Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-wq484 Total loading time: 0 Render date: 2024-04-26T02:01:18.393Z Has data issue: false hasContentIssue false

Science

Published online by Cambridge University Press:  08 July 2017

Andreas Losch
Affiliation:
Universität Bern, Switzerland
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2017

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

Atkins, J.F., Gesteland, R.F. & Cech, T.R. (2011).The RNA World, 3rd edn, Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.Google Scholar
Atlan, H. (1972). L'organisation biologique et la théorie de l'information, Paris: Hermann.Google Scholar
Attwater, J. & Holliger, P. (2014). A synthetic approach to abiogenesis. Nature Meth., 11, 495–8.Google Scholar
Banack, S.A., Metcalf, J.S., Jiang, L., Craighead, D., Ilag, L.L. & Cox, P.A. (2012). Cyanobacteria produce N-(2-aminoethyl) glycine, a backbone for peptide nucleic acids which may have been the first genetic molecules for life on Earth. PLoS ONE, 7(11).CrossRefGoogle ScholarPubMed
Bernal, J.D. (1951). The Physical Basis of Life, London: Routledge and Kegan Paul.Google Scholar
Berthelot, M. (1887). La synthèse chimique, 6th edn, Paris: Félix Alacan.Google Scholar
Cairns-Smith, A.G. (1966). The origin of life and the nature of the primitive gene. J. Theor. Biol., 10, 5388.CrossRefGoogle ScholarPubMed
Cairns-Smith, A.G. (1982). Genetic Takeover and the Mineral Origin of Life, Cambridge: Cambridge University Press.Google Scholar
Canguilhem, G. (1952 (1992)). Aspects du vitalisme. In: La connaissance de la vie, 2nd edn, Paris: J. Vrin.Google Scholar
Canguilhem, G. (1988). Ideology and Rationality in the History of the Life Sciences, Cambridge, Mass.: MIT Press.Google Scholar
Crick, F. (1968). The origin of the genetic code. J. Mol. Biol., 38, 367–79.CrossRefGoogle ScholarPubMed
Darwin, C. (1871). Letter to J.D. Hooker 1 February, Darwin Correspondence Project, “Letter no. 7471,” accessed on 28 September 2016, http://www.darwinproject.ac.uk/DCP-LETT-7471Google Scholar
Darwin, C. (1859 (1992)). On The Origin of Species. L'origine des espèces, trans. Barbier, Edmond, revue par Daniel Becquemont, Paris: GF- Flammarion.Google Scholar
Errington, J. (2013). L-form bacteria, cell walls and the origins of life. Open Biol., 3(1).Google Scholar
Ferris, J.P., Hill, A.R., Liu, R. & Orgel, L.E. (1996). Synthesis of long prebiotic oligomers on mineral surfaces. Nature, 381, 5961.Google Scholar
Friedmann, N., Miller, S.L. & Sanchez, R.A. (1971). Primitive Earth synthesis of nicotinic acid derivatives. Science, 171, 1026–7.Google Scholar
Ganti, T. (2003). Chemoton Theory, 2 vols, New York: Kluwer Academic/Plenum Publishers.Google Scholar
Gilbert, W. (1986). Origin of life: the RNA world. Nature, 319, 618.Google Scholar
Häring, M., Vestergaard, G., Rachel, R., Chen, L., Garrett, R.A. & Prangishvili, D. (2005). Virology: Independent virus development outside a host. Nature, 436, 1101–2.CrossRefGoogle ScholarPubMed
Hesiode, (2001). Théogonie, trans. Backès, Jean-Louis. Paris: Gallimard.Google Scholar
Kirschner, M., Gerhart, J. & Mitchison, T. (2000). Molecular vitalism. Cell, 100(1), 7988.Google Scholar
Lacan, J. (1972). In La Conférence de Louvain 1972 (extraits), ARTE France, INA. 2001.Google Scholar
Lamarck, J.B. (1809 (1994)). Philosophie zoologique, Paris: Flammarion.Google Scholar
Leaver, M., Domınguez-Cuevas, P., Coxhead, J.M., Daniel, R.A. & Errington, J. (2009). Life without a wall or division machine in Bacillus subtilis. Nature, 457, 849–53.Google Scholar
Leclerc, F., Zaccai, G., Vergne, J., Řìhovà, M., Martel, A. & Maurel, M.C. (2016). Self-assembly controls self-cleavage of HHR from ASBVd (−): a combined SANS and modeling study. Sci. Rep., 6(30287).Google Scholar
Maurel, M.C. (1994). Les origines de la vie, Paris: Syros.Google Scholar
Maurel, M.C. (1999). August Weismann et la génération spontanée de la vie, Paris: Kimé.Google Scholar
Maurel, M-C. (2002). Notion d'origines. Actes du colloque, 15 Mai 2001, MNHN, Paris: “Exobiologie, aspects historiques et épistémologiques”, Cahiers François Viète, n°4, 2002.Google Scholar
Maurel, M.C. (2003). Origines de la vie, originalité du vivant. In Maurel, M.C. & Miquel, P.A. (eds.), Nouveaux débats sur le vivant, Paris: Kimé, pp. 921.Google Scholar
Maurel, M.C. & Décout, J.L. (1999). Origins of life: molecular foundations and new approaches. Tetrahedron, 55(11), 3141–82.Google Scholar
Orgel, L.E. (1968). Evolution of the genetic apparatus. J. Mol. Biol., 38(3), 381–93.Google Scholar
