Hostname: page-component-6b989bf9dc-cvxtj Total loading time: 0 Render date: 2024-04-15T04:29:25.433Z Has data issue: false hasContentIssue false

Swansong biospheres II: the final signs of life on terrestrial planets near the end of their habitable lifetimes

Published online by Cambridge University Press:  14 January 2014

Jack T. O'Malley-James
Affiliation:
School of Physics and Astronomy, University of St. Andrews, North Haugh, St. Andrews, Fife, UK
Charles S. Cockell
Affiliation:
UK Centre for Astrobiology, School of Physics and Astronomy, James Clerk Maxwell Building, The King's Buildings, University of Edinburgh, Edinburgh, UK
Jane S. Greaves
Affiliation:
School of Physics and Astronomy, University of St. Andrews, North Haugh, St. Andrews, Fife, UK
John A. Raven*
Affiliation:
Division of Plant Sciences, University of Dundee at TJHI, The James Hutton Institute, Invergowrie, Dundee, UK

Abstract

The biosignatures of life on Earth do not remain static, but change considerably over the planet's habitable lifetime. Earth's future biosphere, much like that of the early Earth, will consist of predominantly unicellular microorganisms due to the increased hostility of environmental conditions caused by the Sun as it enters the late stage of its main sequence evolution. Building on previous work, the productivity of the biosphere is evaluated during different stages of biosphere decline between 1 and 2.8 Gyr from present. A simple atmosphere–biosphere interaction model is used to estimate the atmospheric biomarker gas abundances at each stage and to assess the likelihood of remotely detecting the presence of life in low-productivity, microbial biospheres, putting an upper limit on the lifetime of Earth's remotely detectable biosignatures. Other potential biosignatures such as leaf reflectance and cloud cover are discussed.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2014 

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

Ajtay, G.L., Ketner, P. & Duvigneaud, P. (1979). Terrestrial primary production and phytomass. In The Global Carbon Cycle. Scope 13, ed. Bolin, B.Degens, E.T.Kempe, S. & Ketner, P., pp. 129181. John Wiley and Sons Ltd, Chichester.Google Scholar
Albert, V.A., Jobson, R.W., Michael, T.P. & Taylor, D.J. (2010). The carnivorous bladderwort (Utricularia, Lentibulariaceae): a system inflates. J. Exp. Bot. 61, 59.CrossRefGoogle ScholarPubMed
Allison, S.D. (2006). Brown ground: a soil carbon analogue for the green world hypothesis? Am. Natur. 167, 619627.Google Scholar
Anten, N.P.R., Schieving, F., Medina, E., Werger, M.J.A. & Schufflen, P. (1995). Optimal leaf area indices in C3 and C4 mono- and dicotyledonous species at low and high nitrogen availability. Physiologia Plantarum 95, 541550.Google Scholar
Arakaki, M., Christin, P-A., Nyffeler, R., Lendei, A., Eggli, U., Matthew Ogburn, R., Spriggs, E., Moore, M.J. & Edwards, E.J. (2011). Contemporaneous and recent radiations of the world's major succulent plant lineages. Proc. Natl. Acad. Soc. USA 108, 83798384.Google Scholar
Bailey, J. (2007). Rainbows, polarization, and the search for habitable planets. Astrobiology 7, 320332.CrossRefGoogle ScholarPubMed
Barnes, R.O. & Goldberg, E.D. (1976). Methane production and consumption in anoxic marine sediments. Geology 4, 297300.Google Scholar
