuegiense. Tetrahedron, 68, 35413544.Google Scholar
Carbone, M., Nunez-Pons, L., Ciavatta, M.L., et al. (2014). Occurrence of a taurine derivative in an Antarctic glass sponge. Natural Product Communications, 9, 469470.Google Scholar
Ciaglia, E., Malfitano, A.M., Laezza, C., et al. (2017). Immuno-modulatory and anti-inflammatory effects of dihydrogracilin A, a terpene derived from the marine sponge Dendrilla membranosa. International Journal of Molecular Sciences, 18, 1643.Google Scholar
Clark, M.S., Peck, L.S. (2009a). HSP70 Heat shock proteins and environmental stress in Antarctic marine organisms: a mini-review. Marine Genomics, 2, 1118.Google Scholar
Clark, M.S., Peck, L.S. (2009b). Triggers of the HSP70 stress response: environmental responses and laboratory manipulation in an Antarctic marine invertebrate (Nacella concinna). Cell Stress Chaperones, 14, 649660.Google Scholar
Constable, A.J., Melbourne-Thomas, J., Corney, S.P., et al. (2014). Climate change and Southern Ocean ecosystems I: how changes in physical habitats directly affect marine biota. Global Change Biology, 20, 30043025.Google Scholar
Cutignano, A., Moles, J., Avila, C., Fontana, A. (2015). Granuloside, a unic linear homosesterterpene from the Antarctic nudibranch Charcotia granulosa. Journal of Natural Products, 78, 17611764.Google Scholar
Cutignano, A., Zhang, W., Avila, C., Cimino, G., Fontana, A. (2011). Intrapopulation variability in the terpene metabolism of the Antarctic opisthobranch mollusc Austrodoris kerguelenensis. European Journal of Organic Chemistry, 27, 53835389.Google Scholar
Cutignano, A., De Palma, R., Fontana, A. (2012). A chemical investigation of the Antarctic sponge Lyssodendoryx flabellata. Natural Product Research, 26, 12401248.Google Scholar
Daoust, J., Chen, M., Wang, M., et al. (2013). Sesterterpenoids isolated from a northeastern Pacific Phorbas sp. Journal of Organic Chemistry, 78, 82678273.Google Scholar
Davidson, S.K., Haygood, M.G. (1999). Identification of sibling species of the bryozoan Bugula neritina that produce different anticancer bryostatins and harbor distinct strains of the bacterial symbiont Candidatus endobugula sertula. Biological Bulletin, 196, 273280.Google Scholar
Davis, A.R., Bremner, J.B. (1999). Potential antifouling natural products from ascidians: a review. In: Fingerman, M, Nagabhushanam, R, Thompson, M.F. (eds) Recent Advances in Marine Biotechnology, Vol. III. Science Publishers, New Hampshire, pp. 259308.Google Scholar
Dayton, P.K. (1989). Interdecadal variation in an Antarctic sponge and its predators from oceanographic climate shifts. Science, 245, 14841486.Google Scholar
Dayton, P.K., Robilliard, G.A., Paine, R.T., Dayton, L. B. (1974). Biological accommodation in the benthic community at McMurdo Sound, Antarctica. Ecological Monographs, 44, 105128.Google Scholar
De Broyer, C., Danis, B. (2011). How many species in the Southern Ocean? Towards a dynamic inventory of the Antarctic marine species. Deep-Sea Research Pt II, 58, 517.Google Scholar
De Broyer, C., Koubbi, P., Griffiths, H.J., et al. (2014). Biogeographic Atlas of the Southern Ocean. Scientific Committee on Antarctic Research, Cambridge, UK.Google Scholar
Díaz, J.I., Fusaro, B., Vidal, V., et al. (2017). Macroparasites in Antarctic penguins. In:Klimpel, S, Kuhn, T, Mehlhorn, H (eds) Biodiversity and Evolution of Parasitic Life in the Southern Ocean. Springer International Publishing, Cham, Switzerland, pp. 183204.Google Scholar
Díaz-Marrero, A.R., Brito, I., Dorta, E., et al. (2003). Caminatal, an aldehyde sesterterpene with a novel carbon skeleton from the Antarctic sponge Suberites caminatus. Tetrahedron Letters, 44, 59395942.Google Scholar
Díaz-Marrero, A.R., Brito, I., Cueto, M., San-Martin, A., Darias, J. (2004). Suberitane network, a taxonomical marker for Antarctic sponges of the genus Suberites? Novel sesterterpenes from Suberites caminatus. Tetrahedron Letters, 45, 47074710.Google Scholar
Diyabalanage, T., Amsler, C.D., McClintock, J.B., Baker, B.J. (2006). Palmerolide A, a cytotoxic Macrolide from the Antarctic tunicate Synoicum adareanum. Journal of the American Chemical Society, 128, 56305631.Google Scholar
Downey, R.V., Griffiths, H.J., Linse, K., Janussen, D. (2012). Diversity and distribution patterns in high southern latitude sponges. PLoS One, 7, e41672.Google Scholar
Ducklow, H.W., Fraser, W.R., Meredith, M.P., et al. (2013). West Antarctic Peninsula: an ice-dependent coastal marine ecosystem in transition. Oceanography, 26, 190203.Google Scholar
Duckworth, A.R., Battershill, C.N. (2001). Population dynamics and chemical ecology of New Zealand Demospongiae Latrunculia sp. nov., Polymastia croceus (Poecilosclerida: Latrunculiidae: Polymastiidae). New Zealand Journal of Marine and Freshwater Research, 35, 935949.Google Scholar
Erickson, A.A., Paul, V.J., Alstyne, K.L., Van Kwiatkowski, L.M. (2006). Palatability of macroalgae that use different types of chemical defenses. Journal of Chemical Ecology, 32, 18831895.Google Scholar
Fabry, V.J., Seibel, B.A., Feely, R.A., Orr, J.C. (2008). Impacts of ocean acidification on marine fauna and ecosystem processes. ICES Journal of Marine Sciences, 65, 414432.Google Scholar
Fairhead, V.A., Amsler, C.D., Mcclintock, J.B., Baker, B.J. (2005). Within-thallus variation in chemical and physical defenses in two species of ecologically dominant brown macroalgae from the Antarctic Peninsula. Journal of Experimental Marine Biology and Ecology, 322, 112.Google Scholar
Ferretti, C., Vacca, S., De Ciucis, C., et al. (2009). Growth dynamics and bioactivity variation of the Mediterranean demosponges Agelas oroides (Agelasida, Agelasidae) and Petrosia ficiformis (Haplosclerida, Petrosiidae). Marine Ecology, 30, 110.Google Scholar
