Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-06-12T22:31:39.441Z Has data issue: false hasContentIssue false

Late Proterozoic Transitions in Climate, Oxygen, and Tectonics, and the Rise of Complex Life

Published online by Cambridge University Press:  21 July 2017

Noah J. Planavsky
Affiliation:
Department of Geology and Geophysics, Yale University, New Haven, CT 06520 USA
Lidya G. Tarhan
Affiliation:
Department of Geology and Geophysics, Yale University, New Haven, CT 06520 USA
Eric J. Bellefroid
Affiliation:
Department of Geology and Geophysics, Yale University, New Haven, CT 06520 USA
David A. D. Evans
Affiliation:
Department of Geology and Geophysics, Yale University, New Haven, CT 06520 USA
Christopher T. Reinhard
Affiliation:
School of Earth and Atmospheric Sciences, Georgia Institute of Technology, Atlanta, GA 30332 USA
Gordon D. Love
Affiliation:
Department of Earth Sciences, University of California-Riverside, Riverside, CA 92521 USA
Timothy W. Lyons
Affiliation:
Department of Earth Sciences, University of California-Riverside, Riverside, CA 92521 USA
Get access

Abstract

The transition to the diverse and complex biosphere of the Ediacaran and early Paleozoic is the culmination of a complex history of tectonic, climate, and geochemical development. Although much of this rise occurred in the middle and late intervals of the Neoproterozoic Era (1000–541 million years ago [Ma]), the foundation for many of these developments was laid much earlier, during the latest Mesoproterozic Stenian Period (1200–1000 Ma) and early Neoproterozoic Tonian Period (1000–720 Ma). Concurrent with the development of complex ecosystems, changes in the composition, configuration, and tectonic interaction between continental plates have been proposed as major shapers of both climate and biogeochemical cycling, but there is little support in the geologic record for overriding tectonic controls. Biogeochemical evidence, however, suggests that an expansion of marine oxygen concentrations may have stabilized nutrient cycles and created more stable environmental conditions under which complex, eukaryotic life could gain a foothold and flourish. The interaction of tectonic, biogeochemical, and climate processes, as described in this paper, resulted in the establishment of habitable environments that fostered the Ediacaran and early Phanerozoic radiations of animal life and the emergence of complex, modern-style ecosystems.

Type
Research Article
Copyright
Copyright © 2015 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

Anbar, A. D., Duan, Y., Lyons, T. W., Arnold, G. L., Kendall, B., Creaser, R. A., Kaufman, A. J., Gordon, G. W., Scott, C., Garvin, J., and Buick, R. 2007. A whiff of oxygen before the Great Oxidation Event? Science, 317:19031906.Google Scholar
Anbar, A. D., and Knoll, A. H. 2002. Proterozoic ocean chemistry and evolution: A bioinorganic bridge? Science, 297:11371142.Google Scholar
Anderson, R. P., MacDonald, F. A., Tosca, N. J., Bosak, T., Bold, U., and Briggs, D. E. G. 2014. Taphonomy of eukaryotic microfossils between Cryogenian ice ages explored in the Zavkhan Terrane, southwestern Mongolia [Abstract]. Geological Society of America Abstracts with Programs 46(6):542.Google Scholar
Armstrong, R. A., Lee, C., Hedges, J. I., Honjo, S., and Wakeham, S. G. 2002. A new, mechanistic model for organic carbon fluxes in the ocean based on the quantitative association of POC with ballast minerals. Deep-Sea Research Part II-Topical Studies in Oceanography, 49:219236.Google Scholar
Arnold, G. L., Anbar, A. D., Barling, J., and Lyons, T. W. 2004. Molybdenum isotope evidence for widespread anoxia in Mid-Proterozoic oceans. Science, 304:8790.Google Scholar
Auerbach, D., Planavsky, N., Alfimova, N., Reinhard, C. T., Wang, X., and Asael, D. 2014. A terrestrial Mesoarchean–Mesoproterozoic record of atmospheric oxygen levels [Abstract]. Geological Society of America Abstracts with Programs, 46(6):520.Google Scholar
Bailey, T., McArthur, J., Prince, H., and Thirlwall, M. 2000. Dissolution methods for strontium isotope stratigraphy: whole rock analysis. Chemical Geology, 167:313319.CrossRefGoogle Scholar
Becker, T. W., Conrad, C. P., Buffett, B., and Müller, R. D. 2009. Past and present seafloor age distributions and the temporal evolution of plate tectonic heat transport. Earth and Planetary Science Letters, 278:233242.Google Scholar
Bekker, A. 2015. Questioning myths of the middle Proterozoic time [Abstract]. AGU-GAC-MAC Joint Assembly, Montreal, Quebec.Google Scholar
Berner, R. A. 1991. A model for atmospheric CO2 over Phanerozoic time. American Journal of Science, 29:339376.Google Scholar
Berner, R. A. 2004. The Phanerozoic Carbon Cycle: CO2 and O2 . Oxford University Press.Google Scholar
Berner, R. A., and Kothavala, Z. 2001. GEOCARB III: A revised model of atmospheric CO2 over Phanerozoic time. American Journal of Science, 301:182204.Google Scholar
Bjerrum, C. J., and Canfield, D. E. 2002. Ocean productivity before about 1.9 Gyr ago limited by phosphorus adsorption onto iron oxides. Nature, 417:159162.Google Scholar
Blumenberg, M., Thiel, V., Riegel, W., Kah, L. C., and Reitner, J. 2012. Biomarkers of black shales formed by microbial mats, Late Mesoproterozoic (1.1 Ga) Taoudeni Basin, Mauritania. Precambrian Research, 196–197:113127.CrossRefGoogle Scholar
Bowring, S. A., Myrow, P. M., Landing, E., Ramezani, J., Condon, D., and Hoffmann, K. 2003. Geochronological constraints on Neoproterozoic glaciations and the rise of metazoans. Geological Society of America Abstracts with Programs, 35(6):516.Google Scholar
Bradley, D. C. 2011. Secular trends in the geologic record and the supercontinent cycle. Earth-Science Reviews, 108:1633.CrossRefGoogle Scholar
Brand, U., and Veizer, J. 1980. Chemical diagenesis of a multicomponent carbonate system; 1, Trace elements. Journal of Sedimentary Research, 50:12191236.Google Scholar
Brand, U., Jiang, G., Azmy, K., Bishop, J., and Montañez, I. P. 2012. Diagenetic evaluation of a Pennsylvanian carbonate succession (Bird Spring Formation, Arrow Canyon, Nevada, U.S.A.)—1: Brachiopod and whole rock comparison. Chemical Geology, 308–309:2639.Google Scholar
Brocks, J. J., Love, G. D., Summons, R. E., Knoll, A. H., Logan, G. A., and Bowden, S. A. 2005. Biomarker evidence for green and purple sulphur bacteria in a stratified Palaeoproterozoic sea. Nature, 437:866870.Google Scholar
Brocks, J. J., and Pearson, A. 2005. Building the biomarker tree of life. Molecular Geomicrobiology, 59:233258.Google Scholar
