Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-28T05:43:31.727Z Has data issue: false hasContentIssue false

Part IV - Life and habitability

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. 297 - 354
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

Bedau, A. (2010). An Aristotelean account of minimal chemical life. Astrobiology, 10, 10111020.Google Scholar
Benner, S.A. (2010). Defining life. Astrobiology, 10, 10211030.CrossRefGoogle ScholarPubMed
Bishop, J.L., Aglen, B.L., Pratt, L.M., et al. (2004). A spectroscopic and isotopic study of sediments from the Antarctic Dry Valleys as analogues for potential palaeolakes on Mars. International Journal of Astrobiology, 2, 273287.Google Scholar
Brier, J.A., White, S.N., German, C.R. (2010). Mineral-microbe interactions in deep-sea hydrothermal systems: a challenge for Raman spectroscopy. Philosophical Transactions of the Royal Society A, 368, 30673086.CrossRefGoogle Scholar
Carter, E.A., Hargreaves, M.D., Kee, T.P., Pasek, M. A., Edwards, H.G.M. (2010). A Raman spectroscopic study of a fulgurite. Philosophical Transactions of the Royal Society A, 368, 30873098.Google Scholar
Clark, B.C. (1998). Surviving the limits to life at the surface of Mars. Journal of Geophysical Research Planets, 103, 2854528556.Google Scholar
Cleland, C.E., Chyba, C.F. (2002). Defining life. Origins of Life and Evolution of the Biosphere, 32, 387393.Google Scholar
Cockell, C.S. (2001). Astrobiology and the ethics of new science. Interdisciplinary Science Reviews, 26, 9096.Google Scholar
Cockell, C.S., Knowland, J.R. (1999). Ultraviolet screening compounds. Biological Reviews, 74, 311345.CrossRefGoogle ScholarPubMed
Dickensheets, D.L., Wynn-Williams, D.D., Edwards, H.G.M., Crowder, C., Newton, E.M. (2000). A novel miniature confocal Raman spectrometer system for biomarker analysis on future Mars missions after Antarctic trials. Journal of Raman Spectroscopy, 31, 633635.Google Scholar
Doran, P.T., Wharton, R.A.J., des Marais, D.J., McKay, C.P. (1998). Antarctic palaeolake sediments and the search for extinct life on Mars. Journal of Geophysical Research Planets, 103, 2848128488.Google Scholar
Edwards, H.G.M. (2004). Raman spectroscopic protocol for the molecular recognition of key biomarkers in astrobiological exploration. Origins of Life and Evolution of the Biosphere, 34, 311.Google Scholar
Edwards, H.G.M. (2010). Raman spectroscopic approach to analytical astrobiology: the detection of key geological and biomolecular markers in the search for life. Philosophical Transactions of the Royal Society A, 368, 30593066.CrossRefGoogle ScholarPubMed
Edwards, H.G.M., Newton, E.M. (1999). Applications of Raman spectroscopy for exobiological prospecting. In: Hiscox, J.A. (ed.) The Search for Life on Mars. British Interplanetary Society, London, pp. 7583.Google Scholar
Edwards, H.G.M., Russell, N.C., Wynn-Williams, D.D. (1997). Fourier- transform Raman spectroscopic and scanning electron microscopic study of cryptoendolithic lichens from Antarctica. Journal of Raman Spectroscopy, 30, 685690.3.0.CO;2-X>CrossRefGoogle Scholar
Edwards, H.G.M., Moody, C.D., Jorge Villar, S.E., Wynn-Williams, D.D. (2005). Raman spectroscopic detection of key biomarkers of cyanobacteria and lichen symbiosis in extreme Antarctic habitats: evaluation for Mars lander missions. Icarus, 174, 560571.Google Scholar
Edwards, H.G.M., Sadooni, F., Vítek, P., Jehlička, J. (2010). Raman spectroscopy of the Dukhan sabkha: identification of geological and biogeological molecules in an extreme environment. Philosophical Transactions of the Royal Society A, 368, 30993108.CrossRefGoogle Scholar
Edwards, H.G.M., Hutchinson, I.B., Ingley, R. (2012). The ExoMars Raman spectrometer and the identification of biogeological spectroscopic signatures using a flight-like prototype. Analytical & Bioanalytical Chemistry, 404, 17231731.Google Scholar
Edwards, H.G.M., Hutchinson, I.B., Ingley, R., Jehlicka, J. (2014). Biomarkers and their Raman spectroscopic signatures: a spectral challenge for analytical astrobiology. Philosophical Transactions of the Royal Society A, 372, 20140193.Google Scholar
Eigenbrode, J.L., Summons, R.E., Steele, A., et al. (2018). Organic matter preserved in 3-billion-year-old Mudstones at Gale Crater, Mars. Science, 360, 10961101.Google Scholar
Ellery, A., Wynn-Williams, D.D. (2003). Why Raman spectroscopy on Mars? A case of the right tool for the right job. Astrobiology, 3, 565579.Google Scholar
Farmer, J.D., des Marais, D.J. (1999). Exploring for a record of Ancient Martian Life. Journal of Geophysical Research, 104, 2697726995.Google Scholar
Freeman, K. (1983). Cell astrobiology: shuttle separation of islet cells. Nature, 304, 575.Google Scholar
Freissinet, C., Glavin, D.P., Mahaffy, P.R., et al. and the MSL Science Team (2015). Organic molecules in the Sheepbed mudstone, Gale Crater, Mars. Journal of Geophysical Research Planets, 120, 495514.Google Scholar
Grady, M.M., Verchovsky, A.B., Wright, I.P. (2004). Magmatic carbon in Martian meteorites: attempts to constrain the carbon cycle on Mars. International Journal of Astrobiology, 3, 117124.Google Scholar
Jehlička, J., Edwards, H.G.M. (2014). Raman spectroscopy meets extremophiles on Earth and Mars: studies for the successful search for life. Philosophical Transactions of the Royal Society A, 372, 20140207.Google Scholar
Jehlička, J., Vandenabeele, P., Edwards, H.G.M., Culka, A., Capoun, T. (2010a). Raman spectra of pure biomolecules obtained using a hand-held instrument under cold, high-altitude conditions. Analytical & Bioanalytical Chemistry, 397, 27532760.CrossRefGoogle Scholar
Jehlička, J., Edwards, H.G.M., Culka, A. (2010b).Using portable Raman spectrometers for the identification of organic compounds at low temperatures and high altitudes: exobiological applications. Philosophical Transactions of the Royal Society A, 368, 31093126.Google Scholar
Jorge-Villar, S.E., Edwards, H.G.M. (2005). Raman spectroscopy in astrobiology. Analytical & Bioanalytical Chemistry, 384, 100113.Google Scholar
Jorge-Villar, S.E., Edwards, H.G.M. (2010). Raman spectroscopy of volcanic lavas and inclusions of relevance to astrobiological exploration. Philosophical Transactions of the Royal Society A, 368, 31273136.CrossRefGoogle ScholarPubMed
Lafleur, L.J. (1941). Astrobiology. Astronomical Society of the Pacific Leaflets, 143, 133.Google Scholar
Lederberg, J. (1960). Exobiology: approaches to life beyond the earth. Science, 132, 393.CrossRefGoogle ScholarPubMed
Lin, Y., Goresy, A.E., Hu, S., et al. (2014). NanoSIMS analysis of organic carbon from the Tissint Martian meteorite: evidence for the past existence of subsurface organic-bearing fluids on Mars. Meteoritics Planetary Science, 49, 22012218.Google Scholar