Paecht-Horowitz, M., Berger, J. & Katchalsky, A. (1970). Prebiotic synthesis of polypeptides by heterogeneous polycondensation of amino acid adenylates. Nature, 228, 636.CrossRefGoogle ScholarPubMed
Pali, G., Zucchi, C. & Caglioti, L. (2002). Fundamentals of Life, Paris: Elsevier.Google Scholar
Pinheiro, V.B. & Holliger, P. (2012). The XNA world: progress towards replication and evolution of synthetic genetic polymers. Current Opinion in Chemical Biology, 16(3–4), 245–52.Google Scholar
Popa, R. (2004). Between Necessity and Probability. Searching for the Definition and Origin of Life, Berlin: Springer.Google Scholar
Popper, K. (1974 (1981)). La quête inachevée. Autobiographie intellectuelle, trans. Bouveresse, Renée, Paris: Calmann-Levy.Google Scholar
Roupnel, G. (1945). La nouvelle Siloë, Paris: Grasset 1945.Google Scholar
Szostak, J.W., Bartel, D.P. & Luisi, P.L. (2001). Synthesizing life. Nature, 409, 387–90.CrossRefGoogle ScholarPubMed
Wächtershäuser, G. (1988). Before enzymes and templates: theory of surface metabolism. Microbiological Review, 52(4), 452–84.Google Scholar
Woese, C.R. (1965). On the evolution of the genetic code. PNAS, 54(6), 1546–52.Google Scholar
Zaug, A.J., Grabowski, P.J. & Cech, T.R. (1983). Autocatalytic cyclisation of an excised intervening sequence RNA is a cleavage-ligation reaction. Nature, 301, 578–83.Google Scholar

References

Anglada-Escudé, G., Amado, P. J., Barnes, J. et al. (2016). A terrestrial planet candidate in a temperate orbit around Proxima Centauri. Nature, 536(7617), 437–40.Google Scholar
Arnold, L., Gillet, S., Lardière, O., Riaud, P., & Schneider, J. (2002). A test for the search for life on extrasolar planets – Looking for the terrestrial vegetation signature in the Earthshine spectrum. Astronomy & Astrophysics, 392(1), 231–37.Google Scholar
Barnes, R., Deitrick, R., Luger, R. et al. (2016). The habitability of Proxima Centauri b I: Evolutionary scenarios. arXiv preprint arXiv:1608.06919.Google Scholar
Baross, J., Benner, S., Cody, G. et al. (2007). The limits of organic life in planetary systems Committee on the origins and evolution of life (Vol. 38, pp. 1070): National Reseach Council.Google Scholar
Baum, S. D., Goertzel, B., & Goertzel, T. G. (2011). How long until human-level AI? Results from an expert assessment. Technological Forecasting and Social Change, 78(1), 185–95.Google Scholar
Bessel, F. (1838). On the parallax of 61 Cygni. Monthly Notices of the Royal Astronomical Society, 4, 152–61.Google Scholar
Boyajian, T., LaCourse, D., Rappaport, S. et al. (2016). Planet Hunters X. KIC 8462852 – Where's the flux? Monthly Notices of the Royal Astronomical Society, 457(4), 39884004.Google Scholar
Burke, C. J., Christiansen, J. L., Mullally, F. et al. (2015). Terrestrial planet occurrence rates for the Kepler GK Dwarf sample. The Astrophysical Journal, 809(1), 8.CrossRefGoogle Scholar
CarriganJr, R. A. (2009). IRAS-based whole-sky upper limit on Dyson spheres. The Astrophysical Journal, 698(2), 2075.Google Scholar
Cassan, A., Kubas, D., Beaulieu, J.-P. et al. (2012). One or more bound planets per Milky Way star from microlensing observations. Nature, 481(7380), 167–9.Google Scholar
Catling, D. C. (2014). The Great Oxidation Event Transition. In Turekian, H. D. H. a. K. K, ed., Treatise on Geochemistry, 2nd edn, Vol. 6, Oxford: Elsevier, pp. 177195.CrossRefGoogle Scholar
Catling, D. C. & Kasting, J. (2017). Atmospheric Evolution on Inhabited and Lifeless Worlds, Cambridge: Cambridge University Press.Google Scholar
Crossfield, I. J. (2015). Observations of exoplanet atmospheres. Publications of the Astronomical Society of the Pacific, 127(956), 941.Google Scholar
Dick, S. J. (1993). The search for extraterrestrial intelligence and the NASA High Resolution Microwave Survey (HRMS): historical perspectives. Space science reviews, 64(1–2), 93139.CrossRefGoogle Scholar
Diehl, R., Halloin, H., Kretschmer, K. et al. (2006). Radioactive 26Al from massive stars in the Galaxy. Nature, 439(7072), 45–7.CrossRefGoogle ScholarPubMed
Domagal-Goldman, S. D., Meadows, V. S., Claire, M. W., & Kasting, J. F. (2011). Using biogenic sulfur gases as remotely detectable biosignatures on anoxic planets. Astrobiology, 11(5), 419–41.Google Scholar
Ehman, J. (2010). The Big Ear Wow! signal (30th Anniversary Report). Big Ear Radio Observatory. Available online at http://www.bigear.org/Wow30th/wow30th.htm.Google Scholar
Einstein, A. (1936). Lens-like action of a star by the deviation of light in the gravitational field. Science, 84(2188), 506–7.Google Scholar