Bryan, S. E., and Ernst, R. E. 2008. Revised definition of Large Igneous Provinces (LIPs). Earth-Science Reviews, 86:175202.Google Scholar
Buick, R. 2007. Did the Proterozoic ‘Canfield Ocean’ cause a laughing gas greenhouse? Geobiology, 5:97100.Google Scholar
Butterfield, N. J. 2000. Bangiomorpha pubescens n. gen., n. sp.: implications for the evolution of sex, multicellularity, and the Mesoproterozoic/Neoproterozoic radiation of eukaryotes. Paleobiology, 26:386404.Google Scholar
Butterfield, N. J. 2011. Animals and the invention of the Phanerozoic Earth system. Trends in Ecology & Evolution, 26:8187.Google Scholar
Byrne, B., and Goldblatt, C. 2014a. Radiative forcings for 28 potential Archean greenhouse gases. Climate of the Past Discussions, 10:20112053.Google Scholar
Byrne, B., and Goldblatt, C. 2014b. Radiative forcing at high concentrations of well-mixed greenhouse gases. Geophysical Research Letters, 41:152160.Google Scholar
Canfield, D. E. 1998. A new model for Proterozoic ocean chemistry. Nature, 396:450453.Google Scholar
Canfield, D. E. 2005. The early history of atmospheric oxygen: Homage to Robert M. Garrels: Annual Review of Earth and Planetary Sciences, 33:136.Google Scholar
Canfield, D. E., Poulton, S. W., Knoll, A. H., Narbonne, G. M., Ross, G., Goldberg, T., and Strauss, H. 2008. Ferruginous conditions dominated later Neoproterozoic deep-water chemistry. Science, 321:949952.CrossRefGoogle ScholarPubMed
Canfield, D. E., Poulton, S. W., and Narbonne, G. M. 2007. Late-Neoproterozoic deep-ocean oxygenation and the rise of animal life. Science, 315:9295.Google Scholar
Canfield, D. E., and Raiswell, R. 1999. The evolution of the sulfur cycle. American Journal of Science, 299:697723.Google Scholar
Cavalier-Smith, T. 2002. The neomuran origin of Archaebacteria, the negibacterial root of the universal tree and bacterial megaclassification. International Journal of Systematic and Evolutionary Microbiology, 52:776.Google Scholar
Cawood, P. A., Hawkesworth, C. J., and Dhuime, B. 2012. Detrital zircon record and tectonic setting. Geology, 40:875878.Google Scholar
Charnay, B., Forget, F., Wordsworth, R., Leconte, J., Millour, E., Codron, F., and Spiga, A. 2013. Exploring the faint young Sun problem and the possible climates of the Archean Earth with a 3-D GCM. Journal of Geophysical Research: Atmospheres, 118:10,41410,431.Google Scholar
Clites, E. C., Droser, M. L., and Gehling, J. G. 2012. The advent of hard-part structural support among the Ediacara biota: Ediacaran harbinger of a Cambrian mode of body construction. Geology, 40:307310.Google Scholar
Cohen, P. A., and Knoll, A. H. 2012. Scale microfossils from the mid-Neoproterozoic Fifteenmile Group, Yukon Territory. Journal of Paleontology, 86:775800.Google Scholar
Cole, D., Hodgskiss, M., Gueguen, B., Kunzmann, M., Crockford, P., Gibson, T., Worndle, S., Halverson, G., and Planavsky, N. 2015. Redox conditions in the latest Mesoproterozoic [Abstract]. Goldschmidt Annual Meeting. Prague, Czech Republic. Google Scholar
Dasgupta, R. 2013. Ingassing, storage, and outgassing of terrestrial carbon through geologic time. Reviews in Mineralogy and Geochemistry, 75:183229.Google Scholar
Derry, L. A. 2010. A burial diagenesis origin for the Ediacaran Shuram–Wonoka carbon isotope anomaly. Earth and Planetary Science Letters, 294(1–2): 152162.Google Scholar
Dessert, C., Dupré, B., Gaillardet, J., François, L. M., and Allègre, C. J. 2003. Basalt weathering laws and the impact of basalt weathering on the global carbon cycle. Chemical Geology, 202(3–4): 257273.Google Scholar
Dilling, L., and Alldredge, A. L. 2000. Fragmentation of marine snow by swimming macrozooplankton: A new process impacting carbon cycling in the sea. Deep-Sea Research Part I-Oceanographic Research Papers, 47:12271245.Google Scholar
Donnadieu, Y., Godderis, Y., Ramstein, G., Nedelec, A., and Meert, J. 2004. A “Snowball Earth” climate triggered by continental break-up through changes in runoff. Nature, 428:303306.Google Scholar
Douzery, E. J. P., Snell, E. A., Bapteste, E., Delsuc, F., and Philippe, H. 2004. The timing of eukaryotic evolution: Does a relaxed molecular clock reconcile proteins and fossils? Proceedings of the National Academy of Sciences of the United States of America, 101:1538615391.Google Scholar
Driese, S. G., and Medaris, L. G. 2008. Evidence for biological and hydrological controls on the development of a Paleoproterozoic paleoweathering profile in the Baraboo Range, Wisconsin, USA. Journal of Sedimentary Research, 78:443457.Google Scholar
Driese, S. G., Simpson, E. L., and Eriksson, K. A. 1995. Redoximorphic paleosols in alluvial and lacustrine deposits, 1.8 Ga Lochness Formation, Mount-Isa, Australia—pedogenic processes and implications for paleoclimate. Journal of Sedimentary Research, 65:675689.Google Scholar
Droser, M. L., Gehling, J. G., and Jensen, S. R. 2006. Assemblage palaeoecology of the Ediacara biota: The unabridged edition? Palaeogeography, Palaeoclimatology, Palaeoecology, 232(2–4): 131147.Google Scholar
Ducea, M. N., Saleeby, J. B., and Bergantz, G. 2015. The architecture, chemistry, and evolution of continental magmatic arcs. Annual Review of Earth and Planetary Sciences, 43:299331.Google Scholar
Edmond, J. 1992. Himalayan tectonics, weathering processes, and the strontium isotope record in marine limestones. Science, 258:15941597.Google Scholar
Ellis, A., Johnson, T. M., and Bullen, T. D. 2002. Chromium isotopes and the fate of hexavalent chromium in the environment. Science, 295:20602062.CrossRefGoogle ScholarPubMed
Ellis, A., Johnson, T. M., and Bullen, T. D. 2004. Using chromium stable isotope ratios to quantify Cr(VI) reduction: Lack of sorption effects. Environmental Science & Technology, 38:36043607.Google Scholar
Emerson, S. R., and Huested, S. S. 1991. Ocean anoxia and the concentrations of molybdenum and vanadium in seawater. Marine Chemistry, 34(3–4): 177196.Google Scholar
Endal, A. S., and Schatten, K. H. 1982. The faint young sun-climate paradox: Continental influences. Journal of Geophysical Research: Oceans (1978–2012), 87(C9):72957302.CrossRefGoogle Scholar
Ernst, R. E., Wingate, M. T. D., Buchan, K. L., and Li, Z. X. 2008. Global record of 1600–700 Ma large igneous provinces (LIPs): Implications for the reconstruction of the proposed Nuna (Columbia) and Rodinia supercontinents. Precambrian Research, 160(1–2): 159178.Google Scholar
Ernst, R. E., Bleeker, W., Söderlund, U., and Kerr, A. C. 2013. Large Igneous Provinces and supercontinents: Toward completing the plate tectonic revolution. Lithos, 174:114.Google Scholar