Figuerola, B., Monleón-Getino, T., Ballesteros, M., Avila, C. (2012a). Spatial patterns and diversity of bryozoan communities from the Southern Ocean: South Shetland Islands, Bouvet Island and Eastern Weddell Sea. Systematics and Biodiversity, 10, 109123.Google Scholar
Figuerola, B., Núñez-Pons, L., Vázquez, J., et al. (2012b). Chemical interactions in Antarctic marine benthic ecosystems. In: Cruzado, A (ed.) Marine Ecosystems. InTech, Rijeka, pp. 105126.Google Scholar
Figuerola, B., Núñez-Pons, L., Moles, J., Avila, C. (2013a). Feeding repellence in Antarctic bryozoans. Naturwissenschaften, 100, 10691081.Google Scholar
Figuerola, B., Taboada, S., Monleón-Getino, T., Vázquez, J., Avila, C. (2013b). Cytotoxic activity of Antarctic benthic organisms against the common sea urchin Sterechinus neumayeri. Oceanography, 1, 2.Google Scholar
Figuerola, B., Núñez-Pons, L., Monleón-Getino, T., Avila, C. (2014a). Chemo-ecological interactions in Antarctic bryozoans. Polar Biology, 37, 10171030.Google Scholar
Figuerola, B., Sala-Comorera, L., Angulo-Preckler, C., et al. (2014b). Antimicrobial activity of Antarctic bryozoans: an ecological perspective with potential for clinical applications. Marine Environmental Research, 101, 5259.Google Scholar
Figuerola, B., Angulo-Preckler, C., Núñez-Pons, L., et al. (2017). Experimental evidence of chemical defence mechanisms in Antarctic bryozoans. Marine Environmental Research, 129, 6875.Google Scholar
Fillinger, L., Janussen, D., Lundälv, T., Richter, C. (2013). Rapid glass sponge expansion after climate-induced Antarctic ice shelf collapse. Current Biology, 23, 13301334.Google Scholar
Ford, J., Capon, R.J. (2000). Discorhabdin R: a new antibacterial pyrroloiminoquinone from two latrunculiid marine sponges, Latrunculia sp., Negombata sp. Journal of Natural Products, 63, 15271528.Google Scholar
Fries, J.L. (2016). Chemical investigation of Antarctic marine organisms and their role in modern drug discovery. University of South Florida.Google Scholar
Furrow, F.B., Amsler, C.D., McClintock, J.B., Baker, B.J. (2003). Surface sequestration of chemical feeding deterrents in the Antarctic sponge Latrunculia apicalis as an optimal defence against sea star spongivory. Marine Biology, 143, 443449.Google Scholar
Gavagnin, M., Carbone, M., Mollo, E., Cimino, G. (2003). Austrodoral and austrodoric acid: nor-sesquiterpenes with a new carbon skeleton from the Antarctic nudibranch Austrodoris kerguelenensis. Tetrahedron Letters, 44, 14951498.Google Scholar
Griffiths, H.J., Meijers, A., Bracegirdle, T. (2017). More losers than winners in a century of future Southern Ocean seafloor warming. Nature Climate Change, 7, 749754.Google Scholar
Gutt, J., Barratt, I., Domack, E., et al. (2011). Biodiversity change after climate-induced iceshelf collapse in the Antarctic. Deep-Sea Research Pt. II, 58, 7483.Google Scholar
Harper, M.K., Bugni, T.S., Copp, B.R., et al. (2001). Introduction to the chemical ecology of marine natural products. In: McClintock, J.B. and Baker, B.J. (eds) Marine Chemical Ecology. CRC Press, Boca Raton, FL, pp. 369.Google Scholar
Huang, Y.M., McClintock, J.B., Amsler, C.D., Peters, K.J., Baker, B.J. (2006). Feeding rates of common Antarctic gammarid amphipod on ecologically important sympatric macroalgae. Journal of Experimental Marine Biology and Ecology, 329, 5565.Google Scholar
Iken, K.B., Baker, B.J. (2003). Ainigmaptilones, sesquiterpenes from the Antarctic gorgonian coral Ainigmaptilon antarcticus. Journal of Natural Products, 66, 888890.Google Scholar
Iken, K., Avila, C., Ciavatta, M.L., Fontana, A., Cimino, G. (1998). Hodgsonal, a new drimane sesquiterpene from the mantle of the Antarctic nudibranch Bathydoris hodgsoni. Tetrahedron Letters, 39, 56355638.Google Scholar
Iken, K., Avila, C., Fontana, A., Gavagnin, M. (2002). Chemical ecology and origin of defensive compounds in the Antarctic nudibranch Austrodoris kerguelenensis (Opisthobranchia: Gastropoda). Marine Biology, 141, 101109.Google Scholar
IPCC Core Writing Team (2014). In: Pachauri, R.K, Meyer, L.A (eds) Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. IPCC, Geneva.Google Scholar
Ivanchina, N.V., Kicha, A.A., Kalinovsky, A.I., et al. (2006). Polar steroidal compounds from the Far Eastern starfish Henricia leviuscula. Journal of Natural Products, 69, 224228.Google Scholar
Ivanchina, N.V., Kicha, A.A., Stonik, V.A. (2011). Steroid glycosides from marine organisms. Steroids, 76, 425454.Google Scholar
Jacob, U., Terpstra, S., Brey, T. (2003). High-Antarctic regular sea urchins – the role of depth and feeding in niche separation. Polar Biology, 26, 99104.Google Scholar
Jouiaei, M., Yanagihara, A., Madio, B., et al. (2015). Ancient venom systems: a review on cnidaria toxins. Toxins, 7, 22512271.Google Scholar
Kim, H.J., Kim, W.J., Koo, B-W., et al. (2017). Anticancer activity of sulfated polysaccharides isolated from the Antarctic red seaweed Iridaea cordata. Ocean and Polar Research, 38, 129137.Google Scholar
Koplovitz, G., McClintock, J.B., Amsler, C.D., Baker, B.J. (2009). Palatability and chemical anti-predatory defenses in common ascidians from the Antarctic Peninsula. Aquatic Biology, 7, 8192.Google Scholar
Lambert, G. (2005). Ecology and natural history of the protochordates. Canadian Journal of Zoology, 83, 3450.Google Scholar
Laturnus, F., Wiencke, C., Klöser, H. (1996). Antarctic macroalgae – sources of volatile halogenated organic compounds. Marine Environmental Research, 41, 169181.Google Scholar
Lebar, M.D., Baker, B.J. (2010). Synthesis and structure reassessment of Psammopemmin A. Australian Journal of Chemistry, 63, 862866.Google Scholar
Lebar, M.D., Heimbegner, J.L., Baker, B.J. (2007). Cold-water marine natural products. Natural Product Report, 24, 774797.Google Scholar
