Hostname: page-component-8448b6f56d-jr42d Total loading time: 0 Render date: 2024-04-20T08:59:36.050Z Has data issue: false hasContentIssue false

Authigenesis of native sulphur and dolomite in a lacustrine evaporitic setting (Hellín basin, Late Miocene, SE Spain)

Published online by Cambridge University Press:  17 February 2011

J. LINDTKE
Affiliation:
MARUM, Universität Bremen, 28359 Bremen, Germany
S. B. ZIEGENBALG
Affiliation:
MARUM, Universität Bremen, 28359 Bremen, Germany
B. BRUNNER
Affiliation:
Max-Planck-Institut für Marine Mikrobiologie, 28359 Bremen, Germany
J. M. ROUCHY
Affiliation:
Département Histoire de la Terre, Muséum National d'Histoire Naturelle, 75005 Paris, France
C. PIERRE
Affiliation:
CNRS-UMR 7159, LOCEAN, Univ. P. and M. Curie, 75252 Paris Cedex 05, France
J. PECKMANN*
Affiliation:
Department für Geodynamik und Sedimentologie, Erdwissenschaftliches Zentrum, Universität Wien, Althanstraße 14, 1090 Wien, Austria
*
Author for correspondence: joern.peckmann@univie.ac.at

Abstract

Abundant sulphur is present in the Late Miocene evaporitic sequence of the lacustrine Hellín basin in SE Spain. Weathering of Triassic evaporites controlled the chemical composition of the Miocene lake. The lacustrine deposits comprise gypsum, marlstones, diatomites and carbonate beds. Sulphur-bearing carbonate deposits predominantly consist of early diagenetic dolomite. Abundant dolomite crystals with a spheroidal habit are in accordance with an early formation and point to a microbial origin. The carbon isotopic composition of the dolomite (δ13C values between −10 and −4‰) indicates mixing of lake water carbonate and carbonate derived from the remineralization of organic matter by heterotrophic bacteria. Dolomite precipitated syngenetically under evaporitic conditions as indicated by high oxygen isotope values (δ18O between +6 and +11‰). Nodules of native sulphur are found in gypsum, carbonate beds and marlstone layers. Sulphur formed in the course of microbial sulphate reduction, as reflected by its strong depletion in 34S (δ34S values as low as −17‰). Near to the surface many of the sulphur nodules were in part or completely substituted by secondary gypsum, which still reflects the sulphur isotopic composition of native sulphur (−18 to −10‰). This study exemplifies the role of bacterial sulphate reduction in the formation of dolomite and native sulphur in a semi-enclosed lacustrine basin during Late Miocene time.