Erwin, D. H., Laflamme, M., Tweedt, S. M., Sperling, E. A., Pisani, D., and Peterson, K. J. 2011. The Cambrian conundrum: Early divergence and later ecological success in the early history of animals. Science, 334:10911097.Google Scholar
Evans, D. A. D. 2003. A fundamental Precambrian–Phanerozoic shift in earth's glacial style? Tectonophysics, 375(1–4): 353385.Google Scholar
Evans, D. A., Beukes, N. J., and Kirschvink, J. L. 1997. Low-latitude glaciation in the Palaeoproterozoic Era. Nature, 386:262266.Google Scholar
Evans, D. A. D. 2013. Reconstructing pre-Pangean supercontinents. Geological Society of America Bulletin, 125(11–12): 17351751.Google Scholar
Evans, D. A. D., and Mitchell, R. N. 2011. Assembly and breakup of the core of Paleoproterozoic–Mesoproterozoic supercontinent Nuna. Geology, 39:443446.Google Scholar
Feely, R. A., Trefry, J. H., Lebon, G. T. and German, C. R. 1998. The relationship between P/Fe and V/Fe ratios in hydrothermal precipitates and dissolved phosphate in seawater. Geophysical Research Letters, 25:22532256.Google Scholar
Fendorf, S. E. 1995. Surface reactions of chromium in soils and waters. Geoderma, 67:5571.Google Scholar
Fennel, K., Follows, M., and Falkowski, P. G. 2005. The co-evolution of the nitrogen, carbon and oxygen cycles in the Proterozoic ocean. American Journal of Science, 305:526545.Google Scholar
Fischer, G., and Karakas, G. 2009. Sinking rates and ballast composition of particles in the Atlantic Ocean: Implications for the organic carbon fluxes to the deep ocean. Biogeosciences, 6:85102.Google Scholar
Fischer, G., Karakas, G., Blaas, M., Ratmeyer, V., Nowald, N., Schlitzer, R., Helmke, P., Davenport, R., Donner, B., Neuer, S., and Wefer, G. 2009. Mineral ballast and particle settling rates in the coastal upwelling system off NW Africa and the South Atlantic. International Journal of Earth Sciences, 98:281298.Google Scholar
Fralick, P., and Carter, J. E. 2011. Neoarchean deep marine paleotemperature: Evidence from turbidite successions. Precambrian Research, 191 (1–2), 7884.Google Scholar
Frei, R., Gaucher, C., Poulton, S. W., and Canfield, D. E. 2009. Fluctuations in Precambrian atmospheric oxygenation recorded by chromium isotopes. Nature, 461:250–225.Google Scholar
French, K. L., Hallmann, C., Hope, J. M., Schoon, P. L., Zumberge, J. A., Hoshino, Y., Peters, C. A., George, S. C., Love, G. D., Brocks, J. J., Buick, R., and Summons, R. E. 2015. Reappraisal of hydrocarbon biomarkers in Archean rocks. Proceedings of the National Academy of Sciences, 112:59155920.Google Scholar
Geboy, N. J., Kaufman, A. J., Walker, R. J., Misi, A., De Oliviera, T. F., Miller, K. E., Azmy, K., Kendall, B., and Poulton, S. W. 2013. Re–Os age constraints and new observations of Proterozoic glacial deposits in the Vazante Group, Brazil. Precambrian Research, 238:199213.Google Scholar
Goddéris, Y., Donnadieu, Y., Nédélec, A., Dupré, B., Dessert, C., Grard, A., Ramstein, G., and François, L. M. 2003. The Sturtian ‘Snowball’ glaciation: Fire and ice. Earth and Planetary Science Letters, 211(1–2): 112.Google Scholar
Golonka, J., and Krobicki, M. 2012. Upwelling regime in the Carpathian Tethys: a Jurassic–Cretaceous palaeogeographic and paleoclimatic perspective. Geological Quarterly, 45:1532.Google Scholar
Gomes, M. L., and Hurtgen, M. T. 2015. Sulfur isotope fractionation in modern euxinic systems: Implications for paleoenvironmental reconstructions of paired sulfate-sulfide isotope records. Geochimica et Cosmochimica Acta, 157:3955.Google Scholar
Gough, D. O. 1981. Solar interior structure and luminosity variations, p. 2134 In Domingo, V. (ed.), Physics of Solar Variations. Springer Netherlands.Google Scholar
Graham, A. M. and Bouwer, E. J. 2010. Rates of hexavalent chromium reduction in anoxic estuarine sediments: pH effects and the role of acid volatile sulfides. Environmental Science and Technology, 44:136142.Google Scholar
Grantham, P. J., and Wakefield, L. L. 1988. Variations in the sterane carbon number distributions of marine source rock derived crude oils through geological time. Organic Geochemistry, 12:6173.Google Scholar
Grey, K., and Calver, C. R. 2007. Correlating the Ediacaran of Australia, p. 115135 In Vickers-Rich, P. and Komarower, P. (eds.), Rise and Fall of the Ediacaran Biota. Geological Society of London Special Publication, 286.Google Scholar
Grey, K., and Williams, I. R. 1990. Problematic bedding-plane markings from the middle Proterozoic Manganese Subgroup, Bangemall Basin, Western Australia. Precambrian Research, 46:307327.Google Scholar
Grosjean, E., Love, G. D., Stalvies, C., Fike, D. A., and Summons, R. E. 2009. Origin of petroleum in the Neoproterozoic–Cambrian South Oman Salt Basin. Organic Geochemistry, 40:87110.Google Scholar
Gurnis, M. 1988. Large-scale mantle convection and the aggregation and dispersal of supercontinents. Nature, 332:695699.Google Scholar
Halverson, G. P., Dudás, F. Ö., Maloof, A. C., and Bowring, S. A. 2007. Evolution of the 87Sr/86Sr composition of Neoproterozoic seawater. Palaeogeography, Palaeoclimatology, Palaeoecology, 256(3–4): 103129.Google Scholar
Halverson, G. P., Wade, B. P., Hurtgen, M. T., and Barovich, K. M. 2010. Neoproterozoic chemostratigraphy. Precambrian Research, 182:337350.Google Scholar
Hastings, D. W., Emerson, S. R. and Mix, A. C. 1996. Vanadium in foraminiferal calcite as a tracer for changes in the areal extent of reducing sediments. Paleoceanography, 11:665678.Google Scholar
He, Y. H., Zhao, G. C., Sun, M., and Xia, X. P. 2009. SHRIMP and LA-ICP-MS zircon geochronology of the Xiong'er volcanic rocks: Implications for the Paleo-Mesoproterozoic evolution of the southern margin of the North China Craton. Precambrian Research, 168(3–4): 213222.Google Scholar
Hedges, S. B., Blair, J. E., Venturi, M. L., and Shoe, J. L. 2004. A molecular timescale of eukaryote evolution and the rise of complex multicellular life. BMC Evolutionary Biology, 4:2.Google Scholar
Hoffman, P. F., Kaufman, A. J., Halverson, G. P., and Schrag, D. P. 1998. A Neoproterozoic Snowball Earth. Science, 281:13421346.Google Scholar
Hoffman, P. F., and Schrag, D. P. 2002. The Snowball Earth hypothesis: testing the limits of global change. Terra Nova, 14:129155.Google Scholar
Holland, H. D. 1973. The oceans: A possible source of iron in iron-formations: Economic Geology, 68:11691172.Google Scholar
Hood, A. V. S., and Wallace, M. W. 2015. Extreme ocean anoxia during the Late Cryogenian recorded in reefal carbonates of Southern Australia. Precambrian Research, 261:96111.Google Scholar
Horton, F. 2015. Did phosphorus derived from the weathering of large igneous provinces fertilize the Neoproterozoic ocean? Geochemistry, Geophysics, Geosystems, 16:15252027.Google Scholar