Lee, H., Ahn, J., Lee, Y., Rho, J., Shin, J. (2004). New sesterterpenes from the Antarctic sponge Suberites sp. Journal of Natural Products, 67, 672674.Google Scholar
Li, F., Janussen, D., Peifer, C., Pérez-Victoria, I., Tasdemir, D. (2018). Targeted isolation of Tsitsikammamines from the Antarctic deep-sea sponge Latrunculia biformis by molecular networking and anticancer activity. Marine Drugs, 16, 268.Google Scholar
Ma, W.S., Mutka, T., Vesley, B., et al. (2009). Norselic acids A−E, highly oxidized anti-infective steroids that deter mesograzer predation, from the Antarctic sponge Crella sp. Journal of Natural Products, 72, 18421846.Google Scholar
Mahon, A.R., Amsler, C.D., McClintock, J.B., Amsler, M.O., Baker, B.J. (2003). Tissue-specific palatability and chemical defences against macropredators and pathogens in the common articulate brachiopod Liothyrella uva from the Antarctic Peninsula. Journal of Experimental Marine Biology and Ecology, 290, 197210.Google Scholar
Maier, M.S., Araya, E., Seldes, A.M. (2000). Sulphated polyhydroxysteroids from the Antarctic ophiuroid Gorgonocephalus chilensis. Molecules, 5, 348349.Google Scholar
Maier, M.S., Roccatagliata, A.J., Kuriss, A., et al. (2001). Two new cytotoxic and virucidal trisulphated triterpene glycosides from the Antarctic sea cucumber Staurocucumis liouvillei. Journal of Natural Products, 64, 732736.Google Scholar
Manzo, E., Ciavatta, M.L., Nuzzo, G., Gavagnin, M. (2009). Terpenoid content of the Antarctic soft coral Alcyonium antarcticum. Natural Product Communications, 4, 16151619.Google Scholar
Maschek, J.A., Baker, B.J. (2008). The chemistry of algal secondary metabolism. In: Amsler, C.D. (ed.) Algal Chemical Ecology. Berlin,Springer, pp. 120.Google Scholar
Maschek, J.A., Mevers, E., Diyabalanage, T., et al. (2012). Palmadorine chemodiversity from the Antarctic nudibranch Austrodoris kerguelenensis and inhibition of Jak2-STAT5-dependent HEL leukemia cells. Tetrahedron, 68, 90959104.Google Scholar
McClintock, J.B. (1987). Investigation of the relationship between invertebrate predation and biochemical composition, energy content, spicule armament and toxicity of benthic sponges at McMurdo Sound, Antarctica. Marine Biology, 94, 479487.Google Scholar
McClintock, J.B., Baker, B.J. (1997). Palatability and chemical defense of eggs, embryos and larvae of shallow water Antarctic marine invertebrates. Marine Ecology Progress Series, 154, 121131.Google Scholar
McClintock, J.B., Karentz, D. (1997). Mycosporine-like amino acids in 38 species of subtidal marine organisms from McMurdo Sound. Antarctica. Antarctic Science, 9, 392398.Google Scholar
McClintock, J.B., Heine, J., Slattery, M., Weston, J. (1991). Biochemical and energetic composition, population biology, and chemical defense of the Antarctic ascidian Cnemidocarpa verrucosa lesson. Journal of Experimental Marine Biology and Ecology, 147, 163175.Google Scholar
McClintock, J.B., Slattery, M., Thayer, C.W. (1993). Energy content and chemical defense of the articulate Brachiopod Liothyrella uva (Jackson, 1912) from the Antarctic Peninsula. Journal of Experimental Marine Biology and Ecology, 169, 103116.Google Scholar
McClintock, J.B., Amsler, M.O., Amsler, C.D., et al. (2004). Biochemical composition, energy content and chemical antifeedant and antifoulant defenses of the colonial Antarctic ascidian Distaplia cylindrica. Marine Biology, 145, 885894.Google Scholar
McClintock, J.B., Amsler, C.D., Baker, B.J. (2010). Overview of the chemical ecology of benthic marine invertebrates along the western Antarctic Peninsula. Integrative and Comparative Biology, 50, 967980.Google Scholar
McGovern, T.M., Hellberg, M.E. (2003). Cryptic species, cryptic endosymbionts, and geographic variation in chemical defenses in the bryozoan Bugula neritina. Molecular Ecology, 12, 12071215.Google Scholar
Mellado, G.G., Zubía, E., Ortega, M.J., López-González, P.J. (2005). Steroids from the Antarctic octocoral Anthomastus bathyproctus. Journal of Natural Products, 68, 11111115.Google Scholar
Miyata, Y., Diyabalanage, T., Amsler, C.D., et al. (2007). Ecdysteroids from the Antarctic tunicate Synoicum adareanum. Journal of Natural Products, 70, 18591864.Google Scholar
Moles, J., Núñez-Pons, L., Taboada, S., et al. (2015). Anti-predatory chemical defences in Antarctic benthic fauna. Marine Biology, 162, 18131821.Google Scholar
Moles, J., Wägele, H., Cutignano, A., Fontana, A., Avila, C. (2016). Distribution of granuloside in the Antarctic nudibranch Charcotia granulosa (Gastropoda: Heterobranchia: Charcotiidae). Marine Biology, 163, 5465.Google Scholar
Moles, J., Wägele, H., Cutignano, A., et al. (2017). Giant embryos and hatchlings of Antarctic nudibranchs (Mollusca: Gastropoda: Heterobranchia). Marine Biology, 164, 114126.Google Scholar
Moon, B., Park, Y.C., McClintock, J.B., Baker, B.J. (2000). Structure and bioactivity of erebusinone, a pigment from the Antarctic sponge Isodictya erinacea. Tetrahedron, 56, 90579062.Google Scholar
Morris, B.D., Prinsep, M.R. (1999). Amathaspiramides A-F, novel brominated alkaloids from the marine bryozoan Amathia wilsoni. Journal of Natural Products, 62, 688693.Google Scholar
Noguez, J.H., Diyabalanage, T.K.K., Miyata, Y., et al. (2011). Palmerolide Macrolides from the Antarctic tunicate Synoicum adareanum. Bioorganic & Medicinal Chemistry, 19, 66086614.Google Scholar
Núñez-Pons, L., Avila, C. (2014a). Defensive metabolites from Antarctic invertebrates: does energetic content interfere with feeding repellence? Marine Drugs, 12, 37703791.Google Scholar
Núñez-Pons, L., Avila, C. (2014b). Deterrent activities in the crude lipophilic fractions of Antarctic benthic organisms: chemical defences against keystone predators. Polar Research, 33, 21624.Google Scholar