Type
Original Article
Copyright
Copyright © Cambridge University Press 2011

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

Abd-El-Malek, Y. & Rizk, S. G. 1963 a. Bacterial sulphate reduction and development of alkalinity. I. Experiments with synthetic media. Journal of Applied Bacteriology 26, 713.CrossRefGoogle Scholar
Abd-El-Malek, Y. & Rizk, S. G. 1963 b. Bacterial sulphate reduction and development of alkalinity. II. Laboratory experiments with soils. Journal of Applied Bacteriology 26, 1419.CrossRefGoogle Scholar
Anadón, P., Rosell, L. & Talbot, M. R. 1992. Carbonate replacement of lacustrine gypsum deposits in two Neogene continental basins, eastern Spain. Sedimentary Geology 78, 201–16.CrossRefGoogle Scholar
Ayllón-Quevedo, F., Souza-Egipsy, V., Sanz-Montero, M. E. & Rodríguez-Aranda, J. P. 2007. Fluid inclusion analysis of twinned selenite gypsum beds from the Miocene of the Madrid basin (Spain). Implication on dolomite bioformation. Sedimentary Geology 201, 212–30.Google Scholar
Baker, P. A. & Kastner, M. 1981. Constraints on the formation of sedimentary dolomite. Science 213, 214–16.CrossRefGoogle ScholarPubMed
Bellanca, A., Calvo, J. P., Censi, P., Elizaga, E. & Neri, R. 1989. Evolution of lacustrine diatomite carbonate cycles of Miocene age, southeastern Spain: petrology and isotope geochemistry. Journal of Sedimentary Petrology 59, 4552.Google Scholar
Bolliger, C., Schroth, M. H., Bernasconi, S. M., Kleikemper, J. & Zeyer, J. 2001. Sulfur isotope fractionation during microbial sulfate reduction by toluene-degrading bacteria. Geochimica et Cosmochimica Acta 65, 3289–98.CrossRefGoogle Scholar
Böttcher, M. E. & Parafiniuk, J. 1998. Methane-derived carbonates in a native sulfur deposit: stable isotope and trace element discriminations related to the transformation of aragonite to calcite. Isotopes in Environmental and Health Studies 34, 177–90.CrossRefGoogle Scholar
Boudreau, B. P. 1991. Modelling the sulfide-oxygen reaction and associated pH gradients in porewaters. Geochimica et Cosmochimica Acta 55, 145–59.CrossRefGoogle Scholar
Brunner, B., Bernasconi, S. M., Kleikemper, J. A. & Schroth, M. H. 2005. A model for oxygen and sulfur isotope fractionation in sulfate during bacterial sulfate reduction processes. Geochimica et Cosmochimica Acta 69, 4773–85.Google Scholar
Burns, S. J., McKenzie, J. A. & Vasconcelos, C. 2000. Dolomite formation and biogeochemical cycles in the Phanerozoic. Sedimentology 47, 4961.CrossRefGoogle Scholar
Calvo, J. P. & Elizaga, E. 1989. Sedimentación evaporítica en las cuencas de Cenajo y Las Minas – Camarillas (región de Hellín, Mioceno Superior del área Prebética). In Formaciones evaporíticas de la Cuenca del Ebro y cadenas periféricas, y de la zona de Levante (eds Ortí, F. & Salvany, J. M.), pp. 246–50. Barcelona: University of Barcelona.Google Scholar
Calvo, J. P., Elizaga, E., Lopez Martinez, N., Robles, F. & Usera, J. 1978. El Mioceno superior continental del Prébetico Externo: Evolución del Estrecho Nordbético. Boletín Geológico y Minero 89, 407–26.Google Scholar
Canfield, D. E. 2001. Biogeochemistry of sulfur isotopes. In Stable Isotope Geochemistry. Mineralogical Society of America & Geochemical Society: Reviews in Mineralogy and Geochemistry (eds Valley, J. W. & Cole, D. R.), pp. 607–36. Washington, DC: Mineralogical Society of America.CrossRefGoogle Scholar
Castanier, S., Le Métayer-Levrel, G., Perthuisot, J. P. 1999. Ca-carbonate precipitation and limestone genesis – the microbiogeologist point of view. Sedimentary Geology 126, 923.CrossRefGoogle Scholar
Cerling, T. E., Bowman, J. R. & O'Neil, J. R. 1988. An isotopic study of a fluvial-lacustrine sequence: the Plio-Pleistocene Koobi Fora sequence, East Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 63, 335–56.CrossRefGoogle Scholar
Claypool, G. E., Holser, W. T., Kaplan, I. R., Sakai, H. & Zak, I. 1980. The age curves of sulfur and oxygen isotopes in marine sulfate and their mutual interpretation. Chemical Geology 28, 199260.Google Scholar
Davis, J. B. & Kirkland, D. W. 1979. Bioepigenetic sulfur deposits. Economic Geology and the Bulletin of the Society of Economic Geologists 74, 462–8.Google Scholar
Dessau, G., Jensen, M. L. & Nakai, N. 1962. Geology and isotopic studies of Sicilian sulfur deposits. Economic Geology 57, 410–38.CrossRefGoogle Scholar
Detmers, J., Brüchert, V., Habicht, K. S. & Kuever, J. 2001. Diversity of sulfur isotope fractionations by sulfate-reducing prokaryotes. Applied and Environmental Microbiology 67, 888–94.CrossRefGoogle ScholarPubMed