Howarth, R. W. 1988. Nutrient limitation of net primary production in marine ecosystems. Annual Review of Ecology and Systematics, 19:89110.Google Scholar
Husson, J. M., Maloof, A. C., and Schoene, B. 2012. A syn-depositional age for Earth's deepest 13C excursion required by isotope conglomerate tests. Terra Nova, 24:318325.Google Scholar
Husson, J. M., Maloof, A. C., Schoene, B., Chen, C.Y., and Higgins, J. A. 2015. Stratigraphic expression of Earth's deepest d13C excursion in the Wonoka Formation of South Australia. American Journal of Science, 315:145.Google Scholar
Iversen, M. H., and Poulsen, L. K. 2007. Coprorhexy, coprophagy, and coprochaly in the copepods Calanus helgolandicus, Pseudocalanus elongatus, and Oithona similis . Marine Ecology Progress Series, 350:7989.Google Scholar
Javaux, E. J. 2007. The early eukaryotic fossil record. Eukaryotic Membranes and Cytoskeleton. Origins and Evolution, 607:119.Google Scholar
Javaux, E. J., Knoll, A. H., and Walter, M. R. 2001. Morphological and ecological complexity in early eukaryotic ecosystems. Nature, 412:6669.CrossRefGoogle ScholarPubMed
Javaux, E. J., Knoll, A. H., and Walter, M. R. 2004. TEM evidence for eukaryotic diversity in mid-Proterozoic oceans. Geobiology, 2:121132.Google Scholar
Javaux, E. J., Marshall, C. P., and Bekker, A. 2010. Organic-walled microfossils in 3.2-billion-year-old shallow-marine siliciclastic deposits. Nature, 463:934–8.Google Scholar
Jensen, S. 2003. The Proterozoic and earliest Cambrian trace fossil record; Patterns, problems and perspectives. Integrative and Comparative Biology, 43:219228.Google Scholar
Jensen, S., Droser, M. L., and Gehling, J. G. 2006. A critical look at the Ediacaran trace fossil record, p. 115157 In Xiao, S. and Kaufman, A. J. (eds.), Neoproterozoic Geobiology and Paleobiology, Springer.Google Scholar
Johnson, T. M., and Bullen, T. D. 2004. Mass-dependent fractionation of selenium and chromium isotopes in low-temperature environments, p. 289317 In Johnson, C. M., Beard, B. L., and Albarede, F. (eds.), Geochemistry of Non-Traditional Stable Isotopes. Mineralogical Society of America, Chantilly, VA.Google Scholar
Johnston, D. T. 2011. Multiple sulfur isotopes and the evolution of Earth's surface sulfur cycle. Earth-Science Reviews, 106(1–2): 161183.Google Scholar
Johnston, D. T., Farquhar, J., and Canfield, D. E. 2007. Sulfur isotope insights into microbial sulfate reduction: When microbes meet models. Geochimica et Cosmochimica Acta, 71:39293947.Google Scholar
Johnston, D. T., Poulton, S. W., Dehler, C., Porter, S., Husson, J., Canfield, D. E., and Knoll, A. H. 2010. An emerging picture of Neoproterozoic ocean chemistry: Insights from the Chuar Group, Grand Canyon, USA. Earth and Planetary Science Letters, 290(1–2):6473.Google Scholar
Jones, C., Nomosatryo, S., Crowe, S. A., Bjerrum, C. J., and Canfield, D. E. 2015. Iron oxides, divalent cations, silica, and the early Earth phosphorus crisis. Geology, 43:135138.Google Scholar
Kah, L. C., Crawford, D. C., Bartley, J. K., Kozlov, V. I., Sergeeva, N. D., and Puchkov, V. N. 2007. C-and Sr-isotope chemostratigraphy as a tool for verifying age of Riphean deposits in the Kama-Belaya aulacogen, the east European platform. Stratigraphy and Geological Correlation, 15:1229.Google Scholar
Kah, L. C., Sherman, A. G., Narbonne, G. M., Knoll, A. H., and Kaufman, A. J. 1999. δ13C stratigraphy of the Proterozoic Bylot Supergroup, Baffin Island, Canada: Implications for regional lithostratigraphic correlations. Canadian Journal of Earth Sciences, 36:313332.Google Scholar
Karhu, J. A., and Epstein, S. 1986. The implication of the oxygen isotope records in coexisting cherts and phosphates. Geochimica et Cosmochimica Acta, 50:17451756.Google Scholar
Karhu, J. A., and Holland, H. D. 1996. Carbon isotopes and the rise of atmospheric oxygen. Geology, 24:867870.Google Scholar
Kasting, J. F. 2005. Methane and climate during the Precambrian era. Precambrian Research, 137(3–4): 119129.Google Scholar
Kasting, J. F., Zahnle, K. J., Pinto, J. P., and Young, A. T. 1989. Sulfur, ultraviolet radiation, and the early evolution of life. Origins of Life and Evolution of the Biosphere, 19:95108.Google Scholar
Katsev, S. and Crowe, S.A., 2105, Organic carbon burial efficiencies in sediments: The power law of mineralization revisited. Geology, 43:607610.Google Scholar
Kelemen, P. B., and Manning, C. E. 2015. Reevaluating carbon fluxes in subduction zones, what goes down, mostly comes up. Proceedings of the National Academy of Sciences, doi: 10.1073/pnas.1507889112.Google Scholar
Kendall, B., Creaser, R. A., Gordon, G. W., and Anbar, A. D. 2009. Re–Os and Mo isotope systematics of black shales from the Middle Proterozoic Velkerri and Wollogorang Formations, McArthur Basin, northern Australia. Geochimica et Cosmochimica Acta, 73:25342558.Google Scholar
Kirschvink, J. 1992. Late Proterozoic low-latitude global glaciation: the Snowball Earth, p. 5152 In Schopf, J. W. and Klein, C. (eds.), The Proterozoic Biosphere: A Multidisciplinary Study. Cambridge University Press.Google Scholar
Kirschvink, J., Raub, T., and Fischer, W. 2012. Archean “whiffs of oxygen” go poof! p. 1943. 22nd Goldschmidt Conference Abstracts, Montréal, Canada.Google Scholar
Klaas, C., and Archer, D. E. 2002. Association of sinking organic matter with various types of mineral ballast in the deep sea: Implications for the rain ratio. Global Biogeochemical Cycles, 16(4):63-163–14.Google Scholar
Knauth, L. P., and Kennedy, M. J. 2009. The late Precambrian greening of the Earth. Nature, 460:728732.Google Scholar
Knoll, A. H. 2014. Paleobiological perspectives on early eukaryotic evolution. Cold Spring Harbor Perspectives in Biology, 6:aO16121. doi: 10.1101/cshperspect.a01612.Google Scholar
Knoll, A. H., Javaux, E. J., Hewitt, D., and Cohen, P. 2006. Eukaryotic organisms in Proterozoic oceans. Philosophical Transactions of the Royal Society B-Biological Sciences, 361:10231038.Google Scholar
Kodner, R. B., Summons, R. E., Pearson, A., King, N., and Knoll, A. H. 2008. Sterols in a unicellular relative of the metazoans. Proceedings of the National Academy of Sciences, 105:98979902.Google Scholar
Konhauser, K. O., Lalonde, S. V., Amskold, L., and Holland, H. D. 2007. Was there really an Archean phosphate crisis? Science, 315:12341234.Google Scholar
Kuipers, G., Beunk, F. F., and Van Der Wateren, F. M. 2013. Periglacial evidence for a 1.91–1.89 Ga old glacial period at low latitude, Central Sweden. Geology Today, 29:218221.Google Scholar