Núñez-Pons, L., Avila, C. (2015). Natural products mediating ecological interactions in Antarctic benthic communities: a mini-review of the known molecules. Natural Product Report, 32, 11141130.Google Scholar
Núñez-Pons, L., Forestieri, R., Nieto, R.M., et al. (2010). Chemical defenses of tunicates of the genus Aplidium from the Weddell Sea (Antarctica). Polar Biology, 33, 13191329.Google Scholar
Núñez-Pons, L., Carbone, M., Paris, D., et al. (2012a). Chemoecological studies on hexactinellid sponges from the Southern Ocean. Naturwissenschaften, 99, 353368.Google Scholar
Núñez-Pons, L., Carbone, M., Vázquez, J., et al. (2012b). Natural products from Antarctic colonial ascidians of the genera Aplidium and Synoicum: variability and defensive role. Marine Drugs, 10, 17411764.Google Scholar
Núñez-Pons, L., Rodríguez-Arias, M., Gómez-Garreta, A., Ribera-Siguán, A., Avila, C. (2012c). Feeding deterrency in Antarctic marine organisms: bioassays with the omnivore amphipod Cheirimedon femoratus. Marine Ecology Progress Series, 462, 163174.Google Scholar
Núñez-Pons, L., Carbone, M., Vázquez, J., Gavagnin, M., Avila, C. (2013). Lipophilic defenses from Alcyonium soft corals of Antarctica. Journal of Chemical Ecology, 39, 675685.Google Scholar
Núñez-Pons, L., Nieto, R.M., Avila, C., Jiménez, C., Rodríguez, J. (2015). Mass spectrometry detection of minor new meridianins from the Antarctic colonial ascidians Aplidium falklandicum and Aplidium meridianum. Journal of Mass Spectrometry, 50, 103111.Google Scholar
Núñez-Pons, L., Avila, C., Romano, G., Verde, C., Giordano, D. (2018). UV-protective compounds in marine organisms from the Southern Ocean. Marine Drugs, 16, 336.Google Scholar
Palermo, J.A., Rodrı, M.F., Spagnuolo, C., Seldes, A.M. (2000). Illudalane sesquiterpenoids from the soft coral Alcyonium paessleri: the first natural nitrate esters. Journal of Organic Chemistry, 65, 44824486.Google Scholar
Papaleo, M.C., Fondi, M., Maida, I., et al. (2012). Sponge-associated microbial Antarctic communities exhibiting antimicrobial activity against Burkholderia cepacia complex bacteria. Biotechnology Advances, 30, 272293.Google Scholar
Pasotti, F., Manini, E., Giovannelli, D., et al. (2015). Antarctic shallow water benthos in an area of recent rapid glacier retreat. Marine Ecology, 36, 716733.Google Scholar
Patiño Cano, L.P., Manfredi, R.Q., Pérez, M., et al. (2018). Isolation and antifouling activity of azulene derivatives from the Antarctic gorgonian Acanthogorgia laxa. Chemistry & Biodiversity, 15, e1700425.Google Scholar
Paul, V.J. (1992). Ecological Roles of Marine Natural Products. Cornell University Press, Ithaca, NY.Google Scholar
Pawlik, J.R. (2012). Antipredatory defensive roles of natural products from marine invertebrates. In: Fattorusso, E, Gerwick, W.H., Taglialatela-Scafati, O (eds) Handbook of Marine Natural Products. Springer Netherlands, Dordrecht, pp. 677710.Google Scholar
Peck, L.S. (2018). Antarctic marine biodiversity: adaptations, environments and responses to change. Oceanography and Marine Biology, 56, 105236.Google Scholar
Peters, K.J., Amsler, C.D., McClintock, J.B., van Soest, R.W.M., Baker, B.J. (2009). Palatability and chemical defenses of sponges from the western Antarctic Peninsula. Marine Ecology Progress Series, 385, 7785.Google Scholar
Peters, L., Wright, A.D., Krick, A., König, G.M. (2004). Variation of brominated indoles and terpenoids within single and different colonies of the marine bryozoan Flustra foliacea. Journal of Chemical Ecology, 30, 11651182.Google Scholar
Poloczanska, E.S., Burrows, M.T., Brown, C.J., et al. (2016). Responses of marine organisms to climate change across oceans. Frontiers in Marine Science, 3, 121.Google Scholar
Principe, P.P., Fisher, W.S. (2018). Spatial distribution of collections yielding marine natural products. Journal of Natural Products, 81, 23072320.Google Scholar
Puglisi, M.P., Becerro, M.A. (2018). Life in Extreme Environments: Insights in Biological Capability. CRC Press, Boca Raton, FL.Google Scholar
Puglisi, M.P., Sneed, J.M., Ritson-Williams, R., Young, R. (2019). Marine chemical ecology in benthic environments. Natural Product Report, 36(3), 410–429.Google Scholar
Reyes, F., Fernandez, R., Rodriguez, A., et al. (2008). Aplicyanins A-F, new cytotoxic bromoindole derivatives from the marine tunicate Aplidium cyaneum. Tetrahedron, 64, 51195123.Google Scholar
Rhimou, B., Hassane, R., Nathalie, B., Coppens, Y., Vannes, U.D.B. (2010). Antiviral activity of the extracts of Rhodophyceae from Morocco. African Journal of Biotechnology, 9, 79687975.Google Scholar
Rivera, P. (1996). Plastoquinones and a chromene isolated from the Antarctic brown alga Desmarestia menziesii. Boletín de la Sociedad Chilena de Química, 41, 103105.Google Scholar
Rodríguez Brasco, M.F., Seldes, A.M., Palermo, J.A. (2001). Paesslerins A and B: novel tricyclic sesquiterpenoids from the soft coral Alcyonium paessleri. Organic Letters, 3, 14151417.Google Scholar
Schnitzler, I., Pohnert, G., Hay, M., Boland, W. (2001). Chemical defense of brown algae (Dictyopteris spp.) against the herbivorous amphipod Ampithoe longimana. Oecologia, 126, 515521.Google Scholar
Schoenrock, K.M., Schram, J.B., Amsler, C.D., McClintock, J.B., Angus, R.A. (2015). Climate change impacts on overstory Desmarestia spp. from the western Antarctic Peninsula. Marine Biology, 162, 377389.Google Scholar
Seldes, A.M., Brasco, M.F.R., Franco, L.H., et al. (2007). Identification of two meridianins from the crude extract of the tunicate Aplidium meridianum by tandem mass spectrometry. Natural Product Research, 21, 555563.Google Scholar
Sharp, J.H., Winson, M.K., Porter, J.S. (2007). Bryozoan metabolites: an ecological perspective. Natural Product Report, 24, 659673.Google Scholar