Bender, M. & Conrad, R. (1994). Microbial oxidation of methane, ammonium and carbon monoxide, and turnover of nitrous oxide and nitric oxide in soils. Biogeochemistry 27, 97112.Google Scholar
Benneke, B. & Seager, S. (2012). Atmospheric retrieval for super-earths: uniquely constraining the atmospheric composition with transmission spectroscopy. Astrophys. J. 753, 100121.CrossRefGoogle Scholar
Bernhard, A. (2012). The nitrogen cycle: processes, players, and human impact. Nature Educ. Knowl. 3, 25.Google Scholar
Blair, C.C., D'Hondt, S., Spivack, A.J. & Kingsley, R.H. (2007). Radiolytic hydrogen and microbial respiration in subsurface sediments. Astrobiology 7, 951970.CrossRefGoogle ScholarPubMed
Bohlen, S.R. (1987). Pressure–temperature–time paths and a tectonic model for the evolution of granulites. J. Geol. 95, 617632.Google Scholar
Bongers, L. (1970). Energy generation and utlization in hydrogen bacteria. J. Bacteriol. 104, 145151.Google Scholar
Boyd, E.S., Lange, R.K., Mitchell, A.C., Havig, J.R., Hamilton, T.L., Lafrenière, M.J., Shock, E.L., Peters, J.W. & Skidmore, M. (2011). Diversity, abundance, and potential activity of nitrifying and nitrate-reducing microbial assemblages in a subglacial ecosystem. Appl. Environ. Microbiol. 77, 47784787.CrossRefGoogle Scholar
Burton, M.R., Sawyer, G.M. & Granieri, D. (2013). Deep carbon emissions from volcanoes. Rev. Mineral. Geochem. 75, 323354.Google Scholar
Byrne, N. et al. . (2009). Presence and activity of anaerobic ammonium-oxidizing bacteria at deep-sea hydrothermal vents. ISME J. 3, 117123.Google Scholar
Caldeira, C. & Kasting, J.F. (1992). The life span of the biosphere revisited. Nature 360, 721–273.Google Scholar
Canfield, D.E. (2005). The early history of atmospheric oxygen: homage to Robert M. Garrels. Annu. Rev. Earth Planet. Sci. 33, 136.Google Scholar
Chen, Y., Wu, L., Boden, R., Hillebrand, A., Kumaresan, D., Moussard, H., Baciu, M., Lu, Y. & Murrell, J.C. (2009). Life without light: microbial diversity and evidence of sulfur- and ammonium-based chemolithotrophy in Movile Cave. ISME J. 3, 10931104.Google Scholar
Cnossen, I., Sanz-Forcada, J., Favata, F., Witasse, O., Zegers, T. & Arnold, N.F. (2007). Habitat of early life: solar X-ray and UV radiation at earth's surface 4–3.5 billion years ago. J. Geophys. Res. Planets 112, E02008.Google Scholar
Cockell, C.S. (1999). Life on Venus. Planet. Space Sci. 47, 14871501.Google Scholar
Cockell, C.S. & Airo, A. (2002). On the plausibility of a UV transparent biochemistry. Orig. Life Evol. B 32, 255274.CrossRefGoogle ScholarPubMed
Cockell, C.S. & Raven, J.A. (2007). Ozone and life on the Archaean Earth. Phil. Trans. R. Soc. A 365, 18891901.Google Scholar
Courtillot, V., Davaille, A., Besse, J. & Stock, J. (2003). Three distinct types of hotspots in the Earth's mantle. Earth Plan. Sci. Lett. 205, 295308.Google Scholar
Dalsgaard, T., Canfield, D.E., Petersen, J., Thamdrup, B. & Acuña-González, J. (2003). N2 production by the anammox reaction in the anoxic water column of Golfo Dulce, Costa Rica. Nature 422, 606608.Google Scholar
DeLeon-Rodriguez, N., Lathem, T.L., Rodriguez-R, L.M., Barazesh, J.M., Anderson, B.E., Beyersdorf, A.J., Ziemba, L.D., Bergin, M., Nenes, A. & Konstantinidis, K.T. (2013). Microbiome of the upper troposphere: species composition and prevalence, effects of tropical storms, and atmospheric implications. Proc. Natl. Acad. Soc. USA 110, 25752580.Google Scholar