Marshall, C.P., Olcott Marshall, A. (2010). The potential of Raman spectroscopy for the analysis of diagenetically transformed carotenoids. Philosophical Transactions of the Royal Society A, 368, 31373144.CrossRefGoogle ScholarPubMed
Marshall, C.P., Edwards, H.G.M., Jehlicka, J. (2010). Understanding the application of Raman spectroscopy to the detection of traces of life. Astrobiology, 10, 229243.Google Scholar
Martinez-Frias, J., Hochberg, D. (2007). Classifying science and technology: two problems with the UNESCO system. Interdisciplinary Science Reviews, 32, 315319.Google Scholar
McKay, C.P. (1997). The search for life on Mars. Origins of Life and Evolution of the Biosphere, 27, 263289.Google Scholar
Parnell, J., Cullen, D., Sims, M.R., et al. (2007). Searching for life on Mars: selection of molecular targets for ESA’s Aurora ExoMars mission. Astrobiology, 7, 578604.Google Scholar
Perry, R.S., McLoughlin, N., Lynne, B.Y., et al. (2007). Defining biominerals and organominerals: direct and indirect indicators of life. Sedimentary Geology, 201, 157179.Google Scholar
Pullan, D., Hofmann, B.A., Westall, F., et al. (2008). Identification of morphological biosignatures in Martian analogue field specimens using in situ planetary instrumentation. Astrobiology, 8, 119156.CrossRefGoogle ScholarPubMed
Rull, F., Delgado, A., Martinez-Frias, J. (2010 a) Micro-Raman spectroscopic study of extremely large atmospheric ice conglomerations (megacryometeors). Philosophical Transactions of the Royal Society A, 368, 31453152.Google Scholar
Rull, F., Munoz-Espadas, M.J., Lunar., R., Martinez-Frias, J. (2010 b). Raman spectroscopic study of four Spanish shocked ordinary chondrites: Canellas, Olmedilla de Alarcon, Reliegos and Olivenza. Philosophical Transactions of the Royal Society A, 368, 31533166.CrossRefGoogle Scholar
Shapiro, R., Schulze-Makuch, D. (2009). The search for alien life in our solar system: strategies and priorities. Astrobiology, 9, 19.Google Scholar
Sharma, S.K., Misra, A.K., Clegg, S.M., et al. (2010). Time-resolved remote Raman study of minerals under supercritical CO2 and high temperatures relevant to Venus exploration. Philosophical Transactions of the Royal Society A, 368, 31673192.Google Scholar
Soffen, G.A. (1999). Astrobiology. Life Sciences: ExoBiology, 23, 283288.Google Scholar
Steele, A., McCubbin, F.M., Fries, M.D., et al. (2012). Graphite in the Martian meteorite Allan Hills 84001. American Mineralogist, 97, 12561259.Google Scholar
Strughold, H. (1953). The Red and Green Planet: A Physiological Study of the Possibility of Life on Mars. Albuquerque: University of New Mexico Press.Google Scholar
Summons, R., Amend, J.P., Bish, D., et al. (2011). Preservation of Martian organic and environmental records. Astrobiology, 11, 157181.Google Scholar
Tikhov, G.A. (1953). Astrobiology. Molodaya Gvardia Publishing House, Moscow.Google Scholar
Tirard, S., Morange, M., Lazcano, A. (2010). The definition of life: a brief history of elusive scientific endeavour. Astrobiology, 10, 10031009.Google Scholar
Treado, P.J., Truman, A. (1996). Laser Raman spectroscopy. In: Wdowiak, T.J., Agresti, D.G (eds) Point Clear Exobiology Instrumentation Workshop. University of Alabama Press, Birmingham, AL, pp. 710.Google Scholar
Varnali, T., Edwards, H.G.M. (2010). Ab initio calculations of scytonemin derivatives of relevance to extremophile characterisation by Raman spectroscopy. Philosophical Transactions of the Royal Society A, 368, 31933204.Google Scholar
Vítek, P., Edwards, H.G.M., Jehlicka, J., et al. (2010). Microbial colonisation of halite from the hyper-arid Atacama Desert studied by Raman spectroscopy. Philosophical Transactions of the Royal Society A, 368, 32053221.CrossRefGoogle ScholarPubMed
Vítek, P., Cámara-Gallego, B., Edwards, H.G.M. et al. (2013). Phototrophic community in gypsum crust from the Atacama Desert studied by Raman spectroscopy and microscopic imaging. Geomicrobiology Journal, 30, 399–410.Google Scholar
Webster, C.R., Mahaffy, P.R., Atreya, S.K., et al. and the MSL Science Team (2015). Mars methane detection and variability at Gale Crater. Science, 347, 415417.Google Scholar
Webster, C.R., Mahaffy, P.R., Atreya, S.K., et al. (2018). Background levels of methane in Mars’ atmosphere show strong seasonal variation. Science, 360, 10931096.Google Scholar
Weesie, R.J., Merlin, J.C., Lugtenburg, J., et al. (1999). Semiempirical and Raman spectroscopic study of carotenoids. Biospectroscopy, 5, 1933.3.0.CO;2-E>CrossRefGoogle ScholarPubMed
Westall, F., Cavalazzi, B. (2011). Biosignatures in rocks. In: Thieland, V., Reitner, J. (eds) Encyclopaedia of Geobiology. Springer, Berlin, pp. 189201.Google Scholar
Westall, F., Foucher, F., Bost, N., et al. (2015). Biosignatures on Mars: What, where and how? Implications for the search for Martian Life. Astrobiology, 15, 132.Google Scholar
Wynn-Williams, D.D. (1991). Cyanobacteria in deserts –life at the limit? In: Whitton, B.A, Potts, M (eds) The Ecology of Cyanobacteria: Their Diversity in Time and Space. Kluwer, Academic Press, Dordrecht, The Netherlands, pp. 341366.Google Scholar
Wynn-Williams, D.D. (1999). Antarctica as a model for ancient Mars. In: Hiscox, J.A. (ed.) The Search for Life on Mars. British Interplanetary Society, London, pp. 4957.Google Scholar
Wynn-Williams, D.D., Edwards, H.G.M. (2000a). Antarctic ecosystems as models for extraterrestrial surface habitats. Planetary and Space Sciences, 48, 10651075.Google Scholar
Wynn-Williams, D.D., Edwards, H.G.M. (2000b).Proximal analysis of regolith habitats and protective biomolecules in situ by laser Raman spectroscopy: overview of terrestrial Antarctic habitats and Mars analogs. Icarus, 144, 486503.Google Scholar
Wynn-Williams, D.D., Edwards, H.G.M. (2002). Environmental UV-radiation: biological strategies for protection and avoidance. In: Horneck, G, Baumstarck-Khan, C (eds) Astrobiology: The Quest for the Origins of Life, Springer-Verlag, Berlin, pp. 245260.Google Scholar

References

Aguilera, A., Manrubia, S.C., Gómez, F., Rodríguez, N., Amils, R. (2006). Eukaryotic community distribution and its relationship to water physicochemical parameters in an extreme acidic environment, Río Tinto (Southwestern Spain). Applied and Environmental Microbiology, 72, 53255330.Google Scholar
Aguilera, A., Souza-Egipsy, V., Gomez, F., Amils, R. (2007a). Development and structure of eukaryotic biofilms in an extreme acidic environment, Rio Tinto (SW, Spain). Microbial Ecology, 53, 294305.Google Scholar
Aguilera, A., Zettler, E., Gomez, F., et al (2007b). Distribution and seasonal variability in the benthic eukaryotic community of Rio Tinto (SW, Spain), an acidic, high metal extreme environment. Systematic and Applied Microbiology, 30, 531546.CrossRefGoogle Scholar
Allen, J.F. (2003). State transitions – a question of balance. Science, 299, 15301532.Google Scholar
