Hostname: page-component-848d4c4894-r5zm4 Total loading time: 0 Render date: 2024-07-03T09:47:23.826Z Has data issue: false hasContentIssue false

Isotopic Inferences on Early Ecosystems

Published online by Cambridge University Press:  21 July 2017

Andrew H. Knoll
Affiliation:
Botanical Museum, Harvard University, Cambridge, MA 02138, USA
Donald E. Canfield
Affiliation:
Institute of Biology, Odense University, DK-5230 Odense M, Denmark
Get access

Extract

Long thought to be inaccessible to empirical inquiry, Earth's early biosphere has in recent decades become a central focus of evolutionary and paleobiological research. Knowledge of Precambrian ecosystems comes from three principal sources. The conventional fossil record consists of the compressed and permineralized remains of cyanobacteria, protists and other microorganisms (e.g., Knoll, 1996), complemented by stromatolites and oncolites, the accretionary trace fossils of microbial mat communities (Walter, 1976). Independent inferences about early evolution can be drawn from molecular phylogenies (Pace, 1997). The third principal source of information comprises biogeochemical signatures encrypted in the chemistry of ancient sedimentary rocks. Biomarker molecular fossils and distinctive isotopic compositions record the metabolic activities of organisms not necessarily preserved morphologically (Summons and Walter, 1990). In this paper, we review the inferences about early life and environments that can be drawn from the isotopic records of carbon and sulfur.

Type
Research Article
Copyright
Copyright © 1998 by The Paleontological Society 

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

Aizenshtat, Z., Stoler, A., Cohen, Y. and Nielsen, H. 1983. The Geochemical sulphur enrichment of recent organic matter by polysulfides in the Solar-Lake, p. 279288. In Bjorøy, M. et al. (eds.) Advances in Organic Geochemistry. Wiley, New York.Google Scholar
Aletkar, W., and Rajagopalan, R. 1990, Ribulose bisphosphatecarboxylase activity in halophilic Archaebacteria. Archives of Microbiology, 153: 169174.Google Scholar
Anderson, T. F., and Pratt, L. M. 1995. Isotopic evidence for the origin of organic sulfur and elemental sulfur in marine sediments, p. 378396. In Vairavamurthy, M. A. and Schoonen, M. A. A. (eds.), Geochemical Transformations of Sedimentary Sulfur. American Chemical Society, Washington DC.CrossRefGoogle Scholar
Ayala, F. J. A., Rzhetsky, A., and Ayala, F. 1998. Origin of the metazoan phyla: molecular clocks confirm paleontological estimates. Proceedings of the National Academy of Sciences, USA, 95: 606611.Google Scholar
Bartley, J. K., Pope, M., Knoll, A. H., Semikhatov, M. A., and Petrov, P. Yu. in press. A Vendian-Cambrian boundary succession from the northwestern margin of the Siberian Platform: Stratigraphy, paleontology, chemostratigraphy, and correlation. Geological Magazine.Google Scholar
Bengtson, S. and Zhao, Yue. 1997. Fossilized metazoan embryos from the earliest Cambrian. Science, 277: 16451648.Google Scholar
Berner, R. A., and Canfield, D. E. 1989. A model for atmospheric oxygen over Phanerozoic time. American Journal of Science, 289: 333361.CrossRefGoogle Scholar
Berner, R. A., and Westrich, J. T. 1985. Bioturbation and the early diagenesis of carbon and sulfur. American Journal of Science, 285: 193206.Google Scholar