Silchenko, A.S., Kalinovsky, A.I., Avilov, S.A., et al. (2013). Triterpene glycosides from Antarctic sea cucumbers IV. Turquetoside A, a 3-O-methylquinovose containing disulfated tetraoside from the sea cucumber Staurocucumis turqueti (Vaney, 1906) (= Cucumaria spatha). Biochemical Systematics and Ecology, 51, 4549.Google Scholar
Slattery, M., McClintock, J.B. (1995). Population structure and feeding deterrence in three shallow-water Antarctic soft corals. Marine Biology, 122, 461470.Google Scholar
Solanki, H., Angulo-Preckler, C., Calabro, K., et al. (2018). Suberitane sesterterpenoids from the Antarctic sponge Phorbas areolatus (Thiele, 1905). Tetrahedron Letters, 59, 33533356.Google Scholar
Soldatou, S., Baker, B.J. (2017). Cold-water marine natural products, 2006 to 2016. Natural Product Report, 34, 585626.Google Scholar
Stoecker, D. (1980). Chemical defenses of ascidians against predators. Ecology, 61, 13271334.Google Scholar
Swearingen, III, D.C., Pawlik, J.R. (1998). Variability in the chemical defense of the sponge Chondrilla nucula against predatory reef fishes. Marine Biology, 131, 619627.Google Scholar
Taboada, S., Núñez-Pons, L., Avila, C. (2013). Feeding repellence of Antarctic and sub-Antarctic benthic invertebrates against the omnivorous sea star Odontaster validus. Polar Biology, 36, 1325.Google Scholar
Tian, Y., Li, Y., Zhao, F. (2017). Secondary metabolites from polar organisms. Marine Drugs, 15, 28.Google Scholar
Torssel, K.B.G. (1983). Natural Product Chemistry. A Mechanistic and Biosynthetic Approach to Secondary Metabolism. John Wiley, New York.Google Scholar
Tremblay, N., Abele, D. (2016). Response of three krill species to hypoxia and warming: an experimental approach to oxygen minimum zones expansion in coastal ecosystems. Marine Ecology, 37, 179199.Google Scholar
Turner, J., Bindschadler, R., Convey, P., et al. (2009). Antarctic Climate Change and the Environment: A Contribution to the International Polar Year 2007–2008. Scientific Committee on Antarctic Research,Cambridge, UK.Google Scholar
Turner, J., Barrand, N.E., Bracegirdle, T.J., et al. (2014). Antarctic climate change and the environment: an update. Polar Record, 50, 237259.Google Scholar
Turon, X., Becerro, M.A., Uriz, M.J. (1996). Seasonal patterns of toxicity in benthic invertebrates: the encrusting sponge Crambe crambe (Poecilosclerida). Oikos, 75, 3340.Google Scholar
Vankayala, S.L., Kearns, F.L., Baker, B.J., Larkin, J.D., Woodcock, H.L. (2017). Elucidating a chemical defense mechanism of Antarctic sponges: a computational study. Journal of Molecular Graphics and Modelling, 71, 104115.Google Scholar
Vetter, W., Janussen, D. (2005). Halogenated natural products in five species of Antarctic sponges: compounds with POP-like properties? Environmental Science and Technology, 39, 38893895.Google Scholar
von Salm, J.L., Wilson, N.G., Vesely, B.A., et al. (2014). Shagenes A and B, new tricyclic sesquiterpenes produced by an undescribed Antarctic octocoral. Organic Letters, 16, 26302633.Google Scholar
von Salm, J.L., Witowski, C.G., Fleeman, R.M., et al. (2016). Darwinolide, a new diterpene scaffold that inhibits methicillin-resistant Staphylococcus aureus biofilm from the Antarctic sponge Dendrilla membranosa. Organic Letters, 18, 25962599.Google Scholar
von Salm, J.L., Schoenrock, K.M., McClintock, J.B., Amsler, C.D., Baker, B.J. (2018). The status of marine chemical ecology in Antarctica. Form and function of unique high-latitude chemistry. In: Puglisi, M.P., Becerro, M.A. (eds) Life in Extreme Environments: Insights in Biological Capability. CRC Press, Boca Raton, FL, pp. 2769.Google Scholar
Wang, M., Tietjen, I., Chen, M., et al. (2016). Sesterterpenoids isolated from the sponge Phorbas sp. activate latent HIV-1 provirus expression. Journal of Organic Chemistry, 81, 1132411334.Google Scholar
Wang, Y.J. (2014). The future of marine invertebrates in face of global climate change. Journal of Coastal Development, 17, e105.Google Scholar
Wiencke, C., Clayton, M.N. (2002). Synopses of the Antarctic Benthos. Antarctic Seaweeds. A.R.G. Gantner Verlag KG Ruggell,Liechtenstein.Google Scholar
Wiencke, C., Amsler, C.D., Clayton, M.N. (2014). Macroalgae. In: De Broyer, C, Koubbi, P, Griffiths, H. J., et al. (eds) Biogeographic Atlas of the Southern Ocean. Scientific Committee on Antarctic Research, Cambridge, UK.Google Scholar
Wilkins, S.P., Blum, A.J., Burkepile, D.E., et al. (2002). Isolation of an antifreeze peptide from the Antarctic sponge Homaxinella balfourensis. Cellular and Molecular Life Sciences, 59, 22102215.Google Scholar
Wilson, N.G., Maschek, J.A., Baker, B.J. (2013). A species flock driven by predation? Secondary metabolites support diversification of slugs in Antarctica. PLoS One, 8, e80277.Google Scholar
Winston, J.E. (2010). Life in the colonies: learning the alien ways of colonial organisms. Integrative and Comparative Biology, 50, 919933.Google Scholar
Witowski, C.W. (2015). Investigation of bioactive metabolites from the Antarctic sponge Dendrilla membranosa and marine microorganisms. PhD thesis, University of South Florida.Google Scholar
Young, E.B., Dring, M.J., Savidge, G., Birkett, D.A., Berges, J.A. (2007). Seasonal variations in nitrate reductase activity and internal N pools in intertidal brown algae are correlated with ambient nitrate concentrations. Plant Cell & Environment, 30, 764774.Google Scholar
Young, R.M., von Salm, J.L., Amsler, M.O., et al. (2013). Site-specific variability in the chemical diversity of the Antarctic red alga Plocamium cartilagineum. Marine Drugs, 11, 21262139.Google Scholar
Young, R.M., Schoenrock, K.M., von Salm, J.L., Amsler, C.D., Baker, B.J. (2015). Structure and function of macroalgal natural products. In:Stengel, D and Connan, S (eds) Natural Products from Marine algae. Methods in Molecular Biology. Humana Press, New York, pp. 3973.Google Scholar