Allewalt, J.P., Bateson, M.M., Revsbech, N.P., Slack, K., Ward, D.M. (2006). Effect of temperature and light on growth of and photosynthesis by Synechococcus isolates typical of those predominating in the octopus spring microbial mat community of Yellowstone National Park. Applied and Environmental Microbiology, 72, 544550.CrossRefGoogle ScholarPubMed
Amils, R., González-Toril, E., Fernández-Remolar, D., et al. (2007). Extreme environments as Mars terrestrial analogs: The Rio Tinto case. Planetary and Space Science, 55, 370381.Google Scholar
Angelini, G., Ragni, P., Eposito, D., et al. (2001). A device to study the effect of space radiation on photosynthetic organisms. Physica Medica, XVIII, 267268.Google Scholar
Apte, S.K. (2001). Copying with salinity/water stress: cyanobacteria show the way. Proceedings of the Indian National Science Academy, B67, 285310.Google Scholar
Avron, M., Ben-Amotz, A. (1992). Dunaliella: Physiology, Biochemistry and Biotechnology. Boca Raton, FL: CRC Press.Google Scholar
Axelsson, L., Beer, S. (2001). Carbon limitation. In:Rai, L.C.,Gaur, J.P. (eds) Algal Adaptation to Environmental Stresses. Physiological, Biochemical and Molecular Mechanisms. Springer-Verlag, Berlin/Heidelberg/New York, pp. 2144.Google Scholar
Ballesteros, F., Fernandez-Soto, A., Martínez, V. (2019). Diving into exoplanets: are water seas the most common? Astrobiology. May 2019. Published ahead of print; http://doi.org/10.1089/ast.2017.1720.CrossRefGoogle Scholar
Baqué, M., Hanke, F., Böttger, U., et al. (2018). Protection of cyanobacterial carotenoids’ Raman signatures by Martian mineral analogues after high‐dose gamma irradiation. Journal of Raman Spectroscopy, 49, 16171627.Google Scholar
Baxter, A., Mittler, R., Suzuki, N. (2013). ROS as key players in plant stress signalling. Journal of Experimental Botany, 65, 12291240.Google Scholar
Bekker, A., Holland, H., Wang, P.-L., et al. (2004). Dating the rise of atmospheric oxygen. Nature, 427, 117.Google Scholar
Bhatia, S., Garg, A., Sharma, K., et al. (2011). Mycosporine and mycosporine-like amino acids: A paramount tool against ultra violet irradiation. Pharmacognosy Review, 5, 138146.Google Scholar
Bidigare, R.R., Ondrusek, M.E., Kennicutt, M.C., et al. (1993). Evidence for a photoprotective function for secondary carotenoids of snow algae. Journal of Phycology, 29, 427434.Google Scholar
Brandt, A., de Vera, J.-P., Onofri, S., Ott, S. (2015). Viability of the lichen Xanthoria elegans and its symbionts after 18 months of space exposure and simulated Mars conditions on the ISS. International Journal of Astrobiology, 14, 411425.Google Scholar
Bray, C.M., West, C.E. (2005). DNA repair mechanisms in plants: crucial sensors and effectors for the maintenance of genome integrity. New Phytologist, 168, 511528.CrossRefGoogle ScholarPubMed
Broady, P.A. (1984). Taxonomic and ecological investigations of algae on steam-warmed soil on Mt Erebus, Ross Island, Antarctica. Phycologia, 23, 257271.CrossRefGoogle Scholar
Buick, R. (2008). When did oxygenic photosynthesis evolve? Philosophical Transactions of the Royal Society of London B: Biological Sciences, 363, 27312743.Google Scholar
Burns, B.P., Goh, F., Allen, M., Neilan, B.A. (2004). Microbial diversity of extant stromatolites in the hypersaline marine environment of Shark Bay, Australia. Environmental Microbiology, 6, 10961101.Google Scholar
Chapman, R.L. (2013). Algae: the world’s most important ‘plants’ – an introduction. Mitigation and Adaptation Strategies for Global Change, 18, 512.Google Scholar
Christiansen, J. (2018). Exoplanet catalogs. In: Deeg, H.J., Belmonte, J.A. (eds) Handbook of Exoplanets. Springer International Publishing, Cham, pp. 19331947.Google Scholar
Clegg, M.R., Maberly, S.C., Jones, R.I. (2003). Chemosensory behavioural response of freshwater phytoplanktonic flagellates. Plant Cell and Environment, 27, 123135.CrossRefGoogle Scholar
Cockell, C.S., Knowland, J. (1999). Ultraviolet radiation screening compounds. Biological Reviews, 79, 311345.Google Scholar
Cockell, C.S., McMahon, S. (2019). Lifeless Martian samples and their significance. Nature Astronomy, 3, 468470.CrossRefGoogle Scholar
Cockell, C.S., Rettberg, P., Rabbow, E., Olsson-Francis, K. (2011). Exposure of phototrophs to 548 days in low Earth orbit: microbial selection pressures in outer space and on early earth. ISME Journal, 5, 16711682.Google Scholar
de los Ríos, A., Ascaso, C., Wierzchos, J., Fernández-Valiente, E., Quesada, A. (2004). Microstructural characterization of cyanobacterial mats from the McMurdo Ice Shelf, Antarctica. Applied and Environmental Microbiology, 70, 569580.Google Scholar
de Vera, J. P., Horneck, G., Rettberg, P., Ott, S. (2004). The potential of the lichen symbiosis to cope with the extreme conditions of outer space II: germination capacity of lichen ascospores in response to simulated space conditions. Advances in Space Research, 33, 12361243.Google Scholar
de Vera, J.-P., Schulze-Makuch, D., Khan, A., et al. (2014). Adaptation of an Antarctic lichen to Martian niche conditions can occur within 34 days. Planetary and Space Science, 98, 182190.CrossRefGoogle Scholar
Des Marais, D.J., Jahnke, L.L. (2019). Biosignatures of cellular components and metabolic activity. In: Cavalazzi, B,Westall, F (eds) Biosignatures for Astrobiology. Springer International Publishing, Cham, pp. 5185.CrossRefGoogle Scholar
Doran, P.T., Lyons, W.B., McKnight, D.M. (2010). Life in Antarctic Deserts and Other Cold Dry Environments: Astrobiological Analogs. Cambridge University Press, Cambridge.Google Scholar
Dundas, C.M., Diniega, S., Hansen, C.J., Byrne, S., McEwen, A.S. (2012). Seasonal activity and morphological changes in martian gullies. Icarus, 220, 124143.Google Scholar
Dundas, C.M., Bramson, A.M., Ojha, L., et al. (2018). Exposed subsurface ice sheets in the Martian mid-latitudes. Science, 359, 199201.Google Scholar
Ehling-Schulz, M., Scherer, S. (1999). UV protection in cyanobacteria. European Journal of Phycology, 34, 329338.Google Scholar
Elster, J. (1999). Algal versatility in various extreme environments. In: Seckbach, J (ed.) Enigmatic Microorganisms and Life in Extreme Environments. Kluwer Academic Publishers,Dordrecht/Boston/London, pp. 215227.Google Scholar
Elster, J., Svoboda, J. (1995). In situ simulation and manipulation of a glacial stream ecosystem in the Canadian High Arctic. In:Jenkins, D, Ferrier, R.C., Kirby, C (eds) Ecosystem Manipulation Experiments: Scientific Approaches, Experimental Design and Relevant Results. Commission of the European Communities, Institute of Hydrology, UK and Environment Canada, Brussels, pp. 254263.Google Scholar
Elster, J., Svoboda, J., Kanda, H. (2001). Controlled environmental platform used in temperature manipulation study of a stream periphyton in the Ny-Ålesund, Svalbard. In: Elster, J, Seckbach, J, Vincent, W.F.,Lhotský, O (eds) Algae and Extreme Environments. Cramer, Stuttgart, pp. 6375.Google Scholar
