Hostname: page-component-7c8c6479df-995ml Total loading time: 0 Render date: 2024-03-28T10:51:31.255Z Has data issue: false hasContentIssue false

Sulphur tales from the early Archean world

Published online by Cambridge University Press:  04 April 2016

A. Montinaro*
Affiliation:
Institut für Geologie und Paläontologie, Westfälische Wilhelms-Universität, Corrensstr. 24, 48149 Münster, Germany
H. Strauss
Affiliation:
Institut für Geologie und Paläontologie, Westfälische Wilhelms-Universität, Corrensstr. 24, 48149 Münster, Germany

Abstract

Sedimentary and magmatic rocks and their distinct sulphur isotopic signatures indicate the sources and processes of sulphur cycling, in particular through the analysis of all four stable sulphur isotopes (32S, 33S, 34S and 36S). Research over the past 15 years has substantially advanced our understanding of sulphur cycling on the early Earth, most notably through the discovery of mass-independently fractionated sulphur isotopic signatures. A strong atmospheric influence on the early Archean global sulphur cycle is apparent, much in contrast to the modern world. Diverse microbially driven sulphur cycling is clearly discernible, but its importance for Earth surface environments remains to be quantified.

Type
Research Article
Copyright
Copyright © Cambridge University Press 2016 

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

Alt, J.C. (1995). Sulfur isotopic profile through the oceanic crust: Sulfur mobility and seawater-crustal sulfur exchange during hydrothermal alteration. Geology 23, 585588.Google Scholar
Alt, J.C. & Shanks, W.C. III (2003). Serpentinization of abyssal peridotites from the MARK area, Mid-Atlantic Ridge: sulfur geochemistry and reaction modeling. Geochim. Cosmochim. Acta 67, 641653.Google Scholar
Alt, J.C., Anderson, T.F. & Bonnell, L. (1989). The geochemistry of sulfur in a 1.3 km section of hydrothermally altered oceanic crust, DSDP Hole 504B. Geochim. Cosmochim. Acta 53 10111023.CrossRefGoogle Scholar
Bao, H., Rumble, D. & Lowe, D.R. (2007). The five stable isotope compositions of Fig Tree barites: Implications on sulfur cycle in ca. 3.2 Ga oceans. Geochim. Cosmochim. Acta 71, 48684879.Google Scholar
Beraldi-Campesi, H. (2013). Early life on land and the first terrestrial ecosystems. Ecol. Process. 2, 117.Google Scholar
Böttcher, M.E., Schale, H., Schnetger, B., Wallmann, K. & Brumsack, H.J. (2000). Stable sulfur isotopes indicate net sulfate reduction in near surface sediments of the deep Arabian Sea. Deep Sea Res. II 47, 27692783.Google Scholar
Böttcher, M.E., Thamdrup, B. & Vennemann, T.W. (2001). Oxygen and sulfur isotope fractionation during anaerobic bacterial disproportionation of elemental sulfur. Geochim. Cosmochim. Acta 65, 16011609.Google Scholar
Böttcher, M.E., Thamdrup, B., Gehre, M. & Theune, A. (2005). 34S/34S and 18O/16O fractionation during sulfur disproportionation by Desulfobulbus propionicus. Geomicrobiol. J. 22, 219226.Google Scholar
Böttcher, M.E., Brumsack, H.J. & Dürselen, C.D. (2007). The isotopic composition of modern seawater sulfate: I. Coastal waters with special regard to the North Sea. J. Mar. Syst. 67, 7382.CrossRefGoogle Scholar
Brüchert, V., Knoblauch, C. & Jørgensen, B.B. (2001). Controls on stable sulfur isotope fractionation during bacterial sulfate reduction in Arctic sediments. Geochim. Cosmochim. Acta 65(5), 763776.Google Scholar