Beukes, N. J., Klein, C., Kaufman, A. J., and Hayes, J. M. 1990. Carbonate petrography, kerogen distribution, and carbon and oxygen isotope variations in an Early Proterozoic transition from limestone to iron-formation deposition, Transvaal Supergroup, South Africa. Economic Geology 85, 663690.CrossRefGoogle Scholar
Blair, N. Leu, A., Muñoz, E., Olsen, J., Kwong, E., and Des Marais, D. 1985. Carbon isotopic fractionation in heterotrophic microbial metabolism. Applied and Environmental Microbiology, 50: 9961001.Google Scholar
Brasier, M. D., and Lindsay, J. F. 1998. A billion years of environmental stability and the emergence of eukaryotes: new data from northern Australia. Geology, 26: 555558.Google Scholar
Brune, D. C. 1995. Sulfur compounds as photosynthetic electron donors, p. 847870. In Blankenship, R. E., Madigan, M. T. and Bauer, C. E. (eds.). Anoxygenic Photosynthetic Bacteria. Kluwer, Dordrecht.Google Scholar
Buick, R. 1992. The antiquity of oxygenic photosynthesis: evidence from stromatolites in sulphate-deficient Archaean lakes. Science, 255: 7477.Google Scholar
Buick, R., Des Marais, D., and Knoll, A.H. 1995. Stable isotope compositions of carbonates from the Mesoproterozoic Bangemall Group, Australia: environmental variations, metamorphic effects and stratigraphic trends. Chemical Geology, 123: 153172.Google Scholar
Cameron, E. M. 1982. Sulphate and sulphate reduction in early Precambrian oceans. Nature, 296: 145148.Google Scholar
Cameron, E. M. 1983. Evidence from early Proterozoic anhydrite for sulphur isotopic partitioning in Precambrian oceans. Nature, 304: 5456.Google Scholar
Cameron, E. M., and Hattori, K. 1987. Archean sulphur cycle: evidence from sulphate minerals and isotopically fractionated sulphides in Superior Province, Canada. Chemical Geology (Isotope Geoscience Section), 65: 341358.CrossRefGoogle Scholar
Canfield, D. E. 1996. Evolution of the sulfur cycle, p. 2327. In Bottrell, S. H. (ed.), Fourth International Symposium on the Geochemistry of the Earth's Surface, 1996. Univ. of Leeds, Ilkley.Google Scholar
Canfield, D. E., and Raiswell, R. 1991. Carbonate pecipitation and dissolution: its relevance to fossil preservation, p. 411453. In Allison, P. A. and Briggs, D. E. G. (eds.), Taphonomy: Releasing the Data Locked in the Fossil record. Plenum, London.Google Scholar
Canfield, D. E., and Teske, A. 1996. Late Proterozoic rise in atmospheric oxygen concentration inferred from phylogenetic and sulphur-isotope studies. Nature, 382: 127132.CrossRefGoogle ScholarPubMed
Canfield, D. E., and Thamdrup, B. 1994. The production of 34S-depleted sulfide during bacterial disproportionation of elemental sulfur. Science, 266: 19731975.Google Scholar
Canfield, D. E., Boudreau, B. P., Mucci, A. and Gundersen, J. K. 1998. The early diagenetic formation of organic sulfur in the sediments of Mangrove Lake, Bermuda. Geochemica et Cosmochimica Acta, 62: 767781.Google Scholar
Canfield, D. E., Lyons, T. W. and Raiswell, R. 1996. A model for iron deposition to euxinic Black Sea sediments. American Journal of Science, 296: 818834.CrossRefGoogle Scholar
Canfield, D. E., Thamdrup, B., and Fleischer, S. 1998. Isotope fractionation and sulfur metabolism by pure and enrichment cultures of elemental sulfur disproportionating bacteria. Limnology and Oceanography, 43: 253264.Google Scholar
Chambers, L. A., and Trudinger, P. A. 1979. Microbiological fractionation of stable sulfur isotopes: A review and critique. Geomicrobiological Journal, 1: 249293.CrossRefGoogle Scholar