Elster, J., Kvíderová, J., Hájek, T., Láska, K., Šimek, M. (2012). Impact of warming on Nostoc colonies (Cyanobacteria) in a wet hummock meadow, Spitzbergen. Polish Polar Research, 33, 395420.Google Scholar
Fazi, S., Butturini, A., Tassi, F., et al. (2018). Biogeochemistry and biodiversity in a network of saline–alkaline lakes: implications of ecohydrological connectivity in the Kenyan Rift Valley. Ecohydrology & Hydrobiology, 18, 96106.Google Scholar
Finlay, J.C. (2004). Patterns and controls of lotic algal stable carbon isotope ratios. Limnology and Oceanography, 49, 850861.Google Scholar
Fujita, Y., Ohki, K., Murakami, A. (1985). Cromatic regulation of photosystem composition in the photosynthetic system of red and blue-green algae. Plant and Cell Physiology, 26, 15411548.Google Scholar
Fukuda, S.-y., Yamakawa, R., Hirai, M., et al. (2008). Mechanisms to avoid photoinhibition in a desiccation-tolerant cyanobacterium, Nostoc commune. Plant and Cell Physiology, 49, 488492.Google Scholar
Gelpi, E., Schneider, H., Mann, J., Oró, J. (1970). Hydrocarbons of geochemical significance in microscopic algae. Phytochemistry, 9, 603612.Google Scholar
Gimmler, H. (2001). Acidophilic and acidotolerant algae. In: Rai, L.C., Gaur, J.P. (eds) Algal Adaptation to Environmental Stresses. Physiological, Biochemical and Molecular Mechanisms. Springer-Verlag, Berlin/Heidelberg/New York, pp. 259290.CrossRefGoogle Scholar
Gimmler, H., Degenhardt, B. (2001). Alkaliphilic and alkalitolerant algae. In: Rai, L.C., Gaur, J. P. (eds) Algal Adaptation to Environmental Stresses. Physiological, Biochemical and Molecular Mechanisms. Springer-Verlag, Berlin/Heidelberg/New York, pp. 291322.Google Scholar
Glein, C.R., Baross, J.A., WaiteJr, J.H. (2015). The pH of Enceladus’ ocean. Geochimica et Cosmochimica Acta, 162, 202219.Google Scholar
Gòdia, F., Albiol, J., Montesinos, J.L., et al. (2002). MELISSA: a loop of interconnected bioreactors to develop life support in Space. Journal of Biotechnology, 99, 319330.Google Scholar
Gombos, Z., Wada, H., Hideg, E., Murata, N. (1994a). The unsaturation of membrane lipids stabilizes photosynthesis against heat stress. Plant Physiology, 104, 563567.Google Scholar
Gombos, Z., Wada, H., Murata, N. (1994b). The recovery of photosynthesis from low-temperature photoinhibition is accelerated by the unsaturation of membrane lipids: A mechanism of chilling tolerance. Proceedings of the National Academy of Sciences of the USA, 91, 87878791.Google Scholar
Gómez, F., Walter, N., Amils, R., et al. (2011). Multidisciplinary integrated field campaign to an acidic Martian Earth analogue with astrobiological interest: Rio Tinto. International Journal of Astrobiology, 10, 291305.Google Scholar
Gorbushina, A.A., Krumbein, W.E. (1999). The poikilotrophic micro-organism and its environment. In: Seckbach, J (eds.) Enigmatic Microorganisms and Life in Extreme Environments. Kluwer Academic Press,Dordrecht/Boston/London, pp. 177185.Google Scholar
Hägele, D., Leinfelder, R., Grau, J., Burmeister, E.-G., Struck, U. (2006). Oncoids from the river Alz (southern Germany): tiny ecosystems in a phosphorus-limited environment. Palaeogeography Palaeoclimatology Palaeoecology, 237, 378395.Google Scholar
Henley, W.J. (2001). Algae under desiccation stress. In: Elster, J., Seckbach, J., Vincent, W.F., Lhotský, O. (eds) Algae and Extreme Environments. J. Cramer in der Gebr. Borntraeger Verlagsbuchhandlung,Stuttgart, pp. 443452.Google Scholar
Hindák, F., Kvíderová, J., Lukavský, J. (2013). Growth characteristics of selected thermophilic strains of cyanobacteria using crossed gradients of temperature and light. Biologia, 68, 830837.Google Scholar
Hoham, R.W., Duval, B. (2001). Microbial ecology of snow and freshwater ice with emphasis on snow algae. In: Snow Ecology: An Interdisciplinary Examination of Snow-Covered Ecosystems, eds.Jones, H. G.,Pomeroy, J. W., Walker, D. A., Hoham, R. W. (eds). Cambridge University Press, Cambridge, UK, pp. 168228.Google Scholar
Hoiczyk, E. (2000). Gliding motility in cyanobacteria: observations and possible explanations. Archives of Microbiology, 174, 1117.Google Scholar
Horneck, G., Debus, A., Mani, P., Spry, J.A. (2007). Astrobiology exploratory missions and planetary protection requirements. In: Horneck, G,Rettberg, P (eds) Complete Course in Astrobiology, Wiley VCH,Weinheim, pp. 353397.Google Scholar
Javaux, E. J. (2015). Biomarkers. In: Gargaud, M,Irvine, W.M.,Amils, R (eds) Encyclopedia of Astrobiology. Springer Berlin Heidelberg, Berlin, Heidelberg, pp. 271293.Google Scholar
Jorge Villar, S.E., Edwards, H.G.M. (2006). Raman spectroscopy in astrobiology. Analytical and Bioanalytical Chemistry, 384, 100113.Google Scholar
Joset, F., Jeanjean, R., Hagemann, M. (1996). Dynamics of the response of cyanobacteria to salt stress: Deciphering the molecular events. Physiologia Plantarum, 96, 738744.Google Scholar
Kappen, L., Bölter, M., Kühn, A. (1986). Field measurements of net photosynthesis of lichens in the Antarctic. Polar Biology, 5, 255258.Google Scholar
Kappen, L., Schroeter, B., Scheidegger, C., Sommerkorn, M., Hestmark, G. (1996). Cold resistance and metabolic activity of lichens below 0°C. Advances in Space Research, 18, 119128.Google Scholar
Kereszturi, A. (2012). Review of wet environment types on Mars with focus on duration and volumetric issues. Astrobiology, 12, 586600.Google Scholar
Kiang, N.Y., Segura, A., Tinetti, G., et al. (2007a). Spectral signatures of photosynthesis. II. Coevolution with other stars and atmosphere on extrasolar worlds. Astrobiology, 7, 252274.Google Scholar
Kiang, N.Y., Siefert, J., Govindjee, , Blankenship, R.E. (2007b). Spectral signatures of photosynthesis. I. Review of Earth organisms. Astrobiology, 7, 222251.Google Scholar
Kirst, G.O., Thiel, C., Wolff, H., et al. (1991). Dimethylsulfoniopropionate (DMSP) in icealgae and its possible biological role. Marine Chemistry, 35, 381388.Google Scholar
Komárek, J., Nedbalová, L. (2007). Green cryosestic algae. In:Seckbach, J (ed.) Algae and Cyanobacteria in Extreme Environments. Springer, Dordrecht, pp. 323342.Google Scholar
Kováčik, Ľ., Jezberová, J., Komárková, J., Kopecký, J., Komárek, J. (2011). Ecological characteristics and polyphasic taxonomic classification of stable pigment-types of the genus Chroococcus (Cyanobacteria). Preslia, 83, 145166.Google Scholar
Krause, G.H., Weis, E. (1984). Chlorophyll fluorescence as a tool in plant physiology. II. Interpretation of fluorescence signals. Photosynthesis Research, 5, 139157.Google Scholar
Kregel, K.C. (2002). Invited review: heat shock proteins: modifying factors in physiological stress responses and acquired thermotolerance. Journal of Applied Physiology, 92, 21772186.Google Scholar