Deng, F., Chen, J.M., Plummer, S., Chen, M. & Pise, J. (2006). Algorithm for global leaf area index retrieval using satellite imagery. IEEE Trans. Geosci. Remote Sens. 44, 22192229.Google Scholar
Des Marais, D.J., Harwitt, M.O., Jucks, K.W., Kasting, J.F., Lin, D.N.C., Lunine, J.I., Schneider, J., Seager, S., Traub, W.A. & Woolf, N.J. (2002). Remote sensing of planetary properties and biosignatures on extrasolar terrestrial planets. Astrobiology 2, 153181.Google Scholar
Domagal-Goldman, S., Meadows, V.S., Claire, M.W. & Kasting, J.F. (2011). Using biogenic sulfur gases as remotely detectable biosignatures on anoxic planets. Astrobiology 11, 419441.Google Scholar
Durieux, L., Machado, L.A.T. & Laurent, H. (2003). The impact of deforestation on cloud cover over the Amazon arc of deforestation. Remote Sens. Environ. 86, 132140.CrossRefGoogle Scholar
Ehhalt, D.H. (1974). The atmospheric cycle of methane. Tellus 26, 5870.Google Scholar
Elvidge, C.D. & Lyon, R.J.P. (1985). Influence of rock-soil spectral variation on the assessment of green biomass. Remote Sens. Environ. 17, 265269.Google Scholar
Elwood Madden, M.E., Leeman, J.R., Root, M.J. & Gainey, S. (2011). Reduced sulfur–carbon–water systems on Mars may yield shallow methane hydrate reservoirs. Planet. Space. Sci. 59, 203206.Google Scholar
Ferry, J.G. (2002). Methanogenesis biochemistry. In: Encyclopedia of Life Sciences, John Wiley & Sons, Ltd. Available online at http://onlinelibrary.wiley.com/doi/10.1038/npg.els.0000573/abstract.Google Scholar
Field, C.B., Behrenfeld, M.J., Randerson, J.T. & Falkowski, P. (1998). Primary production of the biosphere: integrating terrestrial and oceanic components. Science 281, 237240.Google Scholar
Filella, I. & Peñuelas, J. (1994). The red edge position and shape as indicators of plant chlorophyll content, biomass and hydric status. Int. J. Remote Sens. 15, 14591470.Google Scholar
Fischer, T.P. (2008). Fluxes of volatiles (H2O, CO2, N2, Cl, F) from arc volcanoes. Geochem. J. 42, 2138.CrossRefGoogle Scholar
Flynn, K.J., Stoecker, D.K., Mitra, A., Raven, J.A., Glibert, P.M., Hansen, P.J., Granéli, E. & Burkholder, J.M. (2013). A case of mistaken identification: the importance of mixotrophs and the clarification of plankton functional-classification. J. Plankton Res. 35, 311.Google Scholar
Ford, E.B., Seager, S. & Turner, E.L. (2003). A model of the temporal variability of optical light from extrasolar terrestrial planets. Astronomical Society of the Pacific Conf. Series vol. 294, pp. 639644.Google Scholar
Ganguly, S. et al. . (2012). Generating global leaf area index from landsat: algorithm formulation and demonstration. Remote Sens. Environ. 122, 185202.Google Scholar
Gantt, B., Meskhidze, N. & Kamykowski, D. (2009). A new physically-based quantification of marine isoprene and primary organic aerosol emissions. Atmos. Chem. Phys. 9, 49154927.Google Scholar
Gardner, J.P. et al. . (2006). The James Webb space telescope. Space Sci. Rev. 123, 485606.Google Scholar
Geib, S.M., Filley, T.R., Hatcher, P.G., Hoover, K., Carlson, J.E., Jimenez-Gasco, M., Nakagawa-Izumi, A., Sleighter, R.L. & Tien, M. (2008). Lignin degradation in wood-feeding insects. Proc. Natl. Acad. Soc. USA 105, 1293212937.Google Scholar