Chang, S., Des Marais, D., Mack, R., Miller, S. L., and Strathearn, G. 1983. Prebiotic organic synthesis and the origin of life, 5392. In Schopf, J. W. (ed.), Earth's Earliest Biosphere. Princeton University Press, Princeton.Google Scholar
Claypool, G. E., and Kaplan, I. R. 1974. The origin and distribution of methane in marine sediments, p. 99139. In Natural Gases in Maine Sediments. Plenum, New York.Google Scholar
Claypool, G. E., Holser, W. T., Kaplan, I. R., Sakai, H., and Zak, I. 1980. The age curves of sulfur and oxygen isotopes in marine sulfate and their mutual interpretation. Chemical Geology, 28: 199260.Google Scholar
Cloud, P. E. 1968. Atmospheric and hydrospheric evolution on the primitive Earth. Science, 160: 729736.Google Scholar
Derry, L. A., Kaufman, A. J., and Jacobsen, S.B. 1992. Sedimentary cycling and environmental change in the late Proterozoic: evidence from stable and radiogenic isotopes. Geochimica et Cosmochimica Acta, 56: 13171329.Google Scholar
Des Marais, D. J. 1994. Tectonic control of the crustal organic carbon reservoir during the Precambrian. Chemical Geology, 114: 303314.Google Scholar
Des Marais, D. J. 1997. Isotopic evolution of the biogeochemical carbon cycle during the Proterozoic Eon. Organic Geochemistry, 27: 185193.CrossRefGoogle Scholar
Des Marais, D. J., Strauss, H., Summons, R. E., and Hayes, J. M. 1992. Carbon isotopic evidence for the stepwise oxidation of the Proterozoic environment. Nature, 359: 605609.CrossRefGoogle ScholarPubMed
Devereux, R., and Stahl, D. A. 1993. Phylogeny of sulfate-reducing bacteria and a perspective for analyzing their natural communities, p. 131160. In Odom, J. M. and Singleton, R. Jr. (eds.), The Sulfate-Reducing Bacteria: Contemporary Perspectives. Springer-Verlag, Berlin.Google Scholar
Dix, G. R., Thomson, M. L., Longstaffe, F. J., and McNutt, R. H. 1995. Systematic decrease of high 13C values with burial in late Archaean (2.8 ga) diagenetic dolomite: evidence for methanogenesis from the Crixás greenstone belt, Brazil, Precambrian Research, 70: 253268.Google Scholar
Eiler, J. M., Mojzsis, S. J., and Arrhenius, G., 1997. Carbon isotope evidence for early life. Nature, 386: 665.Google Scholar
Fallick, A. E., and Hamilton, P. J. 1989. The isotopic geochemistry of ocean waters through time. Transactions of the Royal Society of Edinburgh, Earth Sciences, 80: 177181.Google Scholar
Farquhar, G. D., Ehleringer, J. R., and Hubick, K. T. 1989. Carbon isotope discrimination and photosynthesis. Annual Review of Plant Physiology and Plant Molecular Biology, 40: 503537.Google Scholar
Ferdelman, T. G., Church, T. M., and Luther, G. W. I. 1991. Sulfur enrichment of humic substances in a Delaware salt marsh sediment core. Geochimica et Cosmochimica Acta, 55: 979988.Google Scholar
Francois, R. 1987. A study of sulphur enrichment in the humic fraction of marine sediments during early diagenesis. Geochimica et Cosmochimica Acta, 51: 1727.CrossRefGoogle Scholar
Fry, B., Cox, J., Gest, H., and Hayes, J. M. 1986. Discrimination between 34S and 32S during bacterial metabolism of inorganic sulfur compounds. Journal of Bacteriology, 165: 328330.Google Scholar
Fry, B., Gest, H., and Hayes, J. M. 1984. Isotope effects associated with the anaerobic oxidation of sulfide by the puple photosynthetic bacterium Chromatium vinosum . FEMS Microbiology Letters, 22: 283287.Google Scholar