Kumar, D., Kvíderová, J., Kaštánek, P., Lukavský, J. (2017). The green alga Dictyosphaerium chlorelloides biomass and polysaccharides production determined using cultivation in crossed gradients of temperature and light. Engineering in Life Sciences, 17, 10301038.Google Scholar
Kvíderová, J. (2012). Photochemical performance of the acidophilic red alga Cyanidium sp. in a pH gradient. Origins of Life and Evolution of Biospheres, 42, 223234.Google Scholar
Kvíderová, J., Lukavský, J. (2005). The comparison of ecological characteristics of Stichococcus (Chlorophyta) strains isolated from polar and temperate regions. Algological Studies, 118, 127140.Google Scholar
Kvíderová, J., Elster, J., Šimek, M. (2011). In situ response of Nostoc commune s.l. colonies to desiccation in Central Svalbard, Norwegian High Arctic. Fottea, 11, 8797.Google Scholar
Kvíderová, J., Souquieres, C.-E., Elster, J. (2018). Ecophysiology of photosynthesis of Vaucheria sp. mats in a Svalbard tidal flat. Polar Science.Google Scholar
Kvíderová, J., Elster, J., Komárek, J. (2019). Ecophysiology of cyanobacteria in the Polar Regions. In: Mishra, A.K., Tiwari, D.N., Rai, A.N. (eds) Cyanobacteria. Academic Press, London, San Diego, Cambridge, Oxford, pp. 277302.Google Scholar
Ledford, H.K., Niyogi, K.K. (2005). Singlet oxygen and photo-oxidative stress management in plants and algae. Plant, Cell & Environment, 28, 10371045.Google Scholar
Leigh, J.A., Stahl, D.A., Staley, J.T. (2007). Evolution of metabolism and early microbial communities. In: W.F.I., Sullivan, J.A., Baross (eds) Planets and Life. The Emerging Science of Astrobiology. Cambridge University Press, Cambridge, UK, pp. 222236.Google Scholar
Li, H.Y., Liu, X.X., Wang, Y.L., Hu, H.H., Xu, X.D. (2009). Enhanced expression of antifreeze protein genes drives the development of freeze tolerance in an Antarctica isolate of Chlorella vulgaris. Progress in Natural Science, 19, 10591062.Google Scholar
Lyons, T.W., Reinhard, C.T., Planavsky, N.J. (2014). The rise of oxygen in Earth’s early ocean and atmosphere. Nature, 506, 307.Google Scholar
MacIntyre, H.L., Kana, T.M., Anning, T., Geider, R.J. (2002). Photoacclimation of photosynthesis irradiance response curves and photosynthetic pigments in microalgae and cyanobacteria. Journal of Phycology, 38, 1738.Google Scholar
Mackey, T., Sumner, D., Hawes, I., et al. (2018). Stromatolite records of environmental change in perennially ice-covered Lake Joyce, McMurdo Dry Valleys, Antarctica. Biogeochemistry, 137, 7392.Google Scholar
Marshall, C.P., Edwards, H.G.M., Jehlicka, J. (2010). Understanding the application of Raman spectroscopy to the detection of traces of life. Astrobiology, 10, 229243.Google Scholar
Matsui, K., Nazifi, E., Hirai, Y., et al. (2012). The cyanobacterial UV-absorbing pigment scytonemin displays radical-scavenging activity. Journal of General and Applied Microbiology, 58, 137144.Google Scholar
McEwen, A.S., Ojha, L., Dundas, C.M., et al. (2011). Seasonal flows on warm Martian slopes. Science, 333, 740743.Google Scholar
McKay, C.P., Nienow, J.A., Meyer, M.A., Friedmann, E.I. (1993). Continuous nanoclimate data (1985–1988) from the Ross Desert (McMurdo Dry Valleys) cryptoendolithic microbial ecosystem. In:Bromwich, D.H., Stearns, C. R. (eds) Antarctic Meteorology and Climatology: Studies Based on Automatic Weather Stations. American Geophysical Union,Washington, DC, pp. 201207.Google Scholar
Meleshko, G.I., Shepelev, E.Y., Kordyum, E.I., Šetlík, I., Doucha, J. (1986). Nevesomost i yeyo vliyanie na mikroorganizmy i rasteniya [Weightlessness and its influence on microorganisms and plants]. In: Gurovskiy, N.N. (ed.) Rezultaty medicinskich issledovaniy vypolnenych na orbitalnom nauchno-issledovatelskom komplekse ‘Salyut-6’ – ‘Soyuz’ [Results of medical research carried out onborad scientic-research complex ‘Salyut-6’ – ‘Soyuz’]. Nauka, Moscow, pp. 369391.Google Scholar
Melis, A. (1999). Photosystem-II damage and repair cycle in chloroplasts: what modulates the rate of photodamage in vivo? Trends in Plant Sciences, 4, 130135.Google Scholar
Minagawa, M., Wada, E. (1986). Nitrogen isotope ratios of red tide organisms in the East China Sea: a characterization of biological nitrogen fixation. Marine Chemistry, 19, 245259.Google Scholar
Moon, B.Y., Higashi, S.I., Gombos, Z., Murata, N. (1995). Unsaturation of the membrane lipids of chloroplasts stabilizes the photosynthetic mashinery against low-temperature photoinhibition in transgenic tobacco plants. Proceedings of the National Academy of Sciences of the USA, 92, 62196223.Google Scholar
Moore, L.S., Burne, R. (1994). The modern thrombolites of Lake Clifton, western Australia. In: Bertrand-Sarfati, J, Monty, C (eds) Phanerozoic Stromatolites II. Springer, Dordrecht, pp. 329.Google Scholar
Moorthi, S., Caron, D.A., Gast, R.J., Sanders, R.W. (2009). Mixotrophy: a widespread and important ecological strategy for planktonic and sea-ice nanoflagellates in the Ross Sea, Antarctica. Aquatic Microbial Ecology, 54, 269277.Google Scholar
Nealson, K.H. (1997). The limits of life on Earth and searching for life on Mars. Journal of Geophysical Research: Planets, 102, 2367523686.Google Scholar
Nedbal, L., Soukupová, J., Kaftan, D., Whitmarsh, J., Trtílek, M. (2000). Kinetic imaging of chlorophyll fluorescence using modulated light. Photosynthesis Research, 66, 312.Google Scholar
Nedbal, L., Whitmarsh, J. (2004). Chlorophyll fluorescence imaging of leaves and fruits. In: Papageorgiou, G.C., Govindjee, (eds) Chlorophyll a Fluorescence. A Signature of Photosynthesis. Springer, Dordrecht, pp. 389407.Google Scholar
Nienow, J.A., McKay, C.P., Friedmann, E.I. (1988). Cryptoendolithic microbial environment in the Ross Desert of Antarctica: light in the photosynthetically active region. Microbial Ecology, 16, 271289.Google Scholar
Noffke, N. (2015). Ancient sedimentary structures in the < 3.7 ga Gillespie Lake Member, Mars, that resemble macroscopic morphology, spatial associations, and temporal succession in terrestrial microbialites. Astrobiology, 15, 169192.Google Scholar
Noffke, N., Gerdes, G., Klenke, T., Krumbein, W.E. (2001). Microbially induced sedimentary structures: a new category within the classification of primary sedimentary structures. Journal of Sedimentary Research, 71, 649656.Google Scholar
Ojha, L., Wilhelm, M.B., Murchie, S.L., et al. (2015). Spectral evidence for hydrated salts in recurring slope lineae on Mars. Nature Geoscience, 8, 829832.Google Scholar
Oliver, R.L. (1994). Floating and sinking in gas-vacuoolate cyanobacteria. Journal of Phycology, 30, 161173.CrossRefGoogle Scholar
Olson, J.M. (2006). Photosynthesis in the Archean Era. Photosynthesis Research, 88, 109117.CrossRefGoogle ScholarPubMed