Gillon, M., Triaud, A. H., Demory, B. O. et al. (2017). Seven temperate terrestrial planets around the nearby ultracool dwarf star TRAPPIST-1. Nature, 542(7642), 456–60.Google Scholar
Harman, C., Schwieterman, E., Schottelkotte, J., & Kasting, J. (2015). Abiotic O2 levels on planets around F, G, K, and M stars: Possible false positives for life? The Astrophysical Journal, 812(2), 137.CrossRefGoogle Scholar
Hart, M. H. (1978). The evolution of the atmosphere of the Earth. Icarus, 33(1), 2339.CrossRefGoogle Scholar
Howard, A. W. (2013). Observed properties of extrasolar planets. Science, 340(6132), 572–6.Google Scholar
Hubble, E. P. (1925). Cepheids in spiral nebulae. The Observatory, 48, 139–42.Google Scholar
Jeans, J. (1942). Is there life on the other worlds? Science, 95(2476), 589–92.Google Scholar
Kasting, J. F. (2013). What caused the rise of atmospheric O2? Chemical Geology, 362, 1325.Google Scholar
Kawahara, H., & Fujii, Y. (2010). Global mapping of Earth-like exoplanets from scattered light curves. The Astrophysical Journal, 720(2), 1333.Google Scholar
Kirchner, J. W. (2003). The Gaia hypothesis: conjectures and refutations. Climatic Change, 58(1–2), 2145.CrossRefGoogle Scholar
Knoll, A. H. (2008). Cyanobacteria and Earth history. The Cyanobacteria: Molecular Biology, Genomics, and Evolution, 484.Google Scholar
Kopparapu, R. K., Ramirez, R., Kasting, J. F. et al. (2013). Habitable zones around main-sequence stars: new estimates. The Astrophysical Journal, 765(2), 131.Google Scholar
Krissansen-Totton, J., Bergsman, D. S., & Catling, D. C. (2016). On detecting biospheres from chemical thermodynamic disequilibrium in planetary atmospheres. Astrobiology, 16(1), 3967. doi: 10.1089/ast.2015.1327Google Scholar
Kuhn, J. R. & Berdyugina, S. V. (2015). Global warming as a detectable thermodynamic marker of Earth-like extrasolar civilizations: the case for a telescope like Colossus. International Journal of Astrobiology, 14(03), 401–10.CrossRefGoogle ScholarPubMed
Labeyrie, A. (1999). Snapshots of alien worlds – the future of interferometry. Science, 285(5435), 1864.Google Scholar
Lane, N. (2002). Oxygen: the Molecule that Made the World, Oxford: Oxford University Press.Google Scholar
Léger, A., Fontecave, M., Labeyrie, A. et al. (2011). Is the presence of oxygen on an exoplanet a reliable biosignature? Astrobiology, 11(4), 335–41.Google Scholar
Long, K., Obousy, R., Tziolas, A. et al. (2010). PROJECT ICARUS: Son of Daedalus, Flying Closer to Another Star. arXiv preprint arXiv:1005.3833.Google Scholar
Lovelock, J. E. (1965). A physical basis for life detection experiments. Nature, 207(997), 568–70.Google Scholar
Lovelock, J. E. (1975). Thermodynamics and the recognition of alien biospheres. Proceedings of the Royal Society of London B: Biological Sciences, 189(1095), 167–81.Google Scholar
Lovelock, J. E. & Margulis, L. (1974). Atmospheric homeostasis by and for the biosphere: the Gaia hypothesis. Tellus, 26, 12.Google Scholar
Luger, R. & Barnes, R. (2015). Extreme water loss and abiotic O2 buildup on planets throughout the habitable zones of M dwarfs. Astrobiology, 15(2), 119–43.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(7488), 307–15.Google Scholar
Marois, C., Macintosh, B., Barman, T. et al. (2008). Direct imaging of multiple planets orbiting the star HR 8799. Science, 322(5906), 1348–52.Google Scholar
Martin, A. R. (1978). Project Daedalus: The Final Report on the BIS Starship Study: British Interplanetary Soc.Google Scholar
Mayor, M. & Queloz, D. (1995). A Jupiter-mass companion to a solar-type star. Nature, 378(6555), 355–9.Google Scholar
Oliver, B. (1979). Rationale for the water hole. Communication with Extraterrestrial Intelligence,6(1–2), 71.Google Scholar
Pepe, F., Molaro, P., Cristiani, S. et al. (2014). ESPRESSO: The next European exoplanet hunter. Astronomische Nachrichten, 335(1), 820.Google Scholar
Pohorille, A. & Pratt, L. (2012). Is water the universal solvent for life? Origins of Life and Evolution of Biospheres, 42(5), 405–9.Google Scholar
Postman, M., Brown, T., Koekemoer, A. et al. (2008). Science with an 8-meter to 16-meter optical/UV space telescope. Paper presented at the SPIE Astronomical Telescopes+ Instrumentation.Google Scholar
Preus, A. (2001). Essays in Ancient Greek Philosophy VI: Before Plato (Vol. 6), SUNY Press.Google Scholar
Ribas, I., Bolmont, E., Selsis, F. et al. (2016). The habitability of Proxima Centauri b. I. Irradiation, rotation and volatile inventory from formation to the present. Astronomy & Astrophysics, 596, A111.Google Scholar