Fry, B., Giblin, A., Dornblaser, M. and Peterson, B. 1995. Stable sulfur isotopic compositions of chromium reducible sulfur in lake sediments. In Vairavamurthy, A. and Schoonen, M. A. A. (eds.), Geochemical Transformations of Sedimentary Sulfur. ACS, Washington, DC.Google Scholar
Fuchs, G. 1989. Alternative pathways of autotrophic CO2 fixation, p. 365382. In Schlegel, H. G. and Bowien, B. (eds.), Autotrophic Bacteria. Science Tech, New York.Google Scholar
Fuchs, G., and Stupperich, E. 1985. Evolution of autotrophic CO2 fixation, p. 235249. In Scleifer, K. H. and Stackenbrandt, E. (eds.), Evolution of Prokaryotes, FEMS Symposium 29. Academic Press, New York.Google Scholar
Garrels, R. M., and Lerman, A. 1981. Phanerozoic cycles of sedimentary carbon and sulfur. Proceedings of the National Academy of Sciences, USA, 78: 46524656.Google Scholar
Gelwicks, J. T., Risatti, J. B., and Hayes, J. M. 1994. Carbon isotope effects associated with aceticlastic methanogenesis. Applied and Environmental Microbiology, 60: 467472.Google Scholar
Ghent, E. D., and O'Neil, J. R. 1985. Late Precambrian marbles of unusual carbon isotopic composition, southeastern British Columbia. Canadian Journal of Earth Sciences, 22: 324329.Google Scholar
Goericke, R., Montoya, J. P., and Fry, B. 1994. Physiology of isotopic fractionation in algae and cyanobacteria, p. 187221. In Lajtha, K. and Michener, R. H. (eds.), Stable isotopes in Ecology and Environmental Science. Blackwell, Oxford.Google Scholar
Goldhaber, M. B. and Kaplan, I. R. 1980. Mechanisms of sulfur incorporation and isotope fractionation during early diagenesis in sediments of the Gulf of California. Marine Chemistry, 9: 95143.Google Scholar
Goodwin, A. M., Monster, J., and Thode, H. G. 1976. Carbon and sulfur isotope abundances in Archean iron-formations and early Precambrian life. Economic Geology, 71: 870891.Google Scholar
Habicht, K. S., and Canfield, D. E. 1996. Sulphur isotope fractionation in modern microbial mats and the evolution of the sulphur cycle. Nature, 382: 342343.Google Scholar
Habicht, K. S., and Canfield, D. E. 1997. Sulfur isotope fractionation during bacterial sulfate reduction in organic-rich sediments. Geochimica et Cosmochimica Acta, 61: 53515361.Google Scholar
Habicht, K. S., Canfield., D. E., and Rethmeier, J. in press. Sulfur isotope fractionation during bacterial reduction and disproportionation of thiosulfate and sulfite. Geochimica et Cosmochmica Acta. Google Scholar
Harrison, A. G., and Thode, H. G. 1958. Mechanisms of the bacterial reduction of sulfate from isotope fractiontion studies. Transactions of the Faraday Society, 53: 8492.Google Scholar
Hayes, J. M. 1983. Geochemical evidence bearing on the origin of aerobiosis\ a speculative hypothesis, p. 291301. In Schopf, J. W. (ed.), Earth's Earliest Biosphere. Princeton University Press, Princeton.Google Scholar
Hayes, J. M. 1993. Factors controlling 13C contents of sedimentary compounds: principles and evidence. Marine Geology, 113: 111125.Google Scholar
Hayes, J. M. 1994. Global methanotrophy at the Archean-Proterozoic transition, p. 220236 In Bengtson, S. (ed.), Early Life on Earth. Nobel Symposium 84. Columbia University Press, New York.Google Scholar
Hayes, J. M. 1996. The earliest memories of life on Earth. Nature, 384: 2122.Google Scholar