Olsson-Francis, K., de la Torre, R., Towner, M.C., Cockell, C.S. (2009). Survival of akinetes (resting-state cells of cyanobacteria) in low earth orbit and simulated extraterrestrial conditions. Origins of Life and Evolution of Biosphere, 39, 565579.Google Scholar
Onofri, S., de la Torre, R., de Vera, J.P., et al. (2012). Survival of rock-colonizing organisms after 1.5 years in outer space. Astrobiology, 12, 508516.Google Scholar
Oren, A. (2007). Diversity of organic osmotic compounds and osmotic adaptation in cyanobacteria and algae. In: Seckbach, J (ed.) Algae and Cyanobacteria in Extreme Environments. Springer, Dordrecht, pp. 639655.Google Scholar
Oren, A., Seckbach, J. (2001). Oxygenic photosynthesis in extreme environments. In: Elster, J, Seckbach, J, Vincent, W.F., Lhotský, O (eds) Algae and Extreme Environments. J. Cramer in der Gebr. Borntraeger Verlagsbuchhandlung, Berlin, Stuttgart, pp. 1331.Google Scholar
Parker, B.C., SimmonsJr, G.M., Love, F.G., Wharton Jr, R.A, Seaburg, K.G. (1981). Modern stromatolites in Antarctic dry valley lakes. Bioscience, 31, 656661.Google Scholar
Popova, A.F. (2003). Comparative characteristic of mitochondria ultrastructural organisation in Chlorella cells under altered gravity conditions. Advances in Space Research, 31, 22532259.Google Scholar
Popova, A.F., Sytnik, K.M. (1996). Peculiarities of ultrastructure of Chlorella cells growing aboard the Bion-10 during 12 days. Advances in Space Research, 17, 99102.Google Scholar
Popova, A.F., Sytnik, K.M., Kordyum, E.L., et al. (1989). Ultrastructural and growth indices of Chlorella culture in multicomponent aquatic systems under space flight conditions. Advances in Space Research, 9, 7982.Google Scholar
Potts, M. (1994). Desiccation tolerance of prokaryotes. Microbiology and Molecular Biology Reviews, 58, 755805.Google Scholar
Potts, M. (1999). Mechanisms of desiccation tolerance in cyanobacteria. European Journal of Phycology, 34, 319328.Google Scholar
Rabbow, E., Horneck, G., Rettberg, P., et al. (2009). EXPOSE, an astrobiological exposure facility on the International Space Station – from proposal to flight. Origins of Life and Evolution of Biospheres, 39, 581598.Google Scholar
Rasmussen, B., Fletcher, I.R., Brocks, J.J., Kilburn, M.R. (2008). Reassessing the first appearance of eukaryotes and cyanobacteria. Nature, 455, 1101.Google Scholar
Raven, J.A. (2003). Inorganic carbon concentrating mechanisms in relation to the biology of algae. Photosynthesis Research, 77, 155171.Google Scholar
Rettberg, P., Rabbow, E., Panitz, C., Horneck, G. (2004). Biological space experiments for the simulation of Martian conditions: UV radiation and Martian soil analogues. Advances in Space Research, 33, 12941301.Google Scholar
Sachindra, N.M., Sato, E., Maeda, H., et al. (2007). Radical scavenging and singlet oxygen quenching activity of marine carotenoid fucoxanthin and its metabolites. Journal of Agricultural and Food Chemistry, 55, 85168522.Google Scholar
Seckbach, J., Chapman, D., Garbary, D., Oren, A., Reisser, W. (2007). Algae and cyanobacteria under environmental extremes. In: Seckbach, J (ed.) Algae and Cyanobacteria in Extreme Environments. Springer, Dordrecht, pp. 781786.Google Scholar
Schmitt, F.-J., Campbell, Z.Y., Bui, M.V., et al. (2019). Photosynthesis supported by a chlorophyll f-dependent, entropy-driven uphill energy transfer in Halomicronema hongdechloris cells adapted to far-red light. Photosynthesis Research, 139, 185201.Google Scholar
Schneider, T., Schmid, E., de Castro, J.V., et al. (2011). Structure and function of the symbiosis partners of the lung lichen (Lobaria pulmonaria L. Hoffm.) analyzed by metaproteomics. Proteomics, 11, 27522756.Google Scholar
Schreiber, U., Neubauer, C. (1990). O2-dependent electron flow, membrane energization and the mechanism of non-photochemical quenching of chlorophyll fluorescence. Photosynthesis Research, 25, 279293.Google Scholar
Schulze, E.-D., Beck, E., Müller-, Hohenstein (2005). Plant Ecology. Springer, Berlin, Heidelberg.Google Scholar
Schulze-Makuch, D., Irwin, L.N. (2018). Energy sources and life. In: Schulze-Makuch, D, Irwin, L. N. (eds) Life in the Universe. Expectations and Constraints. Springer Nature Switzerland AG, Cham, pp. 75100.Google Scholar
Šetlík, I., Kordyum, V.A., Meleshko, G.I., et al. (1979). Experiment Chlorella on the board of Salyut 6. In: 29th Congress International Astronautical Federation IAF’78. Pergamon Press, London.Google Scholar
Shehawy, R.M., Kleiner, D. (2001). Nitrogen limitation. In: Rai, L. C., Gaur, J. P. (eds) Algal Adaptation to Environmental Stresses. Physiological, Biochemical and Molecular Mechanisms. Berlin, Heidelberg, New York, Springer-Verlag, pp. 4564.Google Scholar
Soo, R.M., Wood, S.A., Grzymski, J.J., McDonald, I.R., Cary, S.C. (2009). Microbial biodiversity of thermophilic communities in hot mineral soils of Tramway Ridge, Mount Erebus, Antarctica. Environmental Microbiology, 11, 715728.Google Scholar
Squyres, S.W., Knoll, A.H. (2005). Sedimentary rocks at Meridiani planum: origin, diagenesis, and implications for life on Mars. Earth and Planetary Science Letters, 240, 110.CrossRefGoogle Scholar
Stamenković, V., Ward, L.M., Mischna, M., Fischer, W.W. (2018). O2 solubility in Martian near-surface environments and implications for aerobic life. Nature Geoscience, 11, 905909.Google Scholar
Storrie-Lombardi, M.C., Sattler, B. (2009). Laser-induced fluorescence emission (LIFE): in situ nondestructive detection of microbial life in the ice covers of Antarctic lakes. Astrobiology, 9, 659672.Google Scholar
Storrie-Lombardi, M.C., Muller, J.-P., Fisk, M.R., et al. (2009). Laser-Induced Fluorescence Emission (LIFE): searching for Mars organics with a UV-enhanced PanCam. Astrobiology, 9, 953964.Google Scholar
Sytnik, K.M., Popova, A.F., Nechitailo, G.S., Mashinsky, A.L. (1992). Peculiarities of the submicroscopic organization of Chlorella cells cultivated on a solid medium in microgravity. Advances in Space Research, 12, 103107.Google Scholar
Takahashi, S., Badger, M.R. (2011). Photoprotection in plants: a new light on photosystem II damage. Trends in Plant Science, 16, 5360.Google Scholar
Tashyreva, D., Elster, J. (2012). Production of dormant stages and stress resistance of polar cyanobacteria. In: Hanslmeier, A, Kempe, S, Seckbach, J (eds) Life on Earth and Other Planetary Bodies. Springer, Dordrecht, pp. 367386.Google Scholar
Tashyreva, D., Elster, J. (2016). Annual cycles of two cyanobacterial mat communities in hydro-terrestrial habitats of the High Arctic. Microbial Ecology, 71, 887900.Google Scholar
Thannickal, V.J. (2009). Oxygen in the evolution of complex life and the price we pay. American Journal of Respiratory Cell and Molecular Biology, 40, 507510.Google Scholar
van den Hoek, C., Mann, D.G., Jahns, H.M. (1995). Algae: An Introduction to Phycology. Cambridge University Press, Cambridge, UK.Google Scholar