Kunzmann, M., Halverson, G. P., Sossi, P. A., Raub, T. D., Payne, J. L., and Kirby, J. 2013. Zn isotope evidence for immediate resumption of primary productivity after snowball Earth. Geology, 41:2730.Google Scholar
Kuypers, M. M. M., Sliekers, A. O., Lavik, G., Schmid, M., Jorgensen, B. B., Kuenen, J. G., Damste, J. S. S., Strous, M., and Jetten, M. S. M. 2003. Anaerobic ammonium oxidation by anammox bacteria in the Black Sea. Nature, 422:608611.Google Scholar
Lamb, D. M., Awramik, S. M., Chapman, D. J., and Zhu, S. 2009. Evidence for eukaryotic diversification in the ∼1800 million-year-old Changzhougou Formation, North China. Precambrian Research, 173(1–4): 93104.Google Scholar
Lane, N., and Martin, W. 2010. The energetics of genome complexity. Nature, 467:929934.Google Scholar
Large, R. R., Halpin, J. A., Danyushevsky, L. V., Maslennikov, V. V., Bull, S. W., Long, J. A., Gregory, D. D., Lounejeva, E., Lyons, T. W., Sack, P. J., McGoldrick, P. J., and Calver, C. R. 2014. Trace element content of sedimentary pyrite as a new proxy for deep-time ocean–atmosphere evolution. Earth and Planetary Science Letters, 389:209220.Google Scholar
Leavitt, W. D., Halevy, I., Bradley, A. S., and Johnston, D. T. 2013. Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record. Proceedings of the National Academy of Sciences, 110:1124411249.Google Scholar
Lee, C.-T. A., Shen, B., Slotnick, B. S., Liao, K., Dickens, G. R., Yokoyama, Y., Lenardic, A., Dasgupta, R., Jellinek, M., Lackey, J. S., Schneider, T., and Tice, M. M. 2013. Continental arc–island arc fluctuations, growth of crustal carbonates, and long-term climate change. Geosphere, 9:2136.Google Scholar
Lee, C.-T. A., Thurner, S., Paterson, S., and Cao, W. 2015. The rise and fall of continental arcs: Interplays between magmatism, uplift, weathering, and climate. Earth and Planetary Science Letters, 425:105119.Google Scholar
Li, C., Peng, P., Sheng, G. Y., Fu, J. M., and Yan, Y. Z. 2003. A molecular and isotopic geochemical study of Meso- to Neoproterozoic (1.73–0.85 Ga) sediments from the Jixian section, Yanshan Basin, North China. Precambrian Research, 125(3–4): 337356.Google Scholar
Li, Z.-X., Evans, D. A. D., and Halverson, G. P. 2013. Neoproterozoic glaciations in a revised global palaeogeography from the breakup of Rodinia to the assembly of Gondwanaland. Sedimentary Geology, 294:219232.Google Scholar
Li, Z.X., and Zhong, S. 2009. Supercontinent–superplume coupling, true polar wander and plume mobility: Plate dominance in whole-mantle tectonics. Physics of the Earth and Planetary Interiors, 176, 143156.Google Scholar
Liu, C., Wang, Z., Raub, T. D., MacDonald, F. A., and Evans, D. A. 2014. Neoproterozoic cap-dolostone deposition in stratified glacial meltwater plume. Earth and Planetary Science Letters, 404:2232.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
Love, G. D., Grosjean, E., Stalvies, C., Fike, D. A., Grotzinger, J. P., Bradley, A. S., Kelly, A. E., Bhatia, M., Meredith, W., Snape, C. E., Bowring, S. A., Condon, D. J., and Summons, R. E. 2009. Fossil steroids record the appearance of Demospongiae during the Cryogenian period. Nature, 457:718721.Google Scholar
Love, G. D., Snape, C. E., Carr, A. D., and Houghton, R. C. 1995. Release of covalently-bound alkane biomarkers in high yields from kerogen via catalytic hydropyrolysis. Organic Geochemistry, 23(10):981986.Google Scholar
Love, G. D., and Summons, R. E. In press. The molecular record of Cryogenian sponges—a response to Antcliffe (2013). Palaeontology.Google Scholar
Lyons, T. W., Reinhard, C. T., Love, G. D., and Xiao, S. 2012. Geobiology of the Proterozoic Eon. In Knoll, A. H., Canfield, D. E., and Konhauser, K. O. (eds.), Fundamentals of Geobiology. John Wiley & Sons, Chichester. doi: 10.1002/9781118280874.ch20.Google Scholar
Lyons, T. W., Reinhard, C. T., and Planavsky, N. J. 2014. The rise of oxygen in Earth's early ocean and atmosphere. Nature, 506:307315.Google Scholar
Lyons, T. W., and Severmann, S. 2006. A critical look at iron paleoredox proxies: New insights from modern euxinic marine basins. Geochimica et Cosmochimica Acta, 70:56985722.Google Scholar
Marin-Carbonne, J., Robert, F., and Chaussidon, M. 2014. The silicon and oxygen isotope-compositions of Precambrian cherts: A record of oceanic paleo-temperatures? Precambrian Research, 247:223234.Google Scholar
Marriott, C. S., Henderson, G. M., Crompton, R., Staubwasser, M., and Shaw, S. 2004. Effect of mineralogy, salinity, and temperature on Li/Ca and Li isotope composition of calcium carbonate. Chemical Geology, 212(1–2): 515.Google Scholar
Martin, W., and Muller, M. 1998. The hydrogen hypothesis for the first eukaryote. Nature, 392:3741.Google Scholar
McArthur, J., Howarth, R., and Shields, G. 2012. Strontium isotope stratigraphy, p. 127144 In Gradstein, F. M., Ogg, J. G. Schmitz, M. D., and Ogg, G. M. (eds.), The Geologic Time Scale. Elsevier.Google Scholar
McCaffrey, M. A., Moldowan, J. M., Lipton, P. A., Summons, R. E., Peters, K. E., Jeganathan, A., and Watt, D. S. 1994. Paleoenvironmental implications of novel C-30 steranes in Precambrian to Cenozoic age petroleum and bitumen. Geochimica et Cosmochimica Acta, 58:529532.Google Scholar
McKenzie, N. R., Hughes, N. C., Gill, B. C., and Myrow, P. M. 2014. Plate tectonic influences on Neoproterozoic–early Paleozoic climate and animal evolution. Geology, 42:127130.Google Scholar
McKirdy, D. M., Webster, L. J., Arouri, K. R., Grey, K., and Gostin, V. A. 2006. Contrasting sterane signatures in Neoproterozoic marine rocks of Australia before and after the Acraman asteroid impact. Organic Geochemistry, 37:189207.Google Scholar
Meert, J. G. 2012. What's in a name? The Columbia (Paleopangaea/Nuna) supercontinent. Gondwana Research, 21:987993.Google Scholar
Mills, D. B., Ward, L. M., Jones, C., Sweeten, B., Forth, M., Treusch, A. H., and Canfield, D. E. 2014. Oxygen requirements of the earliest animals. Proceedings of the National Academy of Sciences, 111:41684172.Google Scholar
Mitchell, R. L., and Sheldon, N. D. 2010. The ∼1100 Ma Sturgeon Falls paleosol revisited: Implications for Mesoproterozoic weathering environments and atmospheric CO2 levels. Precambrian Research, 183:738748.Google Scholar
Moldowan, J. M., Fago, F. J., Lee, C. Y., Jacobson, S. R., Watt, D. S., Slougui, N. E., Jeganathan, A., and Young, D. C. 1990. Sedimentary 24-n-propylcholestanes, molecular fossils diagnostic of marine algae. Science, 247:309312.Google Scholar
Moreira, D., and Lopez-Garcia, P. 1998. Symbiosis between methanogenic archaea and δ-proteobacteria as the origin of eukaryotes: The syntrophic hypothesis. Journal of Molecular Evolution, 47:517530.Google Scholar