Hayes, J. M., Lambert, I. B., and Strauss, H. 1992. The sulfur-isotopic record, p. 129132. In Schopf, J. W. and Klein, C. (eds.), The Proterozoic Biosphere. Cambridge University Press, Cambridge.Google Scholar
Hayes, J. M., Takigiku, R., Ocampo, R., Callot, H. J., and Albrecht, P. 1987. Isotopic compositions and probable origins of organic molecules in the Eocene Messel shale. Nature, 329, 4851.Google Scholar
Hoffman, P. F., Kaufman, A. J., and Halverson, G. P. 1998. Comings and Goings of Global Glaciations on a Neoproterozoic Tropical Platform in Namibia. Geology, 8: 18.Google Scholar
Holland, H. D., and Beukes, N. J. 1990. A paleoweathering profile from Griqualand West, South Africa: evidence for a dramatic rise in atmospheric oxygen between 2.2 and 1.9 bybp. American Journal of Science, 290-A: 134.Google Scholar
Holland, H. D., and Karhu, J. A. 1996. Carbon isotopes and the rise of atmospheric oxygen. Geology, 24: 867870.Google Scholar
Holser, W. T., Schidlowski, M., Mackenzie, F. T., and Maynard, J. B. 1988. Geochemical cycles of carbon and sulfur, p. 105173. In Gregor, C. B., Garrels, R. M., Mackenzie, F. T. and Maynard, J. B. (eds.), Chemical Cycles in the Evolution of the Earth. Wiley, New York.Google Scholar
Irwin, H., Curtis, C., and Coleman, M. 1977. Isotopic evidence for source of diagenetic carbonates formed during burial of organic-rich sediments. Nature, 269: 209213.Google Scholar
Jørgensen, B. B. 1977. The sulfur cycle of a coastal marine sediment (Limfjorden, Denmark). Limnology and Oceanography, 22: 814832.Google Scholar
Jørgensen, B. B. 1982. Mineralization of organic matter in the sea bed –the role of sulphate reduction. Nature, 296: 643645.Google Scholar
Kah, L. C., Sherman, A. G., Narbonne, G. M., Knoll, A. H., and Kaufman, A. J. in press. 13C stratigraphy of the Proterozoic Bylot Supergroup, Baffin Island, Canada: Implications for regional stratigraphic correlations. Canadian Journal of Earth Sciences.Google Scholar
Kaplan, I.R., and Nissenbaum, A. 1966. Anomalous carbon isotope ratios in nonvolatile organic material. Science, 153: 744745.Google Scholar
Kaplan, I.R., and Rittenberg, S. C. 1964. Microbiological fractionation of sulphur isotopes. Journal of General Microbiology, 34: 195212.Google Scholar
Karhu, J.A., 1993, Paleoproterozoic evolution of the carbon isotope ratios of sedimentary carbonates in the Fennoscandian Shield. Geological Survey of Finland Bulletin, 371: 187.Google Scholar
Kaufman, A.J., 1997, An ice age in the tropics. Nature, 386: 227228.Google Scholar
Kaufman, A.J., and Knoll, A. H. 1995. Neoproterozoic variations in the C-isotopic composition of seawater: stratigraphic and biogeochemical implications. Precambrian Research, 73: 2749.CrossRefGoogle ScholarPubMed
Kaufman, A.J., Knoll, A. H., and Narbonne, G. M. 1997. Isotopes, ice ages, and Neoproterozoic Earth history. Proceedings of the National Academy of Sciences, USA, 94: 66006605.Google Scholar
Kemp, A. L. W., and Thode, H. G. 1968. The mechanism of the bacterial reduction of sulphate and of sulphite from isotope fractionation studies. Geochimica et Cosmochimica Acta, 32: 7191.Google Scholar
Kimura, H., Matsumoto, R., Kakuwa, Y., Hamdi, B., and Zibaseresht, H. 1997. the Vendian-Cambrian 13C record, North Iran: evidence for overturning of the ocean before the Cambrian explosion. Earth and Planetary Science Letters, 147: E1E7.Google Scholar