Robinson, T. D., Ennico, K., Meadows, V. S. et al. (2014). Detection of ocean glint and ozone absorption using LCROSS Earth observations. The Astrophysical Journal, 787(2), 171.CrossRefGoogle Scholar
Robinson, T. D., Meadows, V. S., & Crisp, D. (2010). Detecting oceans on extrasolar planets using the glint effect. The Astrophysical Journal Letters, 721(1), L67.Google Scholar
Rodler, F. & López-Morales, M. (2014). Feasibility studies for the detection of O2 in an Earth-like exoplanet. The Astrophysical Journal, 781(1), 54.Google Scholar
Rubin, V. C., Ford, W. K., & Thonnard, N. (1980). Rotational properties of 21 SC galaxies with a large range of luminosities and radii, from NGC 4605/R= 4kpc/to UGC 2885/R= 122 kpc. The Astrophysical Journal, 238, 471–87.Google Scholar
Sagan, C. (1994). Pale Blue Dot: A Vision of the Human Future in Space, Random House.Google Scholar
Sagan, C., Thompson, W. R., Carlson, R., Gurnett, D., & Hord, C. (1993). A search for life on Earth from the Galileo spacecraft. Nature, 365(6448), 715–21.Google Scholar
Schneider, S. (2016). Superintelligent AI and the postbiological cosmos approach. In Losch, A.. ed., What is Life? On Earth and Beyond, Cambridge: Cambridge University Press.Google Scholar
Schwieterman, E., Cockell, C., & Meadows, V. (2015). Nonphotosynthetic pigments as potential biosignatures. Astrobiology, 15(5), 341–61.Google Scholar
Seager, S., Bains, W., & Hu, R. (2013). Biosignature gases in H2-dominated atmospheres on rocky exoplanets. The Astrophysical Journal, 777(2), 95.Google Scholar
Seager, S., & Sasselov, D. (2000). Theoretical transmission spectra during extrasolar giant planet transits. The Astrophysical Journal, 537(2), 916.Google Scholar
Shapley, H. (1918). Studies based on the colors and magnitudes in stellar clusters. VII. The distances, distribution in space, and dimensions of 69 globular clusters. The Astrophysical Journal, 48, 154–81.Google Scholar
Shklovskii, I. S. & Sagan, C. (1966). Intelligent Life in the Universe, San Francisco, CA: Holden Day.Google Scholar
Shostak, S. (2015). Searching for clever life. Astrobiology, 15(11), 949–50.Google Scholar
Siemion, A. P., Benford, J., Cheng-Jin, J. et al. (2014). Searching for extraterrestrial intelligence with the Square Kilometre Array. arXiv preprint arXiv:1412.4867.Google Scholar
Silburt, A., Gaidos, E., & Wu, Y. (2015). A statistical reconstruction of the planet population around Kepler solar-type stars. The Astrophysical Journal, 799(2), 180.CrossRefGoogle Scholar
Snellen, I., de Kok, R., Birkby, J. et al. (2015). Combining high-dispersion spectroscopy with high contrast imaging: Probing rocky planets around our nearest neighbors. Astronomy & Astrophysics, 576, A59.Google Scholar
Stark, C. C., Cady, E. J., Clampin, M. et al. (2016). A direct comparison of exoEarth yields for starshades and coronagraphs. Paper presented at the SPIE Astronomical Telescopes+ Instrumentation.Google Scholar
Tarter, J. (2001). The search for extraterrestrial intelligence (SETI). Annual Review of Astronomy and Astrophysics, 39(1), 511–48.Google Scholar
Vernadsky, V. I. (1926). The Biosphere, New York: Copernicus Springer-Verlag (English translation of Vernadsky, VI, 1926).Google Scholar
Walker, J. C., Hays, P., & Kasting, J. (1981). A negative feedback mechanism for the long-term stabilization of the Earth's surface temperature. Journal of Geophysical Research, 86(C10), 9776–82.Google Scholar
Webb, S. (2015). If the Universe is Teeming with Aliens…Where is Everybody?: Fifty Solutions to the Fermi Paradox and the Problem of Extraterrestrial Life (Second Edition), Springer Science & Business Media.Google Scholar
Wright, J. T., Cartier, K., Zhao, M., Jontof-Hutter, D., & Ford, E. B. (2015). The Ĝ search for extraterrestrial civilizations with large energy supplies. IV. The signatures and information content of transiting megastructures. The Astrophysical Journal, 816(1), 17.Google Scholar

References

Allwood, A. C., Walter, M. R., Burch, I. W. & Kamber, B. S. (2007). 3.43 billion-year-old stromatolite reef from the Pilbara Craton of Western Australia: Ecosystem-scale insights to early life on Earth. Precambrian Research, 158, 198227.Google Scholar
Amy, P. S. & Haldeman, D. L. (1997). The Microbiology of the Terrestrial Deep Subsurface, Boca Raton: Lewis.Google Scholar
Awramik, S. M. & Grey, K. (2005). Stromatolites: biogenicity, biosignatures, and bioconfusion. In Hoover, R. B, Levin, G. V, Rozanov, A. Y, and Gladstone, G. R, eds., Astrobiology and Planetary Missions, Proceedings of SPIE 5906, 19.Google Scholar