Dravis, J. J. & Yurewicz, D. A. 1985. Enhanced carbonate petrography using fluorescence microscopy. Journal of Sedimentary Petrology 55, 795804.Google Scholar
Dupraz, C., Reid, R. P., Braissant, O., Decho, A. W., Norman, R. S. & Visscher, P. T. 2008. Processes of carbonate precipitation in modern microbial mats. Earth Science Reviews 96, 141–62.CrossRefGoogle Scholar
Feely, H. W. & Kulp, J. L., 1957. Origin of Gulf coast salt-dome sulphur deposits. Bulletin of the American Association of Petroleum Geologists 41, 1802–53.Google Scholar
Folk, R. L. & Land, L. S. 1975. Mg/Ca ratio and salinity: two controls over crystallization of dolomite. Bulletin of the American Association of Petroleum Geologists 59, 60–8.Google Scholar
Füchtbauer, H. 1988. Sedimente und Sedimentgesteine, 4th ed. Stuttgart: Schweizerbart'sche Verlagsbuchhandlung, 1141 pp.Google Scholar
García-Alix, A., Minwer-Barakat, R., Suárez, E. M., Freudenthal, M. & Martín, J. M. 2008. Late Miocene–Early Pliocene climatic evolution of the Granada Basin (southern Spain) deduced from the paleoecology of the micromammal associations. Palaeogeography, Palaeoclimatology, Palaeoecology 265, 214–25.CrossRefGoogle Scholar
Garcia Domingo, A., Lopez Olmedo, F., Jerez Mir, L. & Gallego Coiduras, I. 1980. Mapa Geológico de España, hoja 868 Isso, Escala 1:50.000. Instituto Geológico y Minero de España.Google Scholar
Giese, P., Reutter, K. J., Jacobshagen, V. & Nicolich, R. 1982. Explosion seismic crustal studies in the Alpine Mediterranean region and their implications to tectonic processes. In Alpine Mediterranean Geodynamics (eds Berckhemer, H. & Hsü, K. J.), pp. 3973. Washington, DC: American Geophysical Union.CrossRefGoogle Scholar
Goldhaber, M. B. & Kaplan, I. R. 1974. The sedimentary sulphur cycle. In The Sea (ed. Goldberg, E. B.), pp. 569655. New York: Wiley.Google Scholar
Gray, N. D., Howarth, R., Pickup, R. W., Gwyn Jones, J. & Head, I. M. 1999. Substrate uptake by uncultured bacteria from the genus Achromatium determined by microautoradiography. Applied and Environmental Microbiology 65, 5100–6.Google Scholar
Gunatilaka, A. 1989. Spheroidal dolomites – origin by hydrocarbon seepage? Sedimentology 36, 701–10.Google Scholar
Hardie, L. A. 1987. Dolomitization; a critical view of some current views. Journal of Sedimentary Petrology 57, 166–83.Google Scholar
Hartmann, M. & Nielsen, H. 1969. δ34S-Werte in rezenten Meeressedimenten und ihre Deutung am Beispiel einiger Sedimentprofile aus der westlichen Ostsee. Geologische Rundschau 58, 621–55.Google Scholar
Hentz, T. F. & Henry, C. D. 1989. Evaporite-hosted native sulfur in Trans-Pecos Texas: relation to late-phase Basin and Range deformation. Geology 17, 400–3.2.3.CO;2>CrossRefGoogle Scholar
Hoefs, J. 2004. Stable Isotope Geochemistry, 5th ed. Berlin: Springer, 244 pp.CrossRefGoogle Scholar
Irwin, H., Curtis, C. & Coleman, M. 1977. Isotopic evidence for source of diagenetic carbonates formed during burial of organic-rich sediments. Nature 269, 209–13.CrossRefGoogle Scholar
Janaway, T. M. & Parnell, J. 1989. Carbonate production within the Orcadian basin, northern Scotland: a petrographic and geochemical study. Palaeogeography, Palaeoclimatology, Palaeoecology 70, 89105.CrossRefGoogle Scholar
Jassim, S. Z., Raiswell, R. & Bottrell, S. H. 1999. Genesis of the Middle Miocene stratabound sulphur deposits of northern Iraq. Journal of the Geological Society, London 156, 2539.Google Scholar
Jolivet, L., Augier, R., Robin, C., Suc, J.-P. & Rouchy, J. M. 2006. Lithospheric-scale geodynamic context of the Messinian salinity crisis. Sedimentary Geology 188–189, 933.CrossRefGoogle Scholar
Kahle, C. F. 1965. Possible roles of clay minerals in the formation of dolomite. Journal of Sedimentary Petrology 35, 448–53.Google Scholar
Kaplan, I. R. & Rittenberg, S. C. 1964. Microbiological fractionation of sulphur isotopes. Journal of General Microbiology 34, 195212.CrossRefGoogle ScholarPubMed
Kasten, S. & Jørgensen, B. B. 2000. Sulfate reduction in marine sediments. In Marine Geochemistry (eds Schulz, H. & Zabel, M.), pp. 263–81. Heidelberg: Springer.CrossRefGoogle Scholar
Khalaf, F. I. 1990. Occurrence of phreatic dolocrete within Tertiary clastic deposits of Kuwait, Arabian Gulf. Sedimentary Geology 68, 223–39.CrossRefGoogle Scholar
Krijgsman, W., Garcés, M., Agusti, J., Raffi, I., Taberner, C. & Zachariasse, W. J. 2000. The ‘Tortonian salinity crisis’ of the eastern Betics (Spain). Earth and Planetary Science Letters 181, 497511.CrossRefGoogle Scholar