Canfield, D.E. (1998). A new model for Proterozoic ocean chemistry. Nature 396, 450453.Google Scholar
Canfield, D.E. (2001). Biogeochemistry of sulfur isotopes. Rev. Mineral. Geochem. 43, 607636.Google Scholar
Canfield, D.E. & Raiswell, R. (1999). The evolution of the sulfur cycle. Am. J. Sci. 299, 697723.Google Scholar
Canfield, D.E., Olesen, C.A. & Cox, R.P. (2006). Temperature and its control of isotope fractionation by a sulfate-reducing bacterium. Geochim. Cosmochim. Acta 70(3), 548561.Google Scholar
Canfield, D.E., Farquhar, J. & Zerkle, A.L. (2010). High isotope fractionations during sulfate reduction in a low-sulfate euxinic ocean analog. Geology 38(5), 415418.Google Scholar
Cates, N.L. & Mojzsis, S.J. (2006). Chemical and isotopic evidence for widespread Eoarchean metasedimentary enclaves in southern West Greenland. Geochim. Cosmochim. Acta 70, 42294257.Google Scholar
Claypool, G.E., Holser, W.T., Kaplan, I.R., Sakai, H. & Zak, I. (1980). The age curves for sulfur and oxygen isotopes in marine sulfate and their mutual interpretation. Chem. Geol. 28, 199260.Google Scholar
Crowe, S.A., Paris, G., Katsev, S., Jones, C., Kim, S.T., Zerkle, A.L. & Canfield, D.E. (2014). Sulfate was a trace constituent of Archean seawater. Science 346, 735739.Google Scholar
Detmers, J., Brüchert, V., Habicht, K.S. & Kuever, J. (2001). Diversity of sulfur isotope fractionations by sulfate-reducing prokaryotes. Appl. Environ. Microbiol. 67(2), 888894.CrossRefGoogle ScholarPubMed
Eldridge, C.S., Compston, W., Williams, I.S., Harris, J.W. & Bristow, J.W. (1991). Isotope evidence for the involvement of recycled sediments in diamond formation. Nature 353, 649653.Google Scholar
Farquhar, J., Bao, H. & Thiemens, M. (2000). Atmospheric influence of Earth's earliest sulfur cycle. Science 289, 756758.Google Scholar
Farquhar, J., Wing, B.A., McKeegan, K.D., Harris, J.W., Cartigny, P. & Thiemens, M.H. (2002). Mass-independent sulfur of inclusions in diamond and sulfur recycling on early Earth. Science 298, 23692372.Google Scholar
Farquhar, J., Johnston, D.T., Wing, B.A., Habicht, K.S., Canfield, D.E., Airieau, S. & Thiemens, M.H. (2003). Multiple sulphur isotopic interpretations of biosynthetic pathways: implications for biological signatures in the sulphur isotope record. Geobiology 1, 2736.Google Scholar
Farquhar, J., Peters, M., Johnston, D.T., Strauss, H., Masterson, A., Wiechert, U. & Kaufman, A.J. (2007). Isotopic evidence for Mesoarchaean anoxia and changing atmospheric sulphur chemistry. Nature 449, 706709.Google Scholar
Farquhar, J., Canfield, D.E., Masterson, A., Bao, H. & Johnston, D. (2008). Sulfur and oxygen isotope study of sulfate reduction in experiments with natural populations from Faellestrand, Denmark. Geochim. Cosmochim. Acta 72(12), 28052821.Google Scholar
Fike, D.A., Gammon, C.L., Ziebis, W. & Orphan, V.J. (2008). Micron-scale mapping of sulfur cycling across the oxycline of a cyanobacterial mat: a paired nanoSIMS and CARD-FISH approach. Int. Soc. Microb. Ecol. J. 2(7), 749759.Google Scholar
Fischer, W.W., Fike, D.A., Johnson, J.E., Raub, T.D., Guan, Y., Kirschvink, J.L. & Eiler, J.M. (2014). SQUID-SIMS is a useful approach to uncover primary signals in the Archean sulfur cycle. Proc. Natl. Acad. Sci. U. S. A. 111, 54685473.Google Scholar
Habicht, K.S. & Canfield, D.E. (1997). Sulfur isotope fractionation during bacterial sulfate reduction in organic-rich sediments. Geochim. Cosmochim. Acta 61(24), 53515361.Google Scholar