Cady, S. L., Farmer, J. D., Grotzinger, J. P., Schopf, J. W. & Steele, A. (2003). Morphological biosignatures and the search for life on Mars. Astrobiology, 3, 351–68.Google Scholar
Chan, C. S., Fakra, S. C., Emerson, D., Fleming, E. J. & Edwards, K. J. (2011). Lithotrophic iron-oxidizing bacteria produce organic stalks to control mineral growth: implications for biosignature formation. The ISME Journal, 5, 717–27.Google Scholar
Claus, G. & Nagy, B. (1961). A microbiological examination of some carbonaceous chondrites. Nature, 192, 594.Google Scholar
Daubenton, L. J. M. (1782). Sur les causes qui produisent trois sortes d'herborisations dans les pierres. Memoires Acad. Royale, 667–73.Google Scholar
Garcia-Ruiz, J. M., Hyde, S. T., Carnerup, A. M. et al. (2003). Self-assembled silica-carbonate structures and detection of ancient microfossils. Science, 302, 1194–7.CrossRefGoogle ScholarPubMed
Ghiorse, W. C. & Wilson, J. T. (1988). Microbial ecology of the terrestrial subsurface. Advances in Applied Microbiology, 33, 107–72.CrossRefGoogle ScholarPubMed
Grosch, E. G. & McLoughlin, N. (2014). Reassessing the biogenicity of Earth's oldest trace fossil with implications for biosignatures in the search for early life. Proceedings of the National Academy of Sciences, 111(23), 8380–5.Google Scholar
Grotzinger, J. P. & Rothman, D. H. (1996). An abiotic model for stromatolite morphogenesis. Nature, 383, 423–5.Google Scholar
Hahn, O. (1880). Die Meteorite (Chondrite) und ihre Organismen, Tübingen: Verlag der H. Laupp'schen Buchhandlung.Google Scholar
Hofmann, B. (1990). Reduction spheroids from northern Switzerland: Mineralogy, geochemistry and genetic models. Chemical Geology, 81, 5581.Google Scholar
Hofmann, B. (1991). Mineralogy and geochemistry of reduction spheroids in red beds. Mineralogy and Petrology, 44, 107–24.Google Scholar
Hofmann, B. A. & Farmer, J. D. (2000). Filamentous fabrics in low-temperature mineral assemblages: Are they fossil biomarkers? Implications for the search for a subsurface fossil record on the early Earth and Mars. Planetary and Space Science, 48, 1077–86.CrossRefGoogle Scholar
Hofmann, B. A., Farmer, J. D., von Blanckenburg, F. & Fallick, A. E. (2008). Subsurface filamentous fabrics: An evaluation of possible modes of origins based on morphological and geochemical criteria, with implications for exopalaeontology. Astrobiology, 8, 87117.Google Scholar
Hofmann, B. A. (2011). Reduction spheroids. In Reitner, J. and Thiel, V., eds., Encyclopedia of Geobiology, Dordrecht: Springer, pp. 761–62.Google Scholar
Hofmann, H. J. (1998). Synopsis of Precambrian fossil occurrences in North America. In Lucas, S. B and St-Onge, M. R, eds., Geology of the Precambrian Superior and Grenville Provinces and Precambrian Fossils in North America, Ottawa: Geological Survey of Canada, pp. 271376.Google Scholar
Kalkowski, E. (1908). Oolith und Stromatolith im norddeutschen Buntsandstein. Zeitschrift der deutschen geologischen Gesellschaft, 60, 68125.Google Scholar
Kallmeyer, J., Pockalny, R., Adhikaria, R. R., Smith, D. C. & D'Hondt, S. (2012). Global distribution of microbial abundance and biomass in subseafloor sediment. Proceedings of the National Academy of Sciences, 109(40), 16 213–16.Google Scholar
Kretzschmar, M. (1982). Fossile Pilze in Eisen-Stromatolithen von Warstein (Rheinisches Schiefergebirge). Facies, 7, 237–60.Google Scholar
Lovley, D. R. & Chapelle, F. H. (1995). Deep subsurface microbial processes. Reviews of Geophysics, 33, 365–81.Google Scholar
Mata, S. A. & Bottjer, D. J. (2009). Development of lower Triassic wrinkle structures: Implications for the search for life on other planets. Astrobiology, 9, 895906.Google Scholar
McKay, D. S., Gibson, E. K., Thomas-Keprta, K. L. et al. (1996). Search for past life on Mars: Possible relic biogenic activity in Martian meteorite ALH84001. Science, 273, 924–30.CrossRefGoogle ScholarPubMed
McKinley, J. P., Stevens, T. O. & Westall, F. (2000). Microfossils and paleoenvironments in deep subsurface basalt samples. Geomicrobiology Journal, 17, 4354.Google Scholar
McLoughlin, N., Wilson, L. A. & Brasier, M. D. (2008). Growth of synthetic stromatolites and wrinkle structures in the absence of microbes – implications for the early fossil record. Geobiology, 6, 95105.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, 124.Google Scholar
Noffke, N., Christian, D., Wacey, D. & Hazen, R. M. (2013). Microbially induced sedimentary structures recording an ancient ecosystem in the ca. 3.48 billion-year-old Dresser formation, Pilbara, Western Australia. Astrobiology, 13, 1103–24.Google Scholar