Gilmozzi, R. & Spyromilio, J. (2007). The European extremely large telescope (E-ELT). Messenger 127, 1119.Google Scholar
Girguis, P.R., Orphan, V.J., Hallam, S.J. & DeLong, E.F. (2003). Growth and methane oxidation rates of anaerobic methanotrophic archaea in a continuous-flow bioreactor. Appl. Environ. Microbiol. 69, 54725482.Google Scholar
Givnish, T.J., Burkhardt, E.L., Happel, R.E. & Weintraub, J.D. (1984). Carnivory in the bromeliad Brocchiniareducta, with a cost/benefit model for the general restriction of carnivorous plants to sunny, moist, nutrient-poor habitats. Am. Natur. 124, 479497.Google Scholar
Goldblatt, C. & Watson, A.J. (2012). The runaway greenhouse: implications for future climate change, geoengineering and planetary atmospheres. Phil. Trans. R. Soc. A 370, 41974216.Google Scholar
Goldblatt, C., Robinson, T.D., Zahnle, K.J. & Crisp, D. (2013). Low simulated radiation limit for runaway greenhouse climates. Nature Geosci. 6, 661667.Google Scholar
Grinspoon, D.H. & Bullock, M.A. (2003). Searching for evidence of past oceans on Venus. BAAS 39, 540.Google Scholar
Groombridge, B. & Jenkins, M.D. (2002). Global Biodiversity: Earth's Living Resources in the 21st Century. University of California Press, Berkeley and Los Angeles, California.Google Scholar
Halmer, M.M., Schmincke, H.-U. & Graf, H.-F. (2002). The annual volcanic gas input into the atmosphere, in particular into the stratosphere: a global data set for the past 100 years. J. Volcanol. Geoth. Res. 115, 511528.CrossRefGoogle Scholar
Hegde, S. & Kaltenegger, L. (2013). Colors of extreme exo-Earth environment. Astrobiology 13, 4756.Google Scholar
Huete, A.R., Jackson, R.D. & Post, D.F. (1985). Spectral response of a plant canopy with different soil backgrounds. Remote Sens. Environ. 17, 3753.Google Scholar
Jones, H.G. (2013). Plants and Microclimate. A Quantitative Approach to Environmental Plant Physiology, 3rd edn, Cambridge University Press, Cambridge, ISBN 978-0-521-27959-8 Paperback.Google Scholar
Kallistova, A.Y., Kevbrina, M.V., Nekrasova, V.K., Glagolev, M.V., Serebryanaya, M.I. & Nozhevnikova, A.N. (2005). Methane oxidation in landfill cover soil. Microbiology 74, 608614.Google Scholar
Kallmeyer, J., Pockalny, R., Adhokari, R.R., Smith, D.C. & D'Hondt, S. (2012). Global distribution of microbial abundance and biomass in subseafloor sediment. Proc. Natl. Acad. Soc. USA 109, 1621316216.CrossRefGoogle ScholarPubMed
Kaltenegger, L., Traub, W.A. & Jucks, K.W. (2007). Spectral evolution of an Earth-like planet. Astrophys. J. 658, 598615.Google Scholar
Kaltenegger, L. et al. . (2010). Deciphering spectral fingerprints of habitable exoplanets. Astrobiology 10, 89102.Google Scholar
Kasting, J.F. (1988). Runaway and moist greenhouse atmospheres and the evolution of Earth and Venus. Icarus 74, 472494.Google Scholar
King, A.J., Cragg, S.M., Li, Y., Dymond, J., Guille, M.J., Bowles, D.J., Bruce, N.C., Graham, I.A. & McQueen-Mason, S.J. (2010). Molecular insight into lignocellulose digestion by a marine isopod in the absence of gut microbes. Proc. Natl. Acad. Sci. USA 107, 53455350.Google Scholar