Habicht, K.S., Canfield, D.E. & Rethmeier, J. (1998). Sulfur isotope fractionation during bacterial reduction and disproportionation of thiosulfate and sulfite. Geochim. Cosmochim. Acta 62, 25852595.Google Scholar
Habicht, K.S., Gade, M., Thamdrup, B., Berg, P. & Canfield, D.E. (2002). Calibration of sulfate levels in the Archean ocean. Science 298, 23722374.Google Scholar
Harrison, A.G. & Thode, H.G. (1958). Mechanism of the bacterial reduction of sulphate from isotope fractionation studies. Trans. Faraday Soc. 54, 8492.Google Scholar
Holland, H.D. (2004). The geologic history of seawater. In The Oceans and Marine Chemistry, ed. Elderfield, H., Elsevier, Amsterdam, pp. 583625.Google Scholar
Holland, H.D., Turekian, K.K. & Schlesinger, W.H. (2004). Treatise on Geochemistry – Biogeochemistry, Vol. 8, Elvesier Pergamon, Amsterdam.Google Scholar
Hulston, J.R. & Thode, H.G. (1965). Variations in the S33, S34, and S36 contents of meteorites and their relation to chemical and nuclear effects. J. Geophys. Res. 70, 34753484.Google Scholar
Hu, G., Rumble, D. & Wang, P.L. (2003). An ultraviolet laser microprobe for the in situ analysis of multisulfur isotopes and its use in measuring Archean sulfur isotope mass-independent anomalies. Geochim. Cosmochim. Acta 67, 31013118.Google Scholar
Johnston, D. (2011). Multiple sulfur isotopes and the evolution of Earth's surface sulfur cycle. Earth-Sci. Rev. 106, 161183.Google Scholar
Johnston, D.T., Farquhar, J., Wing, B.A., Kaufman, A.J., Canfield, D.E. & Habicht, K.S. (2005). Multiple sulfur isotope fractionations in biological system: a case study with sulfate reducers and sulfur disproportionation. Am. J. Sci. 305, 645660.Google Scholar
Johnston, D.T., Farquhar, J. & Canfield, D.E. (2007). Sulfur isotopes insight into microbial sulfate reduction: when microbes meet models. Geochim. Cosmochim. Acta 71, 39293947.Google Scholar
Johnston, D.T., Farquhar, J., Habicht, K.S. & Canfield, D.E. (2008). Sulphur isotopes and the search for life: strategies for identifying sulphur metabolisms in the rock record and beyond. Geobiology 6, 425435.Google Scholar
Jørgensen, B.B., Isaksen, M.F. & Jannasch, H.W. (1992). Bacterial sulfate reduction above 100 C in deep-sea hydrothermal vent sediments. Science 258, 17561757.Google Scholar
Kemp, A.L.W. & Thode, H.G. (1968). The mechanism of the bacterial reduction of sulphate and of sulphite from isotope fractionation studies. Geochim. Cosmochim. Acta 32(1), 7191.Google Scholar
Kendall, B., Gordon, G.W., Poulton, S.W. & Anbar, A.D. (2011). Molybdenum isotope constraints on the extent of late Paleoproterozoic ocean euxinia. Earth Planet. Sci. Lett. 307, 450460.Google Scholar
Kump, L.R. & Holland, H.D. (1992). Iron in Precambrian rocks: implications for the global oxygen budget of the ancient Earth. Geochim. Cosmochim. Acta 56, 32173223.Google Scholar
Kurzweil, F., Claire, M., Thomazo, C., Peters, M., Hannington, M. & Strauss, H. (2013). Atmospheric sulfur rearrangement 2.7 billion years ago: evidence for oxygenic photosynthesis. Earth Planet. Sci. Lett. 366, 1726.Google Scholar
Labidi, J., Cartigny, P., Hamelin, C., Moreira, M. & Dosso, L. (2014). Sulfur isotope budget (32S, 33S, 34S and 36S) in Pacific–Antarctic ridge basalts: A record of mantle source heterogeneity and hydrothermal sulfide assimilation. Geochim. Cosmochim. Acta 133, 4767.Google Scholar