Ku, T. C. W., Walter, L. M., Coleman, M. L., Blake, R. E. & Martini, A. M. 1999. Coupling between sulfur recycling and syndepositional carbonate dissolution: evidence from oxygen and sulfur isotope composition of pore water sulfate, South Florida platform, U.S.A. Geochimica et Cosmochimica Acta 63, 2529–46.CrossRefGoogle Scholar
Lalou, C. 1957. Studies on bacterial precipitation of carbonates in sea water. Journal of Sedimentary Petrology 27, 190–5.Google Scholar
Land, L. S. 1998. Failure to precipitate dolomite at 25°C from dilute solution despite 1000-fold oversaturation after 32 years. Aquatic Geochemistry 4, 361–8.CrossRefGoogle Scholar
Lee, C., McKenzie, J. A. & Sturm, M. 1987. Carbon isotope fractionation and changes in flux and composition of particulate matter resulting from biological activity during a sediment trap experiment in lake Greifen, Switzerland. Limnology and Oceanography 32, 8396.Google Scholar
Lumsden, D. N. 1979. Discrepancy between thin-section and x-ray estimates of dolomite in limestones. Journal of Sedimentary Petrology 49, 429–36.Google Scholar
Machel, H. G. 1992. Low-temperature and high-temperature origins of elemental sulfur in diagenetic environments. In Native Sulfur – Developments in Geology and Exploration (eds Wessel, G. R. & Wimberly, B. H.), pp. 322. Littleton, Colorado: Society for Mining, Metallurgy, and Exploration, Inc.Google Scholar
McKenzie, J. 1981. Holocene dolomitization of calcium carbonate sediments from the coastal sabkhas of Abu Dhabi, U.A.E.: a stable isotope study. Journal of Geology 89, 185–98.Google Scholar
McKenzie, J. 1985. Carbon isotopes and productivity in the lacustrine and marine environment. In Chemical Processes in Lakes. Environmental science and technology (ed. Stumm, W.), pp. 99118. New York: Wiley.Google Scholar
Nakai, N. & Jensen, M. L. 1964. The kinetic isotope effect in the bacterial reduction and oxidation of sulfur. Geochimica et Cosmochimica Acta 28, 1893–912.CrossRefGoogle Scholar
Navarro Hervás, F. & Rodríguez Estrella, T. 1985. Caracteristicas morfoestructurales de los diapiros Triasicos de Hellín, Ontur, La Celia, Jumilla, La Rosa y Pinoso, en las provincias de Albacete, Murcia y Alicante. Papeles de Geografica (Fisica) 10, 4969.Google Scholar
Ortí, F., Rosell, L. & Anadón, P. 2003. Deep to shallow lacustrine evaporites in the Libros Gypsum (southern Teruel Basin, Miocene, NE Spain): an occurrence of pelletal gypsum rhythmites. Sedimentology 50, 361–86.CrossRefGoogle Scholar
Ortí, F., Rosell, L. & Anadón, P. 2010. Diagenetic gypsum related to sulfur deposits in evaporites (Libros Gypsum, Miocene, NE Spain). Sedimentary Geology 228, 304–18.Google Scholar
Peckmann, J., Paul, J. & Thiel, V. 1999. Bacterially mediated formation of diagenetic aragonite and native sulfur in Zechstein carbonates (Upper Permian, Central Germany). Sedimentary Geology 126, 205–22.Google Scholar
Peckmann, J. & Thiel, V. 2004. Carbon cycling at ancient methane-seeps. Chemical Geology 205, 443–67.Google Scholar
Pierre, C. & Rouchy, J. M. 1988. Carbonate replacements after sulfate evaporites in the middle Miocene of Egypt. Journal of Sedimentary Research 58, 446–56.Google Scholar
Pirlet, H., Wehrmann, L. M., Brunner, B., Frank, N., Dewanckele, J., van Rooij, D., Foubert, A., Swennen, R., Naudts, L., Boone, M., Cnudde, V. & Henriet, J.-P. 2010. Diagenetic formation of gypsum and dolomite in a cold-water coral mound in the Porcupine Seabight, off Ireland. Sedimentology 57, 786805.Google Scholar
Playà, E., Ortí, F. & Rosell, L. 2000. Marine to non-marine sedimentation in the upper Miocene evaporites of the Eastern Betics, SE Spain: sedimentological and geochemical evidence. Sedimentary Geology 133, 135–66.CrossRefGoogle Scholar
Riccioni, R. M., Brock, P. W. G. & Schreiber, B. C. 1996. Evidence for early aragonite in paleo-lacustrine sediments. Journal of Sedimentary Research 66, 1003–10.Google Scholar
Rouchy, J. M., Taberner, C., Blanc-Valleron, M.-M., Sprovieri, R., Russell, M., Pierre, C., Di Stefano, E., Pueyo, J. J., Caruso, A., Dinarés-Turell, J., Gomis-Coll, E., Wolff, G. A., Cespuglio, G., Ditchfield, P., Pestrea, S., Combourieu-Nebout, N., Santisteban, C. & Grimalt, J. O. 1998. Sedimentary and diagenetic markers of the restriction in a marine basin: the Lorca Basin (SE Spain) during the Messinian. Sedimentary Geology 121, 2355.Google Scholar
Ruckmick, J. C., Wimberly, B. H. & Edwards, A. F. 1979. Classification and genesis of biogenic sulfur deposits. Economic Geology 74