Kopparapu, R.K., Ramirez, R., Kasting, J.F., Eymet, V., Robinson, T.D., Mahadevan, S., Terrien, R.C., Domagal-Goldman, S., Meadows, V. & Deshpande, R. (2013). Habitable zones around main sequence stars: new estimates. Astrophys. J. 765, 131.CrossRefGoogle Scholar
Koropatkin, N.M., Koppenaal, D.W., Pakrasi, H.B. & Smith, T.J. (2007). The Structure of a cyanobacterial bicarbonate transport protein, CmpA*. J. Biol. Chem. 282, 26062614.Google Scholar
Lawson, P.R. et al. . (2008). Terrestrial planet finder interferometer: 2007–2008 progress and plans. ‘Terrestrial Planet Finder Interferometer: 2007–2008 progress and plans’. In Proc. SPIE 7013, Optical and Infrared Interferometry, 70132N (July 14 2008); doi:10.1117/12.786822.Google Scholar
Leake, J.R. & Cameron, D.D. (2010). Physiological ecology of mycoheterotrophy. New Phytol. 185, 601605.Google Scholar
Liu, D.Y., Ding, W.X., Jia, Z.J. & Ca, Z.C. (2011). Relation between methanogenic archaea and methane production potential in selected natural wetland ecosystems across China. Biogeosciences 8, 329338.Google Scholar
Longbottom, R. & Kolbeinsen, L. (2008). Proc. 4th Ulcos Seminar. http://www.ulcos.org/en/docs/seminars/Ref30%20-%20SP12_Longtbottom_FRAME_LKv2.pdf.Google Scholar
Lovelock, J.E. & Whitfield, M. (1982). Life span of the biosphere. Nature 296, 561563.Google Scholar
Lu, J., Vecchi, G.A. & Reichler, T. (2007). Expansion of the Hadley cell under global warming. Geophys. Res. Lett. 34, L06805.Google Scholar
Lynch, R.C., King, A.J., Farías, M.E., Sowell, P., Vitry, C. & Schmidt, S.K. (2012). The potential for microbial life in the highest elevation (>6000 m a.s.l.) mineral soils of the Atacama region. J. Geophys. Res. 117, G02028.Google Scholar
Mahli, Y. (2002). Carbon in the atmosphere and terrestrial biosphere in the 21st century. Phil. Trans. R. Soc. Lond. A 360, 29252945.Google Scholar
Maimaiti, J., Zhang, Y., Yang, J., Cen, Y-P., Layzell, D.B., Peoples, M. & Dong, Z. (2007). Isolation and characterization of hydrogen-oxidizing bacteria induced following exposure of soil to hydrogen gas and their impact on plant growth. Env. Microbiol. 9, 435444.Google Scholar
Mancinelli, R.L. (1995). The regulation of methane oxidation in soil. Appl. Environ. Microbiol. 49, 581605.Google Scholar
McMahon, S., Parnell, J. & Blamey, N.J.F. (2013). Sampling methane in basalt on Earth and Mars. Int. J. Astrobiol. 12, 113122.Google Scholar
McMillan, M. & Rushforth, S.R. (1985). The diatom flora of a steam vent of Kilauea crater, Island of Hawaii. Pacific Sci. 39, 294301.Google Scholar
Miller-Ricci, E. & Fortney, J.J. (2010). The nature of the atmosphere of the transiting super-earth GJ 1214b. Astrophys. J. Lett. 716, L74L79.Google Scholar
Möhler, O., Demott, P.J., Vali, G. & Levin, Z. (2007). Microbiology and atmospheric processes: the role of biological particles in cloud physics. Biosci. Discuss. 4, 25592591.Google Scholar
Mooney, H.A. (1972). The carbon balance of plants. Annu. Rev. Ecol. System. 3, 315346.Google Scholar
Mumma, M.J., Villanueva, G.L., Novak, R.E., Hewagama, T., Bonev, B.P., DiSanti, M.A., Avi, M., Mandell, A.M. & Smith, M.D. (2009). Strong release of methane on mars in northern summer 2003. Science 323, 10411045.Google Scholar
Nicholas Hewitt, C. & Davison, B. (1997). Field measurements of dimethyl sulphide and its oxidation products in the atmosphere. Phil. Trans. R. Soc. Lond. B 352, 183189.Google Scholar