Leavitt, W.D., Halevy, I., Bradley, A.S. & Johnston, D.T. (2013). Influence of sulfate reduction rates on the Phanerozoic sulfur isotope record. Proc. Natl. Acad. Sci. U. S. A. 110, 1124411249.Google Scholar
Melezhik, V.A., Fallick, A.E., Rychanchik, D.V. & Kuznetsov, A.B. (2005). Paleoproterozoic vaporites in Fennoscandia: implications for seawater sulphate, the rise of atmospheric oxygen and local amplification of the _13C excursion. Terra Nova 17, 141148.Google Scholar
Millero, F.J. (2005). Chemical Oceanography, 3rd edn. CRC Press, Boca Raton, Florida.Google Scholar
Mojzsis, S.J., Coath, C.D., Greenwood, J.P., McKeegan, K.D. & Harrison, T.M. (2003). Mass-independent isotope effects in Archean (2.5 to 3.8 Ga) sedimentary sulfides determined by ion microprobe analysis. Geochim. Cosmochim. Acta 67, 16351658.CrossRefGoogle Scholar
Montinaro, A., Strauss, H., Mason, P.R.D., Roerdink, D., Münker, C., Schwarz-Schampera, U., Arndt, N.T., Farquhar, J., Beukes, N.J., Gutzmer, J. & Peters, M. (2015). Paleoarchean sulfur cycling: multiple sulfur isotope constraints from the Barberton Greenstone Belt, South Africa. Precambrian Res. 267, 311322.Google Scholar
Mossman, D.J., Minter, W.E.L., Dutkiewicz, A., Hallbauer, D.K., George, S.C., Hennigh, Q., Reimer, T.O. & Horscroft, F.D. (2008). The indigenous origin of Witwatersrand “carbon”. Precambrian Res. 164, 173186.Google Scholar
Ohmoto, H. & Goldhaber, M.B. (1997). Sulfur and carbon isotopes. Geochem. Hydrothermal Ore Deposits 3, 517611.Google Scholar
Ohmoto, H., Kakegawa, T. & Lowe, D.R. (1993). 3.4-billion-year-old biogenic pyrites from Barberton, South Africa: sulfur isotope evidence. Science 262, 555557.Google Scholar
Ohmoto, H., Watanabe, Y., Ikemi, H., Poulson, S.R. & Taylor, B.E. (2006). Sulphur isotope evidence for an oxic Archaean atmosphere. Nature 442, 908911.Google Scholar
Ohmoto, H., Watanabe, Y., Lasaga, A.C., Naraoka, H., Johnson, I., Brainard, J. & Chorney, A. (2014). Oxygen, iron, and sulfur geochemical cycles on early Earth: Paradigms and contradictions. Geol. Soc. Am. Spec. Pap. 504, 5595.Google Scholar
Oeser, M., Strauss, H., Wolff, P.E., Koepke, J., Peters, M., Garbe-Schönberg, D. & Dietrich, M. (2012). A profile of multiple sulfur isotopes through the Oman ophiolite. Chem. Geol. 312–313, 2746.Google Scholar
Ono, S., Eigenbrode, J.L., Pavlov, A.A., Kharecha, P., Rumble, D., Kasting, J.F. & Freeman, K.H. (2003). New insights into Archean sulfur cycle from mass-independent sulfur isotope records from the Hamersley Basin, Australia. Earth Planet. Sci. Lett. 213, 1530.Google Scholar
Ono, S., Shanks, W.C., Rouxel, O.J. & Rumble, D. (2007). S-33 constraints on the seawater sulfate contribution in modern seafloor hydrothermal vent sulfides. Geochim. Cosmochim. Acta 71, 11701182.Google Scholar
Ono, S., Keller, N.S., Rouxel, O. & Alt, J.C. (2012). Sulfur-33 constraints on the origin of secondary pyrite in altered oceanic basement. Geochim. Cosmochim. Acta 87, 323340.Google Scholar
Paytan, A., Kastner, M., Campbell, D. & Thiemens, M.H. (2004). Seawater sulfur isotope fluctuations in the Cretaceous. Science 304, 16631665.Google Scholar
Pavlov, A.A., Kasting, J.F. (2002). Mass-independent fractionation of sulphur isotopes in Archaean sediments: strong evidence for an anoxic Archaean atmosphere. Astrobiology 2, 2741.Google Scholar