Parnell, J., Brolly, C., Spinks, S. & Bowden, S. (2016). Metalliferous biosignatures for deep subsurface microbial activity. Origins of Life and Evolution of Biospheres, 46, 107–18.Google Scholar
Pedersen, K. (2000). Exploration of deep intraterrestrial microbial life: current perspectives. FEMS Microbiology Letters, 185, 916.CrossRefGoogle ScholarPubMed
Phelps, T. J., Raione, E. G., White, D. C. & Fliermans, C. B. (1989). Microbial activities in deep subsurface environments. Geomicrobiology Journal, 7, 7992.Google Scholar
Porada, H. & Bouougri, E. (2007). ‘Wrinkle structures' – a critical review. In Schieber, J., Bose, P. K, Eriksson, P. G. et al., eds., Atlas of Microbial Mat Features Preserved Within the Clastic Rock Record, Amsterdam: Elsevier, pp. 135–44.Google Scholar
Rasmussen, B., Fletcher, I. R., Brocks, J. J. & Kilburn, M. R. (2008). Reassessing the first appearance of eukaryotes and cyanobacteria. Nature, 455, 1101–5.Google Scholar
Razumovsky, G. (1835). Les agates mousseuses. Bulletin de la société géologique de France, 6, 165–8.Google Scholar
Reith, F. (2011). Life in the deep subsurface. Geology, 39, 287–8.Google Scholar
Riding, R. (2011). The nature of stromatolites: 3,500 million years of history and a century of research. In Reitner, J., Quéric, N.-V and Arp, G., eds., Advances in Stromatolite Geobiology, Lecture Notes in Earth Sciences 131, Berlin: Springer, pp. 2974.Google Scholar
Schumann, G., Manz, W., Reitner, J. & Lustrino, M. (2004). Ancient fungal life in north Pacific Eocene oceanic crust. Geomicrobiology Journal, 21, 241–6.Google Scholar
Spinks, S. C., Parnell, J. & Bowden, S. A. (2010). Reduction spots in the Mesoproterozoic age: implications for life in the early terrestrial record. International Journal of Astrobiology, 9, 209–16.CrossRefGoogle Scholar
Suosaari, E. P., Reid, R. P., Playford, P. E. et al. (2016). New multi-scale perspectives on the stromatolites of Shark Bay, Western Australia. Scientific Reports, 6, 20557.Google Scholar
Teske, A. P. (2005). The deep subsurface biosphere is alive and well. Trends in Microbiology, 13, 402–4.Google Scholar
Trewin, N. H. & Knoll, A. H. (1999). Preservation of Devonian chemotrophic filamentous bacteria in calcite veins. Palaios, 14, 288–94.Google Scholar
Tyler, S. A. & Barghoorn, E. S. (1954). Occurrence of structurally preserved plants in Precambrian rocks of the Canadian Shield. Science, 119, 606–8.CrossRefGoogle ScholarPubMed
Wacey, D., Kilburn, M. R., Saunders, M., Cliff, J. & Brasier, M. D. (2011). Microfossils of sulphur-metabolizing cells in 3.4-billion-year-old rocks of Western Australia. Nature Geoscience, 4, 698702.Google Scholar
Walter, M. R., McLoughlin, S., Drinnan, A. N. & Farmer, J. D. (1998). Paleontology of Devonian thermal spring deposits, Drummond Basin, Australia. Alcheringa, 22, 285314.Google Scholar
Westall, F. (1999). The nature of fossil bacteria: A guide to the search for extraterrestrial life. Journal of Geophysical Research, 104, 16 437–51.Google Scholar
Westall, F. & Folk, R. L. (2003). Exogeneous carbonaceous microstructures in early Archaean cherts and BIFs from the Isua greenstone belt: Implications for the search for life in ancient rocks. Precambrian Research, 126, 313–30.CrossRefGoogle Scholar
Westall, F., Vries, S. T. d., Nijman, W. et al. (2006). The 3.466 Ga “Kitty's Gap Chert,” an early Archean microbial ecosystem. Geological Society of America Special Paper, 405, 105–31.Google Scholar
Williams, A. J., Sumner, D. Y., Alpers, C. N., Karunatillake, S. & Hofmann, B. A. (2015). Preserved filamentous microbial biosignatures in the Brick Flat Gossan, Iron Mountain, California. Astrobiology, 15, 337668.Google Scholar

References

Bada, J. L. & Lazcano, A. (2003). Prebiotic soup: revisiting the Miller experiment. Science, 300, 745–6.CrossRefGoogle ScholarPubMed
Bohr, N. (1933). Light and life. Nature, 133, 457–9.Google Scholar
Bowler, P. J. (1988). The Non-Darwinian Revolution: Reinterpreting a Historical Myth, Baltimore, MD: The Johns Hopkins University Press.Google Scholar
Budin, I. & Szostak, J. W. (2010). Expanding roles for diverse physical phenomena during the origin of life. Annual Review of Biophysics, 39, 245–63.Google Scholar
Burke, J. G. (1991). Cosmic Debris: Meteorites in History, Berkeley, CA: California University Press.Google Scholar
Campos, L. A. (2015). Radium and the Secret of Life, Chicago, IL: Chicago University Press.Google Scholar