O'Malley-James, J.T., Greaves, J.S., Raven, J.A. & Cockell, C.S. (2013). Swansong biospheres: refuges for life and novel microbial biospheres on terrestrial planets near the end of their habitable lifetimes. Int. J. Astrobiol. 12, 99112.Google Scholar
Orphan, V.J., Hinrichs, K.-U., Ussler, W., Paull, C.K., Taylor, L.T., Sylva, S.P., Hayes, J.M. III & Delong, E.F. (2001). Comparative analysis of methane-oxidizing Archaea and sulfate-reducing bacteria in anoxic marine sediments. Appl. Environ. Microbiol. 67, 19221934.Google Scholar
Oze, C., Jones, L.C., Goldsmith, J.I. & Rosenbauer, R.J. (2012). Differentiating biotic from abiotic methane genesis in hydrothermally active planetary surfaces. Proc. Natl. Acad. Sci. USA 109, 97509754.Google Scholar
Pallé, E., Osorio, M.R.Z., Barrena, R., Montañés-Rodríguez, P. & Martín, E.L. (2009). Earth's transmission spectrum from lunar eclipse observations. Nature 459, 814816.Google Scholar
Palmer, P.I., Jacob, D.J., Fiore, A.M. & Martin, R.V. (2003). Mapping isoprene emissions over North America using formaldehyde column observations from space. J. Geophys. Res. 108, 4180.Google Scholar
Pilcher, C.B. (2003). Biosignatures of early Earths. Astrobiology 3, 471486.Google Scholar
Pinto, J.P., Randall Gladstone, G. & Yung, Y.L. (1980). Photochemical production of formaldehyde in earth's primitive atmosphere. Science 210, 183185.Google Scholar
Porembski, S., Theisen, I. & Barthlott, W. (2006). Biomass allocation patterns in terrestrial, epiphytic and aquatic species of Utricularia (Lentibulariaceae). Flora – Morophol. Distrib. Fuct. Ecol. Plants 201, 477482.Google Scholar
Raven, J.A., Beardall, J., Flynn, K.J. & Maberly, S.C. (2009). Phagotrophy in the origins of photosynthesis in eukaryotes and as a complementary mode of nutrition in phototrophs: relation to Darwin's Insectivorous Plants. J. Exp. Bot. 60, 39753987.Google Scholar
Raven, J.A., Giordano, M., Beardall, J. & Maberly, S.C. (2012). Algal evolution in relation to atmospheric CO2: carboxylases, carbon-concentrating mechanisms and carbon oxidation cycles. Phil. Trans. R. Soc. B 367, 493507.Google Scholar
Reith, F. (2012). Life in the deep subsurface. Geology 39, 287288.Google Scholar
Ribas, I., Guinan, E.F., Gúdel, M. & Audard, M. (2005). Evolution of solar activity over time and effects on planetary atmospheres. I. high-energy iraddiances (1–1700 Å). Astrophys. J. 622, 680694.Google Scholar
Rice, B. (2002). Carnivorous plants: classic perspectives and new research. Biologist 49, 245249.Google Scholar
Rosenfeld, D., Rudich, Y. & Lahav, R. (2001). Desert dust suppressing precipitation: a possible desertification feedback loop. Proc. Natl. Acad. Sci. USA 98, 59755980.Google Scholar
Running, S.W., Peterson, D.L., Spanner, M.A. & Teuber, K.B. (1986). Remote sensing of coniferous forest leaf area. Ecology 67, 273276.Google Scholar
Rushby, A.J., Claire, M.W., Osborn, H. & Watson, A.J. (2013). Habitable zone lifetimes of exoplanets around main sequence stars. Astrobiology 13, 833849.Google Scholar
Sagan, C. & Morowitz, H. (1967). Life in the clouds of Venus. Nature 215, 12591260.Google Scholar
Sagan, C. & Saltpeter, E.E. (1976). Particles, environments, and possible ecologies in the Jovian atmosphere. Astrophys. J. 32, 737755.Google Scholar