Peters, M., Strauss, H., Farquhar, J., Ockert, C., Eickmann, B. & Jost, C.L. (2010). Sulfur cycling at the Mid-Atlantic Ridge: a multiple sulfur isotope approach. Chem. Geol. 269(3), 180196.Google Scholar
Philippot, P., Van Zuilen, M., Lepot, K., Thomazo, C., Farquhar, J. & Van Kranendonk, M.J. (2007). Early Archaean microorganisms preferred elemental sulfur, not sulfate. Science 317, 15341537.Google Scholar
Philippot, P., van Zuilen, M. & Rollion-Bard, C. (2012). Variations in atmospheric sulphur chemistry on early Earth linked to volcanic activity. Nat. Geosci. 5, 668674.Google Scholar
Pufahl, P.K. & Hiatt, E.E. (2012). Oxygenation of the Earth's atmosphere–ocean system: a review of physical and chemical sedimentologic responses. Mar. Pet. Geol. 32, 120.Google Scholar
Reinhard, C.T., Raiswell, R., Scott, C., Anbar, A.D. & Lyons, T.W. (2009). A late Archean sulfidic sea stimulated by early oxidative weathering of the continents. Science 326, 713716.Google Scholar
Roerdink, D.L., Mason, P.R.D., Farquhar, J. & Reimer, T. (2012). Multiple sulfur isotopes in Paleoarchean barites identify an important role for microbial sulfate reduction in the early marine environment. Earth Planet. Sci. Lett. 331–332, 177186.Google Scholar
Roerdink, D.L., Mason, P.R.D., Whitehouse, M.J. & Reimer, T. (2013). High-resolution quadruple sulfur isotope analyses of 3.2 Ga pyrite from the Barberton Greenstone Belt in South Africa reveal distinct environmental controls on sulfide isotopic arrays. Geochim. Cosmochim. Acta 117, 203215.Google Scholar
Rouxel, O., Shank, W.C. III, Bach, W., Edwards, K.J. (2008). Integrated Fe- and S-isotope study of seafloor hydrothermal vents at East Pacific Rise 9–10°N. Chem. Geol. 252, 214227.Google Scholar
Rye, R., Holland, H.D. (2000). Life associated with a 2.76 Ga ephemeral pond? Evidence from Mount Roe #2 paleosol. Geology 28, 483486.Google Scholar
Schidlowski, M. (1980). Antiquity and evolutionary status of bacterial sulfate reduction: sulfur isotope evidence. In Limits of Life, Vol. 4, ed. Ponnamperuma, C. & Margulis, L., pp. 159171. Springer, Netherlands. in Ponnamperuma, C., & Margulis, L. (Eds.). (1980). Limits of Life (Vol. 4). Springer Science & Business Media, Netherlands.Google Scholar
Scott, C.T., Bekker, A., Reinhard, C.T., Schnetger, B., Krapez, B., Rumble, D. & Lyons, T.W. (2011). Late Archean euxinic conditions before the rise of atmospheric oxygen. Geology 39, 119122.Google Scholar
Shen, Y. & Buick, R. (2004). The antiquity of microbial sulfate reduction. Earth-Sci. Rev. 64, 243272.Google Scholar
Shen, Y., Buick, R. & Canfield, D.E. (2001). Isotopic evidence for microbial sulphate reduction in the early Archaean era. Nature 410, 7781.Google Scholar
Shen, Y., Farquhar, J., Masterson, A., Kaufman, A.J. & Buick, R. (2009). Evaluating the role of microbial sulfate reduction in the early Archean using quadruple isotope systematics. Earth Planet. Sci. Lett. 279, 383391.Google Scholar
Sim, M.S., Bosak, T. & Ono, S. (2011). Large sulfur isotope fractionation does not require disproportionation. Science 333, 7477.Google Scholar
Stam, M.C., Mason, P.R.D., Pallud, C. & Van Cappellen, P. (2010). Sulfate reducing activity and sulfur isotope fractionation by natural microbial communities in sediments of a hypersaline soda lake (Mono Lake, California). Chem. Geol. 278, 2330.Google Scholar