Carlson, E. A. (1981). Genes, Radiation, and Society: The Life and Work of H J Muller, Ithaca, NY: Cornell University Press.Google Scholar
Darwin, C. (1863). Doctrine of heterogeny and modification of species. Athenaeum, Apr. 25, 554–5.Google Scholar
Darwin, C., Barrett, P. H., Gautrey, P. J. et al. (1987). Charles Darwin's Notebooks. 1836–1844, London: British Museum (Natural History).Google Scholar
Darwin, F. (1887). The Life and Letters of Charles Darwin. Vol. III, London: J. Murray.Google Scholar
Falk, R. & Lazcano, A. (2012). The forgotten dispute: A. I. Oparin and H. J. Muller on the origin of life. History and Philosophy of the Life Sciences, 34, 373–90.Google Scholar
Fara, P. (2002). An Entertainment for Angels: Electricity in the Enlightenment, New York, NY: Columbia University Press.Google Scholar
Farley, J. (1977). The Spontaneous Generation Controversy from Descartes to Oparin, Baltimore, MD: Johns Hopkins University Press.Google Scholar
Fox-Keller, E. (2002). Making Sense of Life: Explaining Biological Development with Models, Metaphors, and Machines, Cambridge, MA: Harvard University Press.Google Scholar
Friedmann, H. C. (1997). From Friedrich Wöhler's urine to Eduard Buchner's alcohol. In Cornish-Bowden, A, ed., New Beer in an Old Bottle: Eduard Buchner and the Growth of Biochemical Knowledge, Valencia: Universitat de Valencia, pp. 67122.Google Scholar
Fry, I. (2006). The origins of research into the origins of life. Endeavour, 30, 24–8.Google Scholar
Garcia-Ruiz, J. M., Hyde, S. T., Carnerup, A. M. et al. (2003). Self-assembled silica-carbonate structures and detection of ancient microfossils. Science, 302, 1194–7.CrossRefGoogle ScholarPubMed
de Goncourt, E. & de Goncourt, J. (1904). Journal des Goncourt–Mémoires de la vie littéraire …: 1862–1865, Paris: E. Fasquelle.Google Scholar
Graham, L. R. (1972). Science and Philosophy in the Soviet Union, New York, NY: Alfred A. Knopf.Google Scholar
Haeckel, E. (1862). Die Radiolarien (Rhizopoda radiaria), Berlin: Verlag von Georg ReimerGoogle Scholar
Haeckel, E. (1876). The History of Creation: or the Development of the Earth and its Inhabitants by the Action of Natural Causes, trans. Lankester, E. R., London: Henry S. King Co.Google Scholar
Herrera, A. L. (1942). A new theory of the origins and nature of life. Science, 96, 14.Google Scholar
Huxley, J. (2010). Evolution: The Modern Synthesis: The Definitive Edition, Cambridge, MA: MIT Press.Google Scholar
Kamminga, H. (1988). Historical perspective: the problem of the origin of life in the context of developments in biology. Origins of Life and Evolution of Biospheres, 18, 110.Google Scholar
Kaufmann, S. A. (1993). The Origins of Order, New York: Oxford University PressGoogle Scholar
Lamarck, J. B. (1914). Zoological Philosophy, trans. Elliot, H, London: Macmillan and Co.Google Scholar
Lazcano, A. (2003). Hooke and generation of molds. Science, 301, 1845.Google Scholar
Lazcano, A. (2007). Prebiotic evolution and the origin of life: is a system-level understanding feasible? In: Rigoutsos, I. and Stephanopoulos, G., eds., Systems Biology, New York, NY: Oxford University Press, pp. 5778.Google Scholar
Lazcano, A. (2008). What is life? A brief historical overview. Chemistry and Biodiversity, 5, 115.Google Scholar
Lazcano, A. (2010). Historical development of origins of life. In Deamer, D. W and Szostak, J., eds., Cold Spring Harbor Perspectives in Biology: The Origins of Life, Cold Spring Harbor, NY: Cold Spring Harbor Press, pp. 116.Google Scholar
Lazcano, A. (2014). The RNA World: stepping out of the shadows. In Trueba, G., ed., Why Does Evolution Matter? The Importance of Understanding Evolution, Newcastle upon Tyne: Cambridge Scholars Publishing, pp. 101–19.Google Scholar
Lehman, I. R., Zimmerman, S.B., Adler, J. et al. (1958). Enzymatic synthesis of deoxyribonucleotic acid. V. Chemical composition of enzymatically synthesized deoxyribonucleic acid. Proceedings of the National Academy of Sciences of the USA, 44: 1191–6.Google Scholar
Lehn, J-M. (2002). Toward self-organization and complex matter. Science, 295, 2400–3.Google Scholar
Leicester, H. M. (1974). Development of Biochemical Concepts from Ancient to Modern Times, Cambridge, MA: Harvard University Press.Google Scholar
Lewontin, R. C. & Levin, B. (2007). Biology Under the Influence: Dialectical Essays on Ecology, Agriculture, and Health, New York, NY: Monthly Review Press.Google Scholar
Mann, T. (1947). Doktor Faustus, trans. Lowe-Porter, H. T, New York, NY: Alfred A. Knopf.Google Scholar
Mayr, E. (1997). This is Biology, Cambridge, MA: Harvard University Press.Google Scholar