Sage, R.F. & Zhu, X-G. (2011). Exploiting the engine of C4 photosynthesis. J. Exp. Bot. 62, 29893000.Google Scholar
Santosh, M. (2010). A synopsis of recent conceptual models on supercontinent tectonics in relation to mantle dynamics, life evolution and surface environment. J.Geodyn. 50, 116133.Google Scholar
Sawyer, C.N. & McCarty, P.L. (1967). Chemistry for Sanitary Engineers. McGraw-Hill Book Company Inc. pp. 446447.Google Scholar
Schmidt, S., Raven, J.A. & Paungfoo-Lonhienne, C. (2013). The mixotrophic nature of photosynthetic plants. Funct. Plant Biol. 40, 425438.Google Scholar
Schrenk, M.O., Edwards, K.J., Goodman, R.M., Hamers, R.J. & Banfield, J.F. (1998). Science 279, 15191522.Google Scholar
Schulze-Makuch, D., Grinspoon, D.H., Abbas, O., Irwin, L.N. & Bullock, M.A. (2004). A sulfur-based survival strategy for putative phototrophic life in the Venusian atmosphere. Astrobiology 4, 1118.Google Scholar
Seager, S., Schrenk, M. & Bains, W. (2012). An astrophysical view of earth-based metabolic Biosignature gases. Astrobiology 12, 6182.CrossRefGoogle ScholarPubMed
Selsis, F. (2004). The atmosphere of terrestrial exoplanets: detection and characterization. Extrasolar Planets: Today and Tomorrow, ASP Conference Series vol. 321, pp. 170182.Google Scholar
Showman, A.P., Cho, J.Y-K. & Menou, K. (2010). Atmospheric circulation of extrasolar planets. In Exoplanets, ed. Seager, S., pp. 471516. University of Arizona Press, Tucson, Arizona.Google Scholar
Siebert, L., Simkin, T. & Kimberly, P. (2010). Volcanoes of the Word, 4th edn, University of California Press, Berkley and Los Angeles, CA.Google Scholar
Smith, L.A., Jim Hendry, M., Wassenaar, L.I. & Lawrence, J. (2012). Rates of microbial elemental sulfur oxidation and 18O and 34S isotopic fractionation under varied nutrient and temperature regimes. Appl. Geochem. 27, 186196.Google Scholar
Smrekar, S.E., Stofan, E.R., Mueller, N., Treiman, A., Elkins-Tanton, L., Herbert, J., Piccioni, G. & Drossart, P. (2010). Recent hotspot volcanism on Venus from VIRTIS emissivity data. Science 328, 605608.Google Scholar
Sørensen, J. (1982). Reduction of ferric iron in anaerobic marine sediment and interaction with reduction of nitrate and sulfate. Appl. Environ. Microbiol. 43, 319323.Google Scholar
Stephenson, R.F. (1997). Historical Eclipses and Earth's Rotation. Cambridge University Press, Cambridge, UK.Google Scholar
Stevens, T.O. (1997). Subsurface Lithoautotrophic Microbial Ecosystems (SLMEs) in igneous rocks: prospects for detection. In Proc. SPIE 3111 Instruments, Methods, and Missions for the Investigation of Extraterrestrial Microorganisms, pp. 358365.Google Scholar
Straub, K.L., Benz, M., Schink, B. & Widdel, F. (1996). Anaerobic, nitrate-dependent microbial oxidation of ferrous iron. Appl. Environ. Microbiol. 62, 14581460.Google Scholar
Strous, M. et al. . (2006). Deciphering the evolution and metabolism of an anammox bacterium from a community genome. Nature 440, 790794.Google Scholar
Suzuki, I., Takeuchi, T.L., Yuthasastrakosol, T.D. & Oh, J.K. (1990). Ferrous iron, and sulfur, oxidation and ferric iron reduction activities of Thiobacillus ferrooxidans are affected by growth on ferrous iron sulfur or a sulfide ore. Appl. Environ. Microbiol. 56, 16201626.Google Scholar