Vázques, M., Pallé, E., Montanés Rodrigues, P. (2010). Earth as a Distant Planet. Springer,New York, Dordrecht, Heidelberg, London.Google Scholar
Wagner, F., Falkner, G. (2001). Phosphate limitation. In: Rai, L.C., Gaur, J.P. (eds) Algal Adaptation to Environmental Stresses. Physiological, Biochemical and Molecular Mechanisms. Springer-Verlag,Berlin, Heidelberg, New York, pp. 65110.Google Scholar
Wang, G., Chen, H., Li, G., et al. (2006). Population growth and physiological characteristics of microalgae in a miniaturized bioreactor during space flight. Acta Astronautica, 58, 264269.Google Scholar
Wang, G.H., Li, G.B., Li, D.H., et al. (2004). Real-time studies on microalgae under microgravity. Acta Astronautica, 55, 131137.Google Scholar
Whitley, D., Goldberg, S.P., Jordan, W.D. (1999). Heat shock proteins: a review of the molecular chaperones. Journal of Vascular Surgery, 29, 748751.Google Scholar
Xiong, F., Lederer, F., Lukavský, J., Nedbal, L. (1996). Screening of freshwater algae (Chlorophyta, Chromophyta) for ultraviolet-B sensitivity of the photosynthetic apparatus. Journal of Plant Physiology, 148, 4248.Google Scholar

References

Aliyu, A.S., Ramli, A.T. (2015). The world’s high background natural radiation areas (HBNRAs) revisited: a broad overview of the dosimetric, epidemiological and radiobiological issues. Radiation Measurements, 73, 5159.Google Scholar
Aronson, R.B., Frederich, M., Price, R., Thatje, S., Alistair Crame, J. (2015). Prospects for the return of shell-crushing crabs to Antarctica. Journal of Biogeography, 42, 17.Google Scholar
Ashcroft, F. (2000). Life at the Extremes. The Science of Survival. Flamingo, London.Google Scholar
Bonnington, C., Gaston, K.J., Evans, K.L., Whittingham, M. (2013). Fearing the feline: domestic cats reduce avian fecundity through trait-mediated indirect effects that increase nest predation by other species. Journal of Applied Ecology, 50, 1524.Google Scholar
Bowler, K., Terblanche, J.S. (2008). Insect thermal tolerance: what is the role of ontogeny, ageing and senescence? Biological Reviews, 83, 339355.Google Scholar
Brown, R.R., Deletic, A., Wong, T.H.F. (2015). How to catalyse collaboration. Nature, 525, 315317.Google Scholar
Childs, D.Z., Metcalf, C.J.E., Rees, M. (2010). Evolutionary bet-hedging in the real world: empirical evidence and challenges revealed by plants. Proceedings of the Royal Society B, 277, 30553064.Google Scholar
Chown, S.L., Terblanche, J.S. (2007). Physiological diversity in insects: ecological and evolutionary contexts. Advances in Insect Physiology, 33, 50152.Google Scholar
Chown, S.L., Hoffmann, A.A., Kristensen, T.N., et al. (2010). Adapting to climate change: a perspective from evolutionary physiology. Climate Research, 43, 315.Google Scholar
Chown, S.L., Sørensen, J.G., Terblanche, J.S. (2011). Water loss in insects: an environmental change perspective. Journal of Insect Physiology, 57, 10701084.CrossRefGoogle ScholarPubMed
Chown, S.L., Clarke, A., Fraser, C.I., et al. (2015). The changing form of Antarctic biodiversity. Nature, 522, 431438.Google Scholar
Christian, K.A., Morton, S.R. (1992). Extreme thermophilia in a Central Australian Ant, Melophorus bagoti. Physiological Zoology, 65, 885905.Google Scholar
Clark, M.S., Verde, C., Fineschi, S., et al. (2020). Life in the extreme environments of our planet under pressure: climate-induced threats and exploitation opportunities. In:di Prisco, G, Huiskes, A, Elster, J, Edwards, H (eds) Life in Extreme Environments. Cambridge University Press, Cambridge.Google Scholar
Clarke, A. (2014). The thermal limits to life on Earth. International Journal of Astrobiology, 13, 141154.Google Scholar
Cooke, S.J., Sack, L., Franklin, C.E., et al. (2013). What is conservation physiology? Perspectives on an increasingly integrated and essential science. Conservation Physiology, 1, cot001.Google Scholar
Corkrey, R., McMeekin, T.A., Bowman, J.P., et al. (2016). The biokinetic spectrum for temperature. PLoS ONE, 11, e0153343.Google Scholar
Cowling, R.M., Esler, K.J., Rundel, P.W. (1999). Namaqualand, South Africa – an overview of a unique winter-rainfall desert ecosystem. Plant Ecology, 142, 321.Google Scholar
Del Giudice, M., Buck, C.L., Chaby, L.E., et al. (2018). What is stress? A systems perspective. Integrative and Comparative Biology, 58, 10191032.Google Scholar
Desmet, P.G., Cowling, R.M. (1999). Biodiversity, habitat and range-size aspects of a flora from a winter-rainfall desert in north-western Namaqualand, South Africa. Plant Ecology, 142, 2333.Google Scholar
Dowd, W.W., Denny, M.W. (2020). A series of unfortunate events: characterizing the contingent nature of physiological extremes using long-term environmental records. Proceedings of the Royal Society B, 287, 20192333.Google Scholar
Eldredge, N. (1986). Information, economics, and evolution. Annual Review of Ecology and Systematics, 17, 351369.Google Scholar
Endler, J.A. (1986). Natural Selection in the Wild. Princeton University Press, Princeton.Google Scholar
Gallagher, A.J., Creel, S., Wilson, R.P., Cooke, S.J. (2017). Energy landscapes and the landscape of fear. Trends in Ecology and Evolution, 32, 8896.Google Scholar
Gaston, K.J., Chown, S.L., Calosi, P., et al. (2009). Macrophysiology: a conceptual reunification. American Naturalist, 174, 595612.Google Scholar
Gaynor, K.M., Brown, J.S., Middleton, A.D., Power, M.E., Brashares, J.S. (2019). Landscapes of fear: spatial patterns of risk perception and response. Trends in Ecology and Evolution, 34, 355368.Google Scholar
Gillson, L., Dawson, T.P., Jack, S., McGeoch, M.A. (2013). Accommodating climate change contingencies in conservation strategy. Trends in Ecology and Evolution, 28, 135142.Google Scholar
Harley, C.D.G., Connell, S.D., Doubleday, Z.A., et al. (2017). Conceptualizing ecosystem tipping points within a physiological framework. Ecology and Evolution, 7, 60356045.Google Scholar
Hawlena, D., Schmitz, O.J. (2010). Physiological stress as a fundamental mechanism linking predation to ecosystem functioning. American Naturalist, 176, 537556.CrossRefGoogle ScholarPubMed
He, Q., Bertness, M.D. (2014). Extreme stresses, niches, and positive species interactions along stress gradients. Ecology, 95, 14371443.Google Scholar
Heurtault, J., Vannier, G. (1990). Thermorésistance chez deux Pseudoscorpiones (Garypidae), l’un du désert de Namibie, l’autre de la région de Gênes (Italie). Acta Zoologica Fennica, 190, 165171.Google Scholar
Hobday, A.J., Alexander, L.V., Perkins, S.E., et al. (2016). A hierarchical approach to defining marine heatwaves. Progress in Oceanography, 141, 227238.Google Scholar
Hoffmann, A.A., Parsons, E.A. (1991). Evolutionary Genetics and Environmental Stress. Oxford University Press, Oxford.Google Scholar