McKay, D. S., Gibson, E. K., Jr., Thomas-Keprta, K. L. et al. (1996). Search for past life on Mars: possible relic biogenic activity in martian meteorite ALH84001. Science, 273, 924–30.Google Scholar
Miller, S. L. (1953). A production of amino acids under possible primitive Earth conditions. Science, 117, 528.Google Scholar
Montillo, R. (2013). The Lady and her Monsters: a Tale of Dissections, Real-Life Dr. Frankensteins, and the Creation of Mary Shelley's Masterpiece, New York, NY: William Morrow.Google Scholar
Morange, M. (2012). The recent evolution of the question “What is Life?”. History and Philosophy of Life Sciences, 34, 425–38.Google Scholar
Muller, H. J. (1926). The gene as the basis of life. Proceedings of the 1st International Congress of Plant Sciences, Ithaca, pp. 897921.Google Scholar
Muller, H. J. (1960). In Tax, S and Callender, C., eds., Evolution after Darwin: The University of Chicago Centennial Discussions, Panel One. The Origin of Life, Chicago, IL: The University of Chicago Press, 69105.Google Scholar
Muller, H. J. (1961). Genetic nucleic acid: Key material in the origin of life. Perspectives in Biology and Medicine, 5, 123.Google Scholar
Muller, H. J. (1966). The gene material as the initiator and the organizing basis of life. American Naturalist, 100, 493502.Google Scholar
Nagy, B. (1975). Carbonaceous Meteorites, Amsterdam: Elsevier Scientific Publishing Co.Google Scholar
Oken, L. (1805). Die Zeugung (The Creation), Bamberg-Würzburg: Joseph Anton Goebhardt.Google Scholar
Oparin, A. I. (1924). Proiskhozhedenie Zhizni (Mosckovskii Rabochii, Moscow), reprinted and translated in Bernal, J. D (1967) The Origin of Life, London: Weidenfeld and Nicolson.Google Scholar
Oparin, A. I. (1938). The Origin of Life, New York, NY: McMillan.Google Scholar
Oparin, A. I. (1961). Life: Its Nature, Origin, and Development, New York: Academic Press.Google Scholar
Oparin, A. I. (1972). The appearance of life in the Universe, in Ponnamperuma, C., ed., Exobiology, Amsterdam: North-Holland, pp. 115.Google Scholar
Orgel, L. E. (2008). The implausibility of metabolic cycles in the primitive Earth. PLoS Biology, 6, 18.Google Scholar
Peretó, J. & Catalá, J. (2007). The renaissance of synthetic biology. Biological Theory, 2, 128–30.Google Scholar
Peretó, J., Bada, J. L., & Lazcano, A. (2009). Charles Darwin and the origins of life. Origins of Life and Evolution of Biospheres, 39, 395406.Google Scholar
Perezgasga, L., Silva, E., Lazcano, A., & Negrón-Mendoza, A. (2003). Herrera's sulfocyanic theory on the origin of life: a critical reappraisal. International Journal of Astrobiology, 2, 16.Google Scholar
Saladino, R., Botta, G., Bizzarri, B. M., Di Mauro, E., & García-Ruiz, J. M. (2016). A global scale scenario for prebiotic chemistry: silica based self-assembled mineral structures and formamide. Biochemistry, 55, 2806–11.Google Scholar
Sankaran, N. (2012). How the discovery of ribozymes cast RNA in the roles of both chicken and egg in origin-of-life theories. Studies in History and Philosophy of Biological and Biomedical Sciences, 43, 741–50.Google Scholar
Soyfer, V. N. (2001). The consequences of political dictatorship for Russian science. Nature Reviews Genetics, 2, 723–9.Google Scholar
Steinbock, O., Cartwright, J. H. E., & Barge, L. M. (2016). The fertile physics of chemical gardens. Physics Today, 69(3), 4451.Google Scholar
Tagliagambe, S. (1978). Science, Philosophy, and Politics in the Soviet Union 1924–1939, Milano: Feltrinelli Editore (in Italian).Google Scholar
Urey, H. C. (1952). The Planets: their Origin and Development, Chicago, IL: University of Chicago Press.Google Scholar
Watson, J. D. & Crick, F. H. C. (1953). A structure for deoxyribose nucleic acid. Nature, 171, 737–8.Google Scholar
Wisniak, J. (2000). Jöns Jacob Berzelius a guide to the perplexed chemist. Chemistry Educator, 5, 343–50.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.

  • Science
  • Edited by Andreas Losch, Universität Bern, Switzerland
  • Book: What is Life? On Earth and Beyond
  • Online publication: 08 July 2017
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.

  • Science
  • Edited by Andreas Losch, Universität Bern, Switzerland
  • Book: What is Life? On Earth and Beyond
  • Online publication: 08 July 2017
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.

  • Science
  • Edited by Andreas Losch, Universität Bern, Switzerland
  • Book: What is Life? On Earth and Beyond
  • Online publication: 08 July 2017
Available formats
×