References

Achberger, A.M. (2016). Structure and Functional Potential of Microbial Communities in Subglacial Lake Whillans and at the Ross Ice Shelf Grounding Zone, West Antarctica. PhD thesis, Louisiana State University.Google Scholar
Achberger, A.M., Christner, B.C., Michaud, A.B., et al. (2016). Microbial community structure of subglacial Lake Whillans, West Antarctica. Frontiers in Microbiology, 7, 256–213. doi:10.3389/fmicb.2016.01457Google Scholar
Achberger, A.M., Michaud, A.B., Vick-Majors, T.J., et al. (2017). Microbiology of subglacial environments. In: Margesin, R. (ed.) Psychrophiles: From Biodiversity to Biotechnology. Springer International Publishing, Cham, Switzerland, pp. 83110.Google Scholar
Bakermans, C., Skidmore, M. (2011). Microbial respiration in ice at subzero temperatures (−4°C to −33°C). Environmental Microbiology Reports, 3, 774782. doi:10.1111/j.1758-2229.2011.00298.xGoogle Scholar
Beulig, F., Røy, H., Glombitza, C., Jørgensen, B.B. (2018). Control on rate and pathway of anaerobic organic carbon degradation in the seabed. Proceedings of the National Academy of Sciences of the USA, 115, 367372. doi:10.1073/pnas.1715789115Google Scholar
Bottrell, S.H., Tranter, M. (2002). Sulphide oxidation under partially anoxic conditions at the bed of the Haut Glacier d’Arolla, Switzerland. Hydrological Processes, 16, 23632368. doi:10.1002/hyp.1012Google Scholar
Boyd, E.S., Lange, R.K., Mitchell, A.C., et al. (2011). Diversity, abundance, and potential activity of nitrifying and nitrate-reducing microbial assemblages in a subglacial ecosystem. Applied and Environmental Microbiology, 77, 47784787. doi:10.1128/AEM.00376-11Google Scholar
Bulat, S.A. (2016). Microbiology of the subglacial Lake Vostok: first results of borehole-frozen lake water analysis and prospects for searching for lake inhabitants. Philosophical Transactions of the Royal Society A, 374, 20140292. doi:10.1098/rsta.2014.0292Google Scholar
Carter, S.P., Fricker, H.A. (2012). The supply of subglacial meltwater to the grounding line of the Siple Coast, West Antarctica. Annals of Glaciology, 53, 267280. doi:10.3189/2012aog60a119Google Scholar
Christner, B.C., Mosley-Thompson, E., Thompson, L.G., Reeve, J.N. (2001). Isolation of bacteria and 16S rDNAs from Lake Vostok accretion ice. Environmental Microbiology, 3, 570577. doi:10.1046/j.1462-2920.2001.00226.xGoogle Scholar
Christner, B.C., Royston-Bishop, G., Foreman, C.M., et al. (2006). Limnological conditions in Subglacial Lake Vostok, Antarctica. Limnology and Oceanography, 51, 24852501.Google Scholar
Christner, B.C., Priscu, J.C., Achberger, A.M., et al. (2014). A microbial ecosystem beneath the West Antarctic ice sheet. Nature, 512, 310313. doi:10.1038/nature13667Google Scholar
Christoffersen, P., Bougamont, M. (2014). Significant groundwater contribution to Antarctic ice streams hydrologic budget. Geophysical Research Letters, 41, 2003–2010. doi:10.1002/2014gl059250Google Scholar
Dieser, M., Hagedorn, B., Christner, B.C., et al. (2014). Molecular and biogeochemical evidence for methane cycling beneath the western margin of the Greenland Ice Sheet. ISME Journal, 8, 23052316. doi:10.1038/ismej.2014.59Google Scholar
Dowdeswell, J.A., Siegert, M.J. (1999). The dimensions and topographic setting of Antarctic subglacial lakes and implications for large-scale water storage beneath continental ice sheets. Geological Society of America Bulletin, 111, 254263.Google Scholar
Doyle, S., Montross, S., Skidmore, M., Christner, B. (2013). Characterizing microbial diversity and the potential for metabolic function at −15°C in the basal ice of Taylor Glacier, Antarctica. Biology, 2, 10341053. doi:10.3390/biology2031034Google Scholar
Fisher, A.T., Mankoff, K.D., Tulaczyk, S.M., Tyler, S.W., Foley, N.; WISSARD Science Team (2015). High geothermal heat flux measured below the West Antarctic Ice Sheet. Science Advances, 1, e1500093. doi:10.1126/sciadv.1500093Google Scholar
Fricker, H.A., Scambos, T., Bindschadler, R., Padman, L. (2007). An active subglacial water system in West Antarctica mapped from space. Science, 315, 15441548. doi:10.1126/science.1136897Google Scholar
Fricker, H.A., Siegfried, M.R., Carter, S.P., Scambos, T.A. (2015). A decade of progress in observing and modelling Antarctic subglacial water systems. Philosophical Transactions of the Royal Society A, 374, 20140294. doi:10.1098/rsta.2014.0294Google Scholar
Gaidos, E., Marteinsson, V., Thorsteinsson, T., et al. (2009). An oligarchic microbial assemblage in the anoxic bottom waters of a volcanic subglacial lake. ISME Journal, 3, 486497. doi:10.1038/ismej.2008.124Google Scholar
Gray, L., Joughin, I., Tulaczyk, S., Spikes, V.B. (2005). Evidence for subglacial water transport in the West Antarctic Ice Sheet through three‐dimensional satellite radar interferometry. Geophysical Research Letters, 32, L03501.Google Scholar
Harrold, Z.R., Skidmore, M.L., Hamilton, T.L., et al. (2015). Aerobic and anaerobic thiosulfate oxidation by a cold-adapted, subglacial chemoautotroph. Applied and Environmental Microbiology, 82, 14861495. doi:10.1128/AEM.03398-15Google Scholar
Hawkings, J.R., Wadham, J.L., Benning, L.G., et al. (2017). Ice sheets as a missing source of silica to the polar oceans. Nature Communications, 8, 14198. doi:10.1038/ncomms14198Google Scholar
Hell, K., Insam, H., Edwards, A., et al. (2013). The dynamic bacterial communities of a melting High Arctic glacier snowpack. ISME Journal, 7, 1814–1826. doi:10.1038/ismej.2013.51Google Scholar
Hodson, A., Brock, B., Pearce, D., Laybourn-Parry, J., Tranter, M. (2015). Cryospheric ecosystems: a synthesis of snowpack and glacial research. Environmental Research Letters, 10, 110201–9. doi:10.1088/1748-9326/10/11/110201Google Scholar