Knoll, A. H. 1991. End of the Proterozoic Eon. Scientific American, 265(4): 6473.Google Scholar
Knoll, A. H. 1992a. Biological and biogeochemical preludes to the Ediacaran radiation, p. 5384. In Lipps, J. H. and Signor, P. W. (eds.), Origin and Early Evolution of the Metazoa. Plenum, New York.Google Scholar
Knoll, A. H. 1992b. The early evolution of eukaryotes: a geological perspective. Science, 256: 622627.Google Scholar
Knoll, A. H. 1996. Archean and Proterozoic paleontology, p. 5180. In Jansonius, J. and McGregor, D. C. (eds.), Palynology: Principles and Applications. American Association of Stratigraphic Palynologists, Tulsa.Google Scholar
Knoll, A. H. and Barghoorn, E. S. 1977. Archean microfossils showing cell division from the Swaziland System, South Africa. Science, 198: 396398.Google Scholar
Knoll, A. H., Bambach, R. K., Canfield, D. E., and Grotzinger, J. P. 1996. Comparative Earth history and Late Permian mass extinction. Science, 273: 452457.Google Scholar
Knoll, A. H., Hayes, J. M., Kaufman, A. J., Swett, K., and Lambert, I. B., 1986. Secular variation in carbon isotope ratios from Upper Proterozoic successions of Svalbard and East Greenland. Nature, 321: 832838.Google Scholar
Knoll, A. H., Kaufman, A. J., and Semikhatov, S. A. 1995. The Proterozoic carbon isotope record: Mesoproterozoic carbonates from Siberia. American Journal of Science, 295: 823850.Google Scholar
Lambert, I. R., and Donnelly, T. H. 1990. The palaeoenvironmental significance of trends in sulphur isotope compositions in the Precambrian: a critical review, p. 261268 In Herbert, H. K. and Ho, S. E. (eds.), Stable Isotopes and Fluid Processes in Mineralization. University of Western Australia, Perth.Google Scholar
Lambert, I. R., Beukes, N. J., Klein, C., and Veizer, J. 1992. Proterozoic mineral deposits through time, p. 5962 In Schopf, J. W. and Klein, C. (eds), The Proterozoic Biosphere. Cambridge University Press, Cambridge, UK.Google Scholar
Lambert, I. R., Donnelly, T. H., Dunlop, J. S. R., and Groves, D. I. 1978. Stable isotope compositions of early Archean sulphate deposits of probable evaporite and volcanogenic origins. Nature, 276: 808810.Google Scholar
Logan, G. A., Hayes, J. M., Hieshima, G. B., and Summons, R. E. 1995. Terminal Proterozoic reorganization of biogeochemical cycles. Nature, 376: 5356.Google Scholar
Logan, G. A., Summons, R. E., and Hayes, J. M. 1997. An isotopic biogeochemical study of Neoproterozoic and Early Cambrian sediments from the Centralian Superbasin, Australia. Geochimica et Cosmochimica Acta, 61: 53915409.Google Scholar
Lowe, D. R. 1994. Abiological origin of desrcibed stromatolites older than 3.2 Ga. Geology, 22:387390.Google Scholar
Margulis, L., and Sagan, D. 1997. Microcosmos. University of California Press, Berkeley, 390 p.Google Scholar
McKay, C. P., and Hartman, H. 1991. Hydrogen peroxide and the evolution of oxygenic photosynthesis. Origins of Life, 21: 157164.Google Scholar
Migdisov, A. A., Cherkovskiy, S. L. and Grinenko, V. A. 1974. The effects of formation conditions on the sulfur isotopes of aquatic sediments. Geochemistry International, 10: 10281047.Google Scholar
Mojzsis, S. J., Arrhenius, G., McKeegan, K. D., Harrison, T. M., Nutman, A.P., and Friend, C. R. L. 1996. Evidence for life on Earth before 3,800 million years ago. Nature, 384: 5559.Google Scholar