Suzuki, I., Chan, C.W. & Takeuchi, T.L. (1992). Oxidation of elemental sulfur to sulfite by Thiobacillus thiooxidans cells. Appl. Environ. Microbiol. 58, 37673769.Google Scholar
Tartakovsky, B., Manuel, M.-F. & Guiot, S.R. (2005). Degradation of trichloroethylene in a coupled anaerobic–aerobic bioreactor: modeling and experiment. Biochem. Eng. J. 26, 7281.Google Scholar
Teske, A.P. (2005). The deep subsurface biosphere is alive and well. TRENDS Microbiol. 13, 402404.Google Scholar
Tie, X., Zhang, R., Brasseur, G. & Lie, W. (2002). Global NOx production by lightning. J. Atmos. Chem. 43, 6174.Google Scholar
Tinetti, G. & EChO Team (2012). EChO – exoplanet characterisation observatory. Exp. Astron. 34, 311353.Google Scholar
Tolli, J.D., Sievert, S.M. & Taylor, C.D. (2006). Unexpected diversity of bacteria capable of carbon monoxide oxidation in a coastal marine environment, and contribution of the roseobacter-associated clade to total CO oxidation. Appl. Environ. Microbiol. 72, 19661973.CrossRefGoogle Scholar
Ulloa, O., Canfield, D.E., DeLong, E.F., Letelier, R.M. & Stewart, F.J. (2012). Microbial oceanography of anoxic oxygen minimum zones. Proc. Natl. Acad. Sci. USA 109, 1599616003.Google Scholar
Walker, J.C.G. (1991). Feedback processes in the biogeochemical cycles of carbon. In Scientists on Gaia, ed. Schneider, S.H. & Boston, P.J., pp. 183190. The MIT Press, Cambridge, Massachusetts.Google Scholar
Wakefield, A.E., Gotelli, N.J., Wittman, S.E. & Ellison, A.M. (2005). Prey addition alters nutrient stoichiometry of the carnivorous plant Sarracenia purpurea. Ecology 86, 17371743.Google Scholar
Wardell, L.J., Kyle, P.R. & Chaffin, C. (2004). Carbon dioxide and carbon monoxide emission rates from an alkaline intra-plate volcano: Mt. Erebus Antarctica. J. Volcanol. Geoth. Res. 131, 109121.Google Scholar
Webster, C.R., Mahaffy, P.R., Atreya, S.K., Flesch, G.J., Christensen, L.E., Farley, K.A. & MSL Science Team (2013). Measurements of Mars Methane at Gale crater by the SAM tunable laser spectrometer on the curiosity rover. 44th Lunar and Planetary Science Conf. held March 18–22, 2013 in The Woodlands, Texas. LPI Contribution 1719: 1366.Google Scholar
Whitman, W.B., Coleman, D.C. & Wiebe, W.J. (1998). Prokaryotes: the unseen majority. Proc. Natl. Acad. Sci.USA 95, 65786583.Google Scholar
Williams, D. & Kasting, J.F. (1997). Habitable planets with high obliquities. Icarus 129, 254267.Google Scholar
Winter, K. & Smith, J.A.C. (eds) (1996). Crassulacean Acid Metabolism. Biochemistry, Ecphysiolgy and Evolution. Springer, Berlin.Google Scholar
Yoshida, M. & Santosh, M. (2011). Future supercontinent assembled in the northern hemisphere. Terra Nova 23, 333338.Google Scholar
Zahnle, K., Freedman, R.S. & Catling, D.C. (2011). Is there methane on Mars? Icarus 212, 493503.Google Scholar
Zarzycki, J., Axen, S.D., Kinney, J.N. & Kerfeld, C.A. (2013). Cyanobacterial-based approaches to improving photosynthesis in plants. J. Exp. Bot. 64, 787798.Google Scholar
Zhou, Z., Takaya, N., Nakamura, A., Yamaguchi, M., Takeo, K. & Shoun, H. (2002). Ammonia fermentation, a novel anoxic metabolism of nitrate by fungi. J. Biol. Chem 277, 18921896.Google Scholar
Zuo, Y. & Jones, R.D. (1996). Photochemical production of carbon monoxide in authentic rainwater. Geophys. Res. Lett. 23, 27692772.Google Scholar