Stam, M.C., Mason, P.R.D., Laverman, A.M., Pallud, C.L. & Cappellen, P.V. (2011). 34S/32S fractionation by sulfate-reducing microbial communities in estuarine sediments. Geochim. Cosmochim. Acta 75, 39033914.CrossRefGoogle Scholar
Strauss, H. (1999). Geological evolution from isotope proxy signals-sulfur. Chem. Geol. 161, 89101.Google Scholar
Strauss, H. (2003). Sulphur isotopes and the early Archean sulphur cycle. Precambrian Res. 126, 349361.Google Scholar
Strauss, H. (2004). 4 Ga of seawater evolution: evidence from the sulfur isotopic composition of sulfate. Geol. Soc. Am. Spec. Pap. 379, 195205.Google Scholar
Surkov, A.V., Böttcher, M.E. & Kuever, J. (2012). Sulphur isotope fractionation during the reduction of elemental sulphur and thiosulphate by Dethiosulfovibrio spp. Isot. Environ. Health Stud. 48(1), 6575.Google Scholar
Thomassot, E., Cartigny, P., Harris, J.W., Lorand, J.P., Rollion-Bard, C. & Chaussidon, M. (2009). Metasomatic diamond growth: A multi-isotope study (13 C, 15 N, 33 S, 34 S) of sulphide inclusions and their host diamonds from Jwaneng (Botswana). Earth Planet. Sci. Lett. 282, 7990.Google Scholar
Tostevin, R., Turchyn, A.V., Farquhar, J., Johnston, D.T., Eldridge, D.L., Bishop, J.K. & McIlvin, M. (2014). Multiple sulfur isotope constraints on the modern sulfur cycle. Earth Planet. Sci. Lett. 396, 1421.Google Scholar
Ueno, Y., Ono, S., Rumble, D. & Maruyama, S. (2008). Quadruple sulfur isotope analysis of ca. 3.5 Ga Dresser Formation: New evidence for microbial sulfate reduction in the early Archean. Geochim. Cosmochim. Acta 72, 56755691.Google Scholar
Wacey, D., McLoughlin, N., Whitehouse, M.J. & Kilburn, M.R. (2010). Two coexisting sulfur metabolisms in a ca. 3400 Ma sandstone. Geology 38, 11151118.Google Scholar
Wacey, D., Kilburn, M.R., Saunders, M., Cliff, J. & Brasier, M.D. (2011a). Microfossils of sulphur-metabolizing cells in 3.4-billion-year-old rocks of Western Australia. Nat. Geosci. 4, 698702.Google Scholar
Wacey, D., Saunders, M., Brasier, M.D. & Kilburn, M.R. (2011b). Earliest microbially mediated pyrite oxidation in~ 3.4 billion-year-old sediments. Earth Planet. Sci. Lett. 301, 393402.Google Scholar
Walker, J.C.G. (1986). Global geochemical cycles of carbon, sulfur and oxygen. Mar. Geol. 70, 159174.Google Scholar
Watanabe, Y., Martini, J.E.J. & Ohmoto, H. (2000). Geochemical evidence for terrestrial ecosystems 2.6 billion years ago. Nature 408, 574578.Google Scholar
Watanabe, Y., Farquhar, J. & Ohmoto, H. (2009). Anomalous fractionations of sulfur isotopes during thermochemical sulfate reduction. Science 324, 370373.Google Scholar
Whitehouse, M.J., Kamber, B.S., Fedo, C.M. & Lepland, A. (2005). Integrated Pb-and S-isotope investigation of sulphide minerals from the early Archaean of southwest Greenland. Chem. Geol. 222, 112131.Google Scholar
Zhao, B., Robb, L.J., Harris, C. & Jordaan, L.J. (2006). Origin of hydrothermal fluids and gold mineralization associated with the Ventersdorp contact reef, Witwatersrand Basin, South Africa: constraints from S, O, and H isotopes. In Processes on the Early Earth. Geological Society of America Special Paper 405, ed. Reimold, W.U. & Gibson, R.L., pp. 333352. Geological Society of America, Boulder, CO.Google Scholar
Zerkle, A.L., Farquhar, J., Johnston, D.T., Cox, R.P. & Canfield, D.E. (2009). Fractionation of multiple sulfur isotopes during phototrophic oxidation of sulfide and elemental sulfur by a green sulfur bacterium. Geochim. Cosmochim. Acta 73(2), 291306.Google Scholar