Muramoto, J. A., Honjo, S., Fry, B., Hay, B. J., Howarth, R. W., and Cisne, J. L. 1991. Sulfur, iron and organic carbon fluxes in the Black Sea: sulfur isotopic evidence for origin of sulfur fluxes. Deep-Sea Research, 38: S1151S1187.Google Scholar
Nursall, J. R. 1959. Oxygen as a prerequisite to the origin of the metazoa. Nature, 183: 11701172.Google Scholar
Ohmoto, H., and Rye, R. O. 1979. Isotopes of sulfur and carbon, p. 509567. In Barnes, H. L. (ed.), Geochemistry of Hydrothermal Ore Deposits. Wiley, New York.Google Scholar
Ohmoto, H., Kakegawa, T. and Lowe, D. R. 1993. 3.4-billion-year-old biogenic pyrites from Barberton, South Africa: Sulfur isotope evidence. Science, 262: 555557.Google Scholar
Pace, N.R. 1997. A molecular view of microbial diversity and the biosphere. Science, 276: 734740.Google Scholar
Perry, E. C., and Ahmad, S. N. 1977. Carbon isotope composition of graphite and carbonate minerals from 3.8-Ae metamorphosed sediments, Isukasia, Greenland. Earth and Planetary Science Letters, 36: 280284.Google Scholar
Pflug, H. D. 1966. Structured organic remains from the Fig Tree Series of the Barberton Mountain Land. University of the Witwatersrand Economic Geology Research Unit, Information Circular, 29: 114.Google Scholar
Phillipe, H., and Adoutte, A. 1995. How reliable is our current view of eukaryotic phylogeny? European Journal of Protistology, 31: 1733.Google Scholar
Raven, J. 1994. Carbon fixation and carbon availability in marine phytoplankton. Photosynthesis Research, 39: 259273.Google Scholar
Ross, G. M., Bloch, J. D., and Krouse, H. R. 1995. Neoproterozoic strata of the southern Canadian Corillera and the isotopic evolution of seawater sulfate. Precambrian Research, 73: 7199.CrossRefGoogle Scholar
Runnegar, B., 1982, A molecular-clock date for the origin of the animal phyla. Lethaia, 14: 199205.Google Scholar
Schidlowski, M. 1993. The initiation of biological processes on Earth: Summary of the empirical evidence, p. 639655 In Engel, M. H. and Macko, S. A. (eds.), Organic Geochemistry. Plenum, New York.Google Scholar
Schidlowski, M., and Aharon, P. 1992. Carbon cycle and carbon isotopic record: geochemical impact of life over 3.8 Ga of Earth history, p. 147175. In Schidlowski, M. et al. (eds.), Early Organic Evolution: Implications for Mineral and Energy Resources. Springer-Verlag, Heidelberg.Google Scholar
Schidlowski, M., Appel, P. W. L., Eichmann, R., and Junge, C. E. 1979. Carbon isotope geochemistry of the 3.7×109-yr-old Isua Sediments, West Greenland: Implications for the Archaean carbon and oxygen cycles. Geochimica et Cosmochimica Acta, 43: 189199.Google Scholar
Schidlowski, M., Eichman, R., and Junge, C. E. 1976. Carbon isotope geochemistry of the Precambrian Lomagundi carbonate province, Rhodesia. Geochimica et Cosmochimica Acta, 40: 449455.Google Scholar
Schidlowski, M., Hayes, J. M., and Kaplan, I. R. 1983. Isotopic inferences of ancient biochemistries: carbon, sulufr, hydrogen, and nitrogen, p. 149186. In Schoipf, J. W. (ed.), Earth's Earliest Biosphere. Princeton University Press, Princeton.Google Scholar
Schoell, M., and Wellmer, F.-W. 1981. Anomalous 13C depletion in early Precambrian graphites from Superior Province, Canada. Nature, 290: 696699.Google Scholar
Schopf, J. W. (Ed.) 1983. Earth's Earliest Biosphere: Its Origin and Evolution. Princeton University Press, Princeton, N.J., 543 p.Google Scholar
Schopf, J. W. (Ed.) 1993. Microfossils of the Early Archean Apex Chert: New evidence of the antiquity of life. Science, 260: 640646.Google Scholar