Moroz, L. L., Kocot, K. M., Citarella, M. R., Dosung, S., Norekian, T. P., Povolotskaya, I. S., Grigorenko, A. P., Dailey, C., Berezikov, E., Buckley, K. M., Ptitsyn, A., Reshetov, D., Mukherjee, K., Moroz, T. P., Bobkova, Y., Yu, F. H., Kapitonov, V. V., Jurka, J., Bobkov, Y. V., Swore, J. J., Girardo, D. O., Fodor, A., Gusev, F., Sanford, R., Bruders, R., Kittler, E., Mills, C. E., Rast, J. P., Derelle, R., Solovyev, V. V., Kondrashov, F. A., Swalla, B. J., Sweedler, J. V., Rogaev, E. I., Halanych, K. M., and Kohn, A. B. 2014. The ctenophore genome and the evolutionary origins of neural systems. Nature, 510:109114.Google Scholar
Morse, J. W., Zullig, J. J., Bernstein, L. D., Millero, F. J., Milne, P. J., Mucci, A., and Choppin, G. R. 1985. Chemistry of calcium carbonate-rich shallow water sediments in the Bahamas. American Journal of Science, 285:147185.Google Scholar
Müller, R. D., Sdrolias, M., Gaina, C., and Roest, W. R. 2008. Age, spreading rates, and spreading asymmetry of the world's ocean crust. Geochemistry, Geophysics, Geosystems, 9(4)119.Google Scholar
Nance, R. D., Murphy, J. B., and Santosh, M. 2014. The supercontinent cycle: A retrospective essay. Gondwana Research, 25:429.Google Scholar
Neuweiler, F., Turner, E. C., and Burdige, D. J. 2009. Early Neoproterozoic origin of the metazoan clade recorded in carbonate rock texture. Geology, 37:475478.Google Scholar
Noordmann, J., Weyer, S., Montoya-Pino, C., Dellwig, O., Neubert, N., Eckert, S., Paetzel, M., and Bottcher, M. E. 2015. Uranium and molybdenum isotope systematics in modern euxinic basins: Case studies from the central Baltic Sea and the Kyllaren Fjord (Norway). Chemical Geology, 396:182195.Google Scholar
Nursall, J. 1959. Oxygen as a prerequisite to the origin of the Metazoa. Nature, 183:11701172.Google Scholar
Ohnemueller, F., Prave, A. R., Fallick, A. E., and Kasemann, S. A. 2014. Ocean acidification in the aftermath of the Marinoan glaciation. Geology, 42:11031106.Google Scholar
Olson, S. L., Kump, L. R., and Kasting, J. F. 2013. Quantifying the areal extent and dissolved oxygen concentrations of Archean oxygen oases. Chemical Geology, 362:3543.Google Scholar
Palike, H., Lyle, M. W., Nishi, H., Raffi, I., Ridgwell, A., Gamage, K., Klaus, A., Acton, G., Anderson, L., Backman, J., Baldauf, J., Beltran, C., Bohaty, S. M., Bownpaul, , Busch, W., Channell, J. E. T., Chun, C. O. J., Delaney, M., Dewangan, P., Dunkley Jones, T., Edgar, K. M., Evans, H., Fitch, P., Foster, G. L., Gussone, N., Hasegawa, H., Hathorne, E. C., Hayashi, H., Herrle, J. O., Holbourn, A., Hovan, S., Hyeong, K., Iijima, K., Ito, T., Kamikuri, S.-I., Kimoto, K., Kuroda, J., Leon-Rodriguez, L., Malinverno, A., Moore, T. C. Jr., Murphy, B. H., Murphy, D. P., Nakamura, H., Ogane, K., Ohneiser, C., Richter, C., Robinson, R., Rohling, E. J., Romero, O., Sawada, K., Scher, H., Schneider, L., Sluijs, A., Takata, H., Tian, J., Tsujimoto, A., Wade, B. S., Westerhold, T., Wilkens, R., Williams, T., Wilson, P. A., Yamamoto, Y., Yamamoto, S., Yamazaki, T., and Zeebe, R. E. 2012. A Cenozoic record of the equatorial Pacific carbonate compensation depth. Nature, 488:609614.Google Scholar
Palmer, M. R., and Edmond, J. M. 1989. The strontium isotope budget of the modern ocean. Earth and Planetary Science Letters, 92:1126.Google Scholar
Pan, Y. M., and Stauffer, M. R. 2000. Cerium anomaly and Th/U fractionation in the 1.85 Ga Flin Flon Paleosol: Clues from REE- and U-rich accessory minerals and implications for paleoatmospheric reconstruction. American Mineralogy, 85:898911.Google Scholar
Pang, K., Tang, Q., Schiffbauer, J. D., Yao, J., Yuan, X., Wan, B., Chen, L., Ou, Z., and Xiao, S. 2013. The nature and origin of nucleus-like intracellular inclusions in Paleoproterozoic eukaryote microfossils. Geobiology, 11:499510.Google Scholar
Pang, K., Tang, Q., Yuan, X.-L., Wan, B., and Xiao, S. 2015. A biomechanical analysis of the early eukaryotic fossil Valeria and new occurrence of organic-walled microfossils from the Paleo-Mesoproterozoic Ruyang Group. Palaeoworld, doi:10.1016/j.palwor.2015.04.002.Google Scholar
Partin, C. A., Bekker, A., Planavsky, N. J., Scott, C. T., Gill, B. C., Li, C., Podkovyrov, V., Maslov, A., Konhauser, K. O., Lalonde, S. V., Love, G. D., Poulton, S. W., and Lyons, T. W. 2013a. Large-scale fluctuations in Precambrian atmospheric and oceanic oxygen levels from the record of U in shales. Earth and Planetary Science Letters, 369:284293.Google Scholar
Partin, C. A., Lalonde, S. V., Planavsky, N. J., Bekker, A., Rouxel, O. J., Lyons, T. W., and Konhauser, K. O. 2013b. Uranium in iron formations and the rise of atmospheric oxygen. Chemical Geology, 362:8290.Google Scholar
Pavlov, A. A., Hurtgen, M. T., Kasting, J. F., and Arthur, M. A. 2003. Methane-rich Proterozoic atmosphere? Geology, 31:8790.Google Scholar
Pavlov, A. A., Kasting, J. F., Brown, L. L., Rages, K. A., and Freedman, R. 2000. Greenhouse warming by CH4 in the atmosphere of early Earth. Journal of Geophysical Research: Planets (1991–2012), 105(E5):1198111990.Google Scholar
Pawlowska, M. M., Butterfield, N. J., and Brocks, J. J. 2013. Lipid taphonomy in the Proterozoic and the effect of microbial mats on biomarker preservation. Geology, 41:103106.Google Scholar
Pecoits, E., Konhauser, K. O., Aubet, N. R., Heaman, L. M., Veroslavsky, G., Stern, R. A., and Gingras, M. K. 2012. Bilaterian burrows and grazing behavior at > 585 million years ago. Science, 336:16931696.Google Scholar
Pehrsson, S. J., Eglington, B. M., Evans, D. A. D., Huston, D., and Reddy, S. M. 2015. Metallogeny and its link to orogenic style during the Nuna supercontinent cycle. The Geological Society of London Special Publications, 424. doi:10.1144/SP424.5.Google Scholar
Peng, Y. B., Bao, H. M., and Yuan, X. L. 2009. New morphological observations for Paleoproterozoic acritarchs from the Chuanlinggou Formation, North China. Precambrian Research, 168(3–4): 223232.Google Scholar
Petit, J. R., Jouzel, J., Raynaud, D., Barkov, N. I., Barnola, J. M., Basile, I., Bender, M., Chappellaz, J., Davis, M., Delaygue, G., Delmotte, M., Kotlyakov, V. M., Legrand, M., Lipenkov, V. Y., Lorius, C., Pepin, L., Ritz, C., Saltzman, E., and Stievenard, M. 1999. Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature, 399:429436.Google Scholar
Pisarevsky, S. A., Elming, S.-Å., Pesonen, L. J., and Li, Z.-X. 2014. Mesoproterozoic paleogeography: Supercontinent and beyond. Precambrian Research, 244:207225.Google Scholar