Jouzel, J., Petit, J.R., Souchez, R., et al. (1999). More than 200 meters of lake ice above subglacial Lake Vostok, Antarctica. Science, 286, 21382141.Google Scholar
Jørgensen, B.B. (1982). Mineralization of organic matter in the sea bed – the role of sulphate reduction. Nature, 296, 643645. doi:10.1038/296643a0Google Scholar
Jørgensen, B.B. (2011). Deep subseafloor microbial cells on physiological standby. Proceedings of the National Academy of Sciences of the USA, 108, 1819318194. doi:10.1073/pnas.1115421108Google Scholar
Kapitsa, A.P., Ridley, J.K., de Q Robin, G., Siegert, M.J., Zotikov, I.A. (1996). A large deep freshwater lake beneath the ice of central East Antarctica. Nature, 381, 684686. doi:10.1038/381684a0Google Scholar
Karl, D.M., Bird, D.F., Bjorkman, K., et al. (1999). Microorganisms in the accreted ice of Lake Vostok, Antarctica. Science, 286, 21442147. doi:10.1126/science.286.5447.2144Google Scholar
Lanoil, B., Skidmore, M., Priscu, J.C., et al. (2009). Bacteria beneath the West Antarctic Ice Sheet. Environmental Microbiology, 11, 609615. doi:10.1111/j.1462-2920.2008.01831.xGoogle Scholar
LaRowe, D., Amend, J. (2014). Energetic constraints on life in marine deep sediments. In: Kallmeyer, J., Wagner, D. (eds) Microbial Life of the Deep Biosphere. De Gruyter, Berlin, Boston, pp. 279302.Google Scholar
Lin, L.-H., Wang, P.-L., Rumble, D., et al. (2006). Long-term sustainability of a high-energy, low-diversity crustal biome. Science, 314, 479482. doi:10.1126/science.1127376Google Scholar
Lipps, J.H., Ronan, T.E., DeLaca, T.E. (1979). Life below the Ross Ice Shelf, Antarctica. Science, 203, 447449. doi:10.1126/science.203.4379.447Google Scholar
Llubes, M., Lanseau, C., Rémy, F. (2006). Relations between basal condition, subglacial hydrological networks and geothermal flux in Antarctica. Earth and Planetary Science Letters, 241, 655662. doi:10.1016/j.epsl.2005.10.040Google Scholar
Lollar, B.S., Onstott, T.C., Lacrampe-Couloume, G., Ballentine, C.J. (2014). The contribution of the Precambrian continental lithosphere to global H2 production. Nature, 516, 379382. doi:10.1038/nature14017Google Scholar
McKay, C.P., Hand, K.P., Doran, P.T., Andersen, D.T., Priscu, J.C. (2003). Clathrate formation and the fate of noble and biologically useful gases in Lake Vostok, Antarctica. Geophysical Research Letters, 30, 124. doi:10.1029/2003GL017490Google Scholar
Michaud, A.B., Skidmore, M.L., Mitchell, A.C., et al. (2016). Solute sources and geochemical processes in Subglacial Lake Whillans. West Antarctica, 44, 347350. doi:10.1130/G37639.1Google Scholar
Michaud, A.B., Dore, J.E., Achberger, A.M., et al. (2017). Microbial oxidation as a methane sink beneath the West Antarctic Ice Sheet. Nature Publishing Group, 10, 18. doi:10.1038/ngeo2992Google Scholar
Mikucki, J.A., Priscu, J.C. (2007). Bacterial diversity associated with Blood Falls, a subglacial outflow from the Taylor Glacier, Antarctica. Applied and Environmental Microbiology, 73, 40294039. doi:10.1128/AEM.01396-06Google Scholar
Mikucki, J.A., Foreman, C.M., Sattler, B., Lyons, W.B., Priscu, J.C. (2004). Geomicrobiology of Blood Falls: an iron-rich saline discharge at the terminus of the Taylor Glacier, Antarctica. Aquatic Geochemistry, 10, 199220. doi:10.1007/s10498-004-2259-xGoogle Scholar
Mikucki, J.A., Schrag, D.P., Mikucki, J.A., et al. (2009). A contemporary microbially maintained subglacial ferrous ‘ocean’.Science, 324, 397400. doi:10.1126/science.1167350Google Scholar
Mikucki, J.A., Lee, P.A., Ghosh, D., et al. (2016). Subglacial Lake Whillans microbial biogeochemistry: a synthesis of current knowledge. Philosophical Transactions of the Royal Society A, 374, pii: 20140290. doi:10.1098/rsta.2014.0290Google Scholar
Mitchell, A.C., Lafrenière, M.J., Skidmore, M.L., Boyd, E.S. (2013). Influence of bedrock mineral composition on microbial diversity in a subglacial environment. Geology, 41, 855858. doi:10.1130/G34194.1Google Scholar
Montross, S.N., Skidmore, M., Tranter, M., Kivimaki, A.L., Parkes, R.J. (2013). A microbial driver of chemical weathering in glaciated systems. Geology, 41, 215218. doi:10.1130/G33572.1Google Scholar
National Research Council (2007). Exploration of Antarctic Subglacial Aquatic Environments: Environmental and Scientific Stewardship. The National Academies Press, Washington, DC.Google Scholar
Oswald, G., Robin, G.Q. (1973). Lakes beneath the Antarctic ice sheet. Nature, 245, 251254. doi:10.1038/245251a0Google Scholar
Pattyn, F. (2010). Antarctic subglacial conditions inferred from a hybrid ice sheet/ice stream model. Earth and Planetary Science Letters, 295, 451461. doi:10.1016/j.epsl.2010.04.025Google Scholar
Petit, J.R., Jouzel, J., Raynaud, D., et al. (1999). Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature, 399, 429436.Google Scholar
Post, A.L., Galton-Fenzi, B.K., Riddle, M.J., et al. (2014). Modern sedimentation, circulation and life beneath the Amery Ice Shelf, East Antarctica. Continental Shelf Research, 74, 7787. doi:10.1016/j.csr.2013.10.010Google Scholar
Prestrud-Anderson, S., Drever, J.I., Humphrey, N.F. (1997). Chemical weathering in glacial environments. Geology, 25, 399.Google Scholar
Priscu, J.C., Christner, B.C. (2004). Earth’s icy biosphere. Microbial Diversity and Prospecting, 130145.Google Scholar
Priscu, J.C., Adams, E.E., Lyons, W.B., et al. (1999). Geomicrobiology of subglacial ice above Lake Vostok, Antarctica. Science, 286, 21412144.Google Scholar
Priscu, J.C., Tulaczyk, S., Studinger, M., et al. (2008). Antarctic subglacial water: origin, evolution, and ecology. In: Vincent, W.F., Laybourn-Parry, J (eds) Polar Lakes and Rivers. Oxford University Press Inc., New York, pp. 119135.Google Scholar