Schopf, J. W., and Barghoorn, E. S. 1967. Alga-like fossils from the Early Precambrian of South Africa. Science, 156: 508512.Google Scholar
Shanks, W. C. III, and Seyfried, W. E. Jr. 1987. Stable isotope studies of vent fluids and chimney minerals, southern Juan de Fuca Ridge: sodium metasomatism and seawater sulfate reduction. Journal of Geophysical Research, 92: 11,387–11,399.Google Scholar
Sogin, M. L. 1994. The origin of eukaryotes and evolution into major kingdoms, p. 181192. In Bengtson, S. (ed.), Early Life on Earth. Nobel Symposium 84. Columbia University Press, New York.Google Scholar
Sørensen, J., Christensen, D., and Jørgensen, B. B. 1981. Volatile fatty acids and hydrogen as substrates for sulfate-reducing bacteria in anaerobic marine sediment. Applied and Environmental Microbiology, 42: 511.Google Scholar
Stackebrandt, E., Stahl, D. A., and Devereux, R. 1995. Taxonomic relationships, p. 4987. In Barton, L. L. (ed.), Sulfate-Reducing Bacteria. Plenum, New York.Google Scholar
Stetter, K. O. 1996. Hyperthermophiles in the History of Life, p. 110. In Bock, G. R. and Goode, J. A. (eds.), Evolution of Hydrothermal Ecosystems on Earth (and Mars?). Wiley, New York.Google Scholar
Strauss, H. 1993. The sulfur isotopic record of Precambrian sulfates: new data and a critical evaluation of the existing record. Precambrain Research, 63: 225246.Google Scholar
Strauss, H., and Moore, T. 1992. Abundances and isotopic compositions of carbon and sulfur species in whole rock and kerogen samples, p. 709798. In Schopf, J. W. and Klein, C. (eds.), The Proterozoic Biosphere. Cambridge University Press, Cambridge.Google Scholar
Summons, R. E., and Walter, M. R. 1990. Molecular fossils and microfossils of prokaryotes and protists from Proterozoic sediments. American Journal of Science, 290A: 212244.Google Scholar
Veizer, J., Hoefs, J., Lowe, D. R., and Thurston, P. C. 1989. Geochemistry of Precambrian carbonates: II. Archean greenstone belts and Archean sea water. Geochimica et Cosmochimica Acta, 53: 859871.Google Scholar
Walsh, M. M. 1992. Microfossils and possible microfossils from the Early Archean Onverwacht Group, Barberton Mountain Land, South Africa. Precambrian Research, 54: 271293.Google Scholar
Walter, M. R. (ed.) 1976. Stromatolites. Elsevier, Amsterdam, 790 p.Google Scholar
Wray, G. A., Levinton, J. S., and Shapiro, L. H. 1996. Molecular evidence for deep Precambrian divergences among metazoan phyla. Science, 270: 13191325.Google Scholar
Wickham, S. M., and Peters, M. T. 1993. High 13C Neoproterozoic carbonate rocks in western North America. Geology, 21: 165168.Google Scholar
Woese, C.R. 1987. Bacterial evolution. Microbiological Reviews, 51: 221271.Google Scholar
Woese, C.R., Kandler, O., and Wheeler, M. 1990. Towards a natural system of organisms: proposal for the domains Archaea, Bacteria, and Eucarya. Proceedings of the National Academy of Sciences, USA, 87: 45764579.Google Scholar
Woodruff, L. G., and Shanks, W. C. III. 1987. Sulfur isotope study of chimney minerals and vent fluids from 21°N, East Pacific Rise: hydrothermal sulfur sources and disequilibrium sulfate reduction. Journal of Geophysical Research, 93: 45624572.Google Scholar
Xiao, S.Y., Zhang, A. H. and Knoll, . 1998. Three-dimensional preservation of algae and animal embryos in a Neoproterozoic phosphorite. Nature, 391: 553558.Google Scholar