Houston, J. (2006). Variability of precipitation in the Atacama Desert: its causes and hydrological impact. International Journal of Climatology, 26, 21812198.Google Scholar
Huey, R.B. (1991). Physiological consequences of habitat selection. American Naturalist, 137, S91S115.Google Scholar
Jutfelt, F., Norin, T., Ern, R., et al. (2018). Oxygen- and capacity-limited thermal tolerance: blurring ecology and physiology. Journal of Experimental Biology, 221, jeb169615.Google Scholar
Kearney, M.R., Simpson, S.J., Raubenheimer, D., Kooijman, S.A.L.M. (2012). Balancing heat, water and nutrients under environmental change: a thermodynamic niche framework. Functional Ecology, 27, 950966.Google Scholar
Kearney, M.R., McGeoch, M.A., Chown, S.L. (2019). Where do functional traits come from? The role of theory and models. Biodiversity Information Science and Standards; doi:10.3897/biss.3.37500Google Scholar
Kim, B.M., Amores, A., Kang, S., et al. (2019). Antarctic blackfin icefish genome reveals adaptations to extreme environments. Nature Ecology and Evolution, 3, 469478.Google Scholar
Krogh, A. (1929). The progress of physiology. American Journal of Physiology, 90, 243251.Google Scholar
Lauder, G.V. (1982). Historical biology and the problem of design. Journal of Theoretical Biology, 97, 5767.Google Scholar
Leibold, M.A., Chase, J.M. (2018). Metacommunity Ecology. Princeton University Press, Princeton, NJ.Google Scholar
Levine, S. (1985). A definition of stress? In:Moberg, G.P. (ed.) Animal Stress. Springer, Berlin, pp. 5169.Google Scholar
Loarie, S.R., Duffy, P.B., Hamilton, H., et al. (2009). The velocity of climate change. Nature, 462, 10521055.Google Scholar
Makarieva, A.M., Gorshkov, V.G., Li, B.-L., Chown, S.L. (2006). Size- and temperature-independence of minimum life-supporting metabolic rates. Functional Ecology, 20, 8396.Google Scholar
Marchant, D.R., Head, J.W. (2007). Antarctic dry valleys: microclimate zonation, variable geomorphic processes, and implications for assessing climate change on Mars. Icarus, 192, 187222.Google Scholar
Merino, N., Aronson, H.S., Bojanova, D.P., et al. (2019). Living at the extremes: extremophiles and the limits of life in a planetary context. Frontiers in Microbiology, 10, 780.Google Scholar
Mildrexler, D.J., Zhao, M., Running, S.W. (2011). Satellite finds highest land skin temperatures on Earth. Bulletin of the American Meteorological Society, 92, 855860.Google Scholar
Murata, F., Hayashi, T., Matsumoto, J., Asada, H. (2007). Rainfall on the Meghalaya plateau in northeastern India – one of the rainiest places in the world. Natural Hazards, 42, 391399.Google Scholar
Ortega y Gasset, J. (1932). The Revolt of the Masses. W.W. Norton, New York.Google Scholar
Peck, L. (2018). Antarctic marine biodiversity: adaptations, environments and responses to change. Oceanography and Marine Biology. An Annual Review, 56, 105236.Google Scholar
Peck, L. (2020). The ecophysiology of responding to change in polar marine benthos. In: di Prisco, G, Huiskes, A, Elster, J, Edwards, H (eds.) Life in Extreme Environments. Cambridge University Press, Cambridge.Google Scholar
Popper, K. (1945). The Open Society and Its Enemies. Volume II. The High Tide of Prophecy: Hegel, Marx and the Aftermath. Routledge & Kegan Paul, London & New York.Google Scholar
Pörtner, H.O. (2001). Climate change and temperature-dependent biogeography: oxygen limitation of thermal tolerance in animals. Naturwissenschaften, 88, 137146.Google Scholar
Pörtner, H.-O., Bock, C., Mark, F.C. (2018). Connecting to ecology: a challenge for comparative physiologists? Response to ‘Oxygen- and capacity-limited thermal tolerance: blurring ecology and physiology’. Journal of Experimental Biology, 221, jeb169615.Google Scholar
Ravaux, J., Hamel, G., Zbinden, M., et al. (2013). Thermal limit for metazoan life in question: in vivo heat tolerance of the Pompeii worm. PLoS ONE, 8, e64074.Google Scholar
Ricklefs, R.E., Wikelski, M. (2002). The physiology/life-history nexus. Trends in Ecology and Evolution, 17, 462468.Google Scholar
Sahai, R., Nyman, L.-A. (1997). The Boomerang Nebula: the coldest region of the universe? The Astrophysical Journal, 487, L155L159.Google Scholar
Scambos, T.A., Campbell, G.G., Pope, A., et al. (2018). Ultralow surface temperatures in East Antarctica from satellite thermal infrared mapping: the coldest places on earth. Geophysical Research Letters, 45, 61246133.Google Scholar
Scott, G.R., Hawkes, L.A., Frappell, P.B., et al. (2015). How bar-headed geese fly over the Himalayas. Physiology, 30, 107115.Google Scholar
Sibly, R.M., Calow, P. (1986). Physiological Ecology of Animals. An Evolutionary Approach. Blackwell Scientific Publications, Oxford.Google Scholar
Simons, A.M. (2011). Modes of response to environmental change and the elusive empirical evidence for bet hedging. Proceedings of the Royal Society B, 278, 16011609.Google Scholar
Sørensen, J.G., White, C.R., Duffy, G.A., Chown, S.L. (2018). A widespread thermodynamic effect, but maintenance of biological rates through space across life’s major domains. Proceedings of the Royal Society B, 285, 20181775.Google Scholar
Southwood, T.R.E. (1988). Tactics, strategies and templets. Oikos, 52, 318.Google Scholar
Tufto, J. (2000). The evolution of plasticity and nonplastic spatial and temporal adaptations in the presence of imperfect environmental cues. American Naturalist, 156, 121130.Google Scholar
Tufto, J. (2015). Genetic evolution, plasticity, and bet-hedging as adaptive responses to temporally autocorrelated fluctuating selection: a quantitative genetic model. Evolution, 69, 20342049.Google Scholar
Turner, J.S. (2000). The Extended Organism. The Physiology of Animal-Built Structures. Harvard University Press, Cambridge, MA.Google Scholar
Van Dover, C.L., German, C.R., Speer, K.G., Parson, L.M., Vrijenhoek, R.C. (2002). Evolution and biogeography of deep-sea vent and seep invertebrates. Science, 295, 12531257.Google Scholar
Verberk, W.C., Overgaard, J., Ern, R., et al. (2016). Does oxygen limit thermal tolerance in arthropods? A critical review of current evidence. Comparative Biochemistry and Physiology A, 192, 6478.Google Scholar
Violle, C., Reich, P.B., Pacala, S.W., Enquist, B.J., Kattge, J. (2014). The emergence and promise of functional biogeography. Proceedings of the National Academy of Sciences of the USA, 111, 1369013696.Google Scholar
Wehner, R., March, A.C., Wehner, S. (1992). Desert ants on a thermal tightrope. Nature, 357, 586587.Google Scholar
West-Eberhard, M.J. (2003). Developmental Plasticity and Evolution. Oxford University Press, New York.Google Scholar
Wu, G.C., Wright, J.C. (2015). Exceptional thermal tolerance and water resistance in the mite Paratarsotomus macropalpis (Erythracaridae) challenge prevailing explanations of physiological limits. Journal of Insect Physiology, 82, 17.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
×