Priscu, J.C., Achberger, A.M., Cahoon, J.E., et al. (2013). A microbiologically clean strategy for access to the Whillans Ice Stream subglacial environment. Antarctic Science, 25, 637647. doi:10.1017/S0954102013000035Google Scholar
Purcell, A.M., Mikucki, J.A., Achberger, A.M., et al. (2014). Microbial sulfur transformations in sediments from Subglacial Lake Whillans. Frontiers in Microbiology, 5, 594. doi:10.3389/fmicb.2014.00594Google Scholar
Riddle, M.J., Craven, M., Goldsworthy, P.M., Carsey, F. (2007). A diverse benthic assemblage 100 km from open water under the Amery Ice Shelf, Antarctica. Paleoceanography, 22. doi:10.1029/2006pa001327Google Scholar
Rignot, E., Jacobs, S., Mouginot, J., Scheuchl, B. (2013). Ice-shelf melting around Antarctica. Science, 341, 266270. doi:10.1126/science.1235798Google Scholar
Robinson, N.J., Williams, M.J.M., Barrett, P.J., et al. (2010). Observations of flow and ice-ocean interaction beneath the McMurdo Ice Shelf, Antarctica. Journal of Geophysical Research, 115, C03025. doi:10.1029/2008JC005255Google Scholar
Rogers, S., Shtarkman, Y., Koçer, Z., et al. (2013). Ecology of subglacial lake Vostok (Antarctica), based on metagenomic/metatranscriptomic analyses of accretion ice. Biology, 2, 629650. doi:10.3390/biology2020629Google Scholar
Sharp, M., Parkes, J., Cragg, B., Fairchild, I.J., Lamb, H. (1999). Widespread bacterial populations at glacier beds and their relationship to rock weathering and carbon cycling. Geology, 27, 107. doi:10.1130/0091-7613(1999)0272.3.co;2Google Scholar
Siegert, M.J., Ellis-Evans, J.C., Tranter, M., et al. (2001). Physical, chemical and biological processes in Lake Vostok and other Antarctic subglacial lakes. Nature, 414, 603609. doi:10.1038/414603aGoogle Scholar
Siegert, M.J., Carter, S., Tabacco, I., Popov, S., Blankenship, D.D. (2005). A revised inventory of Antarctic subglacial lakes. Antarctic Science, 17, 453. doi:10.1017/S0954102005002889Google Scholar
Siegert, M.J., Clarke, R.J., Mowlem, M., Ross, N. (2012). Clean access, measurement, and sampling of Ellsworth Subglacial Lake: a method for exploring deep Antarctic subglacial lake environments. Review of Geophysics, 50, RG1003. doi:10.1029/2011rg000361Google Scholar
Siegert, M.J., Makinson, K., Blake, D., et al. (2014). An assessment of deep hot-water drilling as a means to undertake direct measurement and sampling of Antarctic subglacial lakes: experience and lessons learned from the Lake Ellsworth field season 2012/13. Annals of Glaciology, 55, 5973. doi:10.3189/2014AoG65A008Google Scholar
Siegert, M.J., Ross, N., Le Brocq, A.M. (2016). Recent advances in understanding Antarctic subglacial lakes and hydrology. Philosophical Transactions of The Royal Society A Mathematical Physical and Engineering Sciences, 374, 20140306. doi:10.1098/rsta.2014.0306Google Scholar
Siegfried, M.R., Fricker, H.A., Roberts, M., Scambos, T.A., Tulaczyk, S. (2014). A decade of West Antarctic subglacial lake interactions from combined ICESat and CryoSat-2 altimetry. Geophysical Research Letters, 41, 891898. doi:10.1002/2013gl058616Google Scholar
Skidmore, M. (2011). Microbial communities in Antarctic subglacial aquatic environments. In: Antarctic Subglacial Aquatic Environments. American Geophysical Union. Oxford University Press Inc., New York, pp. 6181.Google Scholar
Skidmore, M.L., Foght, J.M., Sharp, M.J. (2000). Microbial life beneath a high Arctic glacier. Applied and Environmental Microbiology, 66, 32143220. doi:10.1128/AEM.66.8.3214-3220.2000.UpdatedGoogle Scholar
Skidmore, M., Tranter, M., Tulaczyk, S., Lanoil, B. (2010). Hydrochemistry of ice stream beds – evaporitic or microbial effects? Hydrological Processes, 24, 517523. doi:10.1002/hyp.7580Google Scholar
Tranter, M. (2003). Geochemical weathering in glacial and proglacial environments. In: Singh, V.P., Singh, P, Haritashay, U.K. (eds) Treatise on Geochemistry. Elsevier, the Netherlands, pp. 189205.Google Scholar
Tranter, M., Skidmore, M., Wadham, J. (2005). Hydrological controls on microbial communities in subglacial environments. Hydrological Processes, 19, 995998. doi:10.1002/hyp.5854Google Scholar
Tulaczyk, S., Tulaczyk, S., Mikucki, J.A., et al. (2014). WISSARD at Subglacial Lake Whillans, West Antarctica: scientific operations and initial observations. Annals of Glaciology, 55, 5158. doi:10.3189/2014aog65a009Google Scholar
Uemura, T., Taniguchi, M., Shibuya, K. (2011). Submarine groundwater discharge in Lützow-Holm Bay, Antarctica. Geophysical Research Letters, 38, L08402. doi:10.1029/2010GL046394Google Scholar
Vick-Majors, T.J., Achberger, A., Santibáñez, P.A., et al. (2015). Biogeochemistry and microbial diversity in the marine cavity beneath the McMurdo Ice Shelf, Antarctica. Limnology and Oceanography, 61, 572586. doi:10.1002/lno.10234Google Scholar
Vick-Majors, T.J., Mitchell, A.C., Achberger, A.M., et al. (2016). Physiological ecology of microorganisms in subglacial Lake Whillans. Frontiers in Microbiology, 7, 116. doi:10.3389/fmicb.2016.01705Google Scholar
Wadham, J.L., Tranter, M., Skidmore, M., et al. (2010). Biogeochemical weathering under ice: Size matters. Global Biogeochemical Cycles, 24, GB3025. doi:10.1029/2009gb003688Google Scholar
Wadham, J.L., Arndt, S., Tulaczyk, S., et al. (2012). Potential methane reservoirs beneath Antarctica. Nature, 488, 633637. doi:10.1038/nature11374Google Scholar
Wingham, D.J., Siegert, M.J., Shepherd, A., Muir, A.S. (2006). Rapid discharge connects Antarctic subglacial lakes. Nature, 440, 10331036. doi:10.1038/nature04660Google Scholar
Wright, A., Siegert, M. (2012). A fourth inventory of Antarctic subglacial lakes. Antarctic Science, 24, 659664. doi:10.1017/S095410201200048XGoogle Scholar
Zotikov, I.A. (2006). The Antarctic Subglacial Lake Vostok. Springer-Verlag ,Berlin/Heidelberg/New York.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
×