Planavsky, N. 2009. Early Neoproterozoic origin of the metazoan clade recorded in carbonate rock texture: COMMENT. Geology, 37(9):E195E195.Google Scholar
Planavsky, N. J., Bekker, A., Hofmann, A., Owens, J. D., and Lyons, T. W. 2012. Sulfur record of rising and falling marine oxygen and sulfate levels during the Lomagundi event. Proceedings of the National Academy of Sciences, 109:1830018305.Google Scholar
Planavsky, N. J., McGoldrick, P., Scott, C. T., Li, C., Reinhard, C. T., Kelly, A. E., Chu, X. L., Bekker, A., Love, G. D., and Lyons, T. W. 2011. Widespread iron-rich conditions in the mid-Proterozoic ocean. Nature, 477:448–U95.Google Scholar
Planavsky, N. J., Reinhard, C. T., Wang, X., Thomson, D., McGoldrick, P., Rainbird, R. H., Johnson, T., Fischer, W. W., and Lyons, T. W. 2014. Low Mid-Proterozoic atmospheric oxygen levels and the delayed rise of animals. Science, 346:635638.Google Scholar
Planavsky, N., Rouxel, O., Bekker, A., Lalonde, S., Konhauser, K. O., Reinhard, C. T., and Lyons, T. W. 2010. The evolution of the marine phosphate reservoir. Nature, 467:10881090.Google Scholar
Ploug, H., Iversen, M. H., and Fischer, G. 2008. Ballast, sinking velocity, and apparent diffusivity within marine snow and zooplankton fecal pellets: Implications for substrate turnover by attached bacteria. Limnology and Oceanography, 53:18781886.Google Scholar
Pogge Von Strandmann, P. A. E., and Henderson, G. M. 2014. The Li isotope response to mountain uplift. Geology, doi: 10.1130/G36162.1.Google Scholar
Porter, S. 2011. The rise of predators. Geology, 39:607608.Google Scholar
Porter, S. M., Meisterfeld, R., and Knoll, A. H. 2003. Vase-shaped microfossils from the Neoproterozoic Chuar Group, Grand Canyon: A classification guided by modern testate amoebae. Journal of Paleontology, 77:409429.Google Scholar
Poulton, S. W., and Canfield, D. E. 2005. Development of a sequential extraction procedure for iron: implications for iron partitioning in continentally derived particulates. Chemical Geology, 214(3–4): 209221.Google Scholar
Poulton, S. W., and Canfield, D. E. 2011. Ferruginous conditions: A dominant feature of the ocean through Earth's history. Elements, 7:107112.Google Scholar
Poulton, S. W., Fralick, P. W., and Canfield, D. E. 2004. The transition to a sulphidic ocean similar to 1.84 billion years ago. Nature, 431:173177.Google Scholar
Pratt, L. M., Summons, R. E., and Hieshima, G. B. 1991. Sterane and triterpane biomarkers in the Precambrian Nonesuch Formation, North American Midcontinent Rift. Geochimica et Cosmochimica Acta, 55(3):911916.Google Scholar
Pufahl, P. K., and Hiatt, E. E. 2012. Oxygenation of the Earth's atmosphere–ocean system: A review of physical and chemical sedimentologic responses. Marine and Petroleum Geology, 32:120.Google Scholar
Rauch, J. N., and Pacyna, J. M. 2009. Earth's global Ag, Al, Cr, Cu, Fe, Ni, Pb, and Zn cycles. Global Biogeochemical Cycles, 23:GB2001, doi: 10.1029/2008GB003376.Google Scholar
Redfield, A. C. 1934. On the proportions of organic derivations in sea water and their relation to the composition of plankton, p. 177192 In Daniel, R. J. (ed.), James Johnstone Memorial Volume, University Press of Liverpool, Liverpool, UK.Google Scholar
Reinhard, C. T., Lalonde, S. V., and Lyons, T. W. 2013a. Oxidative sulfide dissolution on the early Earth. Chemical Geology, 362:4455.Google Scholar
Reinhard, C. T., Planavsky, N. J., Robbins, L. J., Partin, C. A., Gill, B. C., Lalonde, S. V., Bekker, A., Konhauser, K. O., and Lyons, T. W. 2013b. Proterozoic ocean redox and biogeochemical stasis: Proceedings of the National Academy of Sciences, 110:53575362.Google Scholar
Richard, F. C. and Bourg, A. C. M. 1991. Aqueous geochemistry of chromium—a review. Water Research 25, 807816.Google Scholar
Ridgwell, A., and Zeebe, R. E. 2005. The role of the global carbonate cycle in the regulation and evolution of the Earth system. Earth and Planetary Science Letters, 234(3–4): 299315.Google Scholar
Robert, F., and Chaussidon, M. 2006. A palaeotemperature curve for the Precambrian oceans based on silicon isotopes in cherts. Nature, 443:969972.Google Scholar
Rodrigues, J. B., Pimentel, M. M., Buhn, B., Matteini, M., Dardenne, M. A., Alvarenga, C. J. S., and Armstrong, R. A. 2012. Provenance of the Vazante Group: New U–Pb, Sm–Nd, Lu–Hf isotopic data and implications for the tectonic evolution of the Neoproterozoic Brasilia Belt. Gondwana Research, 21(2–3): 439450.Google Scholar
Rooney, A. D., Strauss, J. V., Brandon, A. D., and MacDonald, F. A. 2015. A Cryogenian chronology: Two long-lasting synchronous Neoproterozoic glaciations. Geology, 43:459462.Google Scholar
Rowley, D. B. 2002. Rate of plate creation and destruction: 180 Ma to present. Geological Society of America Bulletin, 114:927933.Google Scholar
Runnegar, B. 1991. Precambrian oxygen levels estimated from the biochemistry and physiology of early eukaryotes. Palaeogeography, Palaeoclimatology, Palaeoecology, 97(1–2): 97111.Google Scholar
Rye, R., and Holland, H. D. 1998. Paleosols and the evolution of atmospheric oxygen: a critical review. American Journal of Science, 298:621672.Google Scholar
Sahoo, S. K., Planavsky, N. J., Kendall, B., Wang, X. Q., Shi, X. Y., Scott, C., Anbar, A. D., Lyons, T. W., and Jiang, G. Q. 2012. Ocean oxygenation in the wake of the Marinoan glaciation. Nature, 489:546549.Google Scholar
Schauble, E. A., Rossman, G. R., and Taylor, H. P. J. 2004. Theoretical estimates of equilibrium chromium-isotope fractionations. Chemical Geology, 205:99114.Google Scholar
Schoenberg, R., Zink, S., Staubwasser, M., and Von Blanckenburg, F. 2008. The stable Cr isotope inventory of solid Earth reservoirs determined by double spike MC-ICP-MS. Chemical Geology, 249(3–4): 294306.Google Scholar
Schrag, D. P., Berner, R. A., Hoffman, P. F., and Halverson, G. P. 2002. On the initiation of a Snowball Earth. Geochemistry Geophysics Geosystems, 3:doi: 10.1029/2001gc000219.Google Scholar
Schrag, D. P., Higgins, J. A., MacDonald, F. A., and Johnston, D. T. 2013. Authigenic carbonate and the history of the global carbon cycle. Science, 339:540543.Google Scholar
Schulz, H. N., Brinkhoff, T., Ferdelman, T. G., Marine, M. H., Teske, A., and Jorgensen, B. B. 1999. Dense populations of a giant sulfur bacterium in Namibian shelf sediments. Science, 284:493495.Google Scholar
Schwark, L., and Empt, P. 2006. Sterane biomarkers as indicators of Palaeozoic algal evolution and extinction events. Palaeogeography, Palaeoclimatology, Palaeoecology, 240(1–2): 225236.Google Scholar
Scott, C., Lyons, T. W., Bekker, A., Shen, Y., Poulton, S., Chu, X., and Anbar, A. 2008. Tracing the stepwise oxygenation of the Proterozoic biosphere. Nature, 452:456459.Google Scholar
Scott, C., Wing, B. A., Bekker, A., Planavsky, N. J., Medvedev, P., Bates, S. M., Yun, M., and Lyons, T. W. 2014. Pyrite multiple-sulfur isotope evidence for rapid expansion and contraction of the early Paleoproterozoic seawater sulfate reservoir. Earth and Planetary Science Letters, 389:95104.Google Scholar
Sheldon, N. D. 2013. Causes and consequences of low atmospheric pCO(2) in the late Mesoproterozoic. Chemical Geology, 362:224231.Google Scholar
Singh, H. B., and Kasting, J. F. 1988. Chlorine-hydrocarbon photochemistry in the marine troposphere and lower stratosphere. Origins of Life and Evolution of the Biosphere, 7:261285.Google Scholar
Sim, M. S., Bosak, T., and Ono, S. 2011. Large sulfur isotope fractionation does not require disproportionation. Science, 333:7477.Google Scholar
Spang, A., Saw, J. H., Jorgensen, S. L., Zaremba-Niedzwiedzka, K., Martijn, J., Lind, A. E., Van Eijk, R., Schleper, C., Guy, L., and Ettema, T. J. G. 2015. Complex Archaea that bridge the gap between prokaryotes and eukaryotes. Nature, 521:173179.Google Scholar
Sperling, E. A., Halverson, G. P., Knoll, A. H., MacDonald, F. A., and Johnston, D. T. 2013. A basin redox transect at the dawn of animal life. Earth and Planetary Science Letters, 371–372:143155.Google Scholar
Stanley, G. D. Jr., and Stürmer, W. 1983. The first fossil ctenophore from the lower Devonian of West Germany. Nature, 303:518520.Google Scholar
Strother, P. K., Battison, L., Brasier, M. D., and Wellman, C. H. 2011. Earth's earliest non-marine eukaryotes. Nature, 473:505509.Google Scholar
Su, W., Li, H., Xu, L., Jia, S., Geng, J., Zhou, H., Wang, Z., and Pu, H. 2012. Luoyu and Ruyang Group at the south margin of the North China Craton (NCC) should belong in the Mesoproterozoic Changchengian System: Direct constraints from the LA-MC-ICPMS U-Pb age of the tuffite in the Luoyukou Formation, Ruzhou, Henan, China. Geological Survey and Research, 35:96108.Google Scholar
Summons, R. E., Bradley, A. S., Jahnke, L. L., and Waldbauer, J. R. 2006. Steroids, triterpenoids and molecular oxygen. Philosophical Transactions of the Royal Society B-Biological Sciences, 361:951968.Google Scholar
Summons, R. E., Brassell, S. C., Eglinton, G., Evans, E., Horodyski, R. J., Robinson, N., and Ward, D. M. 1988. Distinctive hydrocarbon biomarkers from fossiliferous sediment of the late Proterozoic Walcott Member, Chuar Group, Grand Canyon, Arizona. Geochimica et Cosmochimica Acta, 52:26252637.Google Scholar
Summons, R. E., Thomas, J., Maxwell, J. R., and Boreham, C. J. 1992. Secular and environmental constraints on the occurrence of dinosterane in sediments. Geochimica et Cosmochimica Acta, 56:24372444.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
Tarhan, L. G., Droser, M. L., and Gehling, J. G. 2015. Depositional and preservational environments of the Ediacara Member, Rawnsley Quartzite (South Australia): Assessment of paleoenvironmental proxies and the timing of ‘ferruginization.’ Palaeogeography, Palaeoclimatology, Palaeoecology, 434:413.Google Scholar
Thomson, D., Rainbird, R. H., and Dix, G. 2014. Architecture of a Neoproterozoic intracratonic carbonate ramp succession: Wynniatt Formation, Amundsen Basin, Arctic Canada. Sedimentary Geology, 299:119138.Google Scholar
Towe, K. M. 1970. Oxygen-collagen priority and the early metazoan fossil record. Proceedings of the National Academy of Sciences, 65:781788.Google Scholar
Turner, J. T. 2002. Zooplankton fecal pellets, marine snow and sinking phytoplankton blooms. Aquatic Microbial Ecology, 27:57102.Google Scholar
Tyrrell, T. 1999. The relative influences of nitrogen and phosphorus on oceanic primary production. Nature, 400:525531.Google Scholar
Ventura, G. T., Kenig, F., Grosjean, E., and Summons, R. E. 2004. Biomarker analysis of solvent extractable organic matter from the late Neoproterozoic Kwagunt Formation, Chuar Group (∼800–742 Ma), Grand Canyon. Geological Society of America Abstracts with Programs, Vol. 36(5):170.Google Scholar
Volkman, J. K. 2003. Sterols in microorganisms. Applied Microbiology and Biotechnology, 60:495506.Google Scholar
Volkman, J. K. 2005. Sterols and other triterpenoids: source specificity and evolution of biosynthetic pathways. Organic Geochemistry, 36:139159.Google Scholar
Walter, M. R., Du, R. L., and Horodyski, R. J. 1990. Coiled carbonaceous megafossils from the middle Proterozoic of Jixian (Tianjin) and Montana. American Journal of Science, 290A:133148.Google Scholar
Watson, B. 2008. Quiet revolution in the geochemical sciences. Elements, 4:219220.Google Scholar
Whelan, N. V., Kocot, K. M., Moroz, L. L., and Halanych, K. M. 2015. Error, signal, and the placement of Ctenophora sister to all other animals. Proceedings of the National Academy of Sciences, 112:57735778.Google Scholar
Williams, G. E. 2005. Subglacial meltwater channels and glaciofluvial deposits in the Kimberley Basin, Western Australia: 1.8 Ga low-latitude glaciation coeval with continental assembly. Journal of the Geological Society, 162:111124.Google Scholar
Williams, T. A., Foster, P. G., Cox, C. J., and Embley, T. M. 2013. An archaeal origin of eukaryotes supports only two primary domains of life. Nature, 504:231236.Google Scholar
Williams, T. A., Foster, P. G., Nye, T. M. W., Cox, C. J., and Embley, T. M. 2012. A congruent phylogenomic signal places eukaryotes within the Archaea. Proceedings of the Royal Society B-Biological Sciences, 279:48704879.Google Scholar
Woese, C. R., Kandler, O., and Wheelis, M. L. 1990. Towards a natural system of organisms—proposal for the domains Archaea, Bacteria, and Eucarya. Proceedings of the National Academy of Sciences, 87:45764579.Google Scholar
Wolf, E. T., and Toon, O. B. 2014. Controls on the Archean climate system investigated with a global climate model. Astrobiology, 14:241253.Google Scholar
Wordsworth, R., and Pierrehumbert, R. 2013. Hydrogen-nitrogen greenhouse warming in Earth's early atmosphere. Science, 339:6467.Google Scholar
Yoon, H. S., Hackett, J. D., Ciniglia, C., Pinto, G., and Bhattacharya, D. 2004. A molecular timeline for the origin of photosynthetic eukaryotes. Molecular Biology and Evolution, 21:809818.Google Scholar
Zbinden, E. A., Holland, H. D., Feakes, C. R. and Dobos, S. K. 1988. The Sturgeon Falls Paleosol and the composition of the atmosphere 1.1 Ga Bp. Precambrian Research, 42:141163.Google Scholar
Zink, S., Schoenberg, R., and Staubwasser, M. 2010. Isotopic fractionation and reaction kinetics between Cr(III) and Cr(VI) in aqueous media. Geochimica et Cosmochimica Acta, 74:57295745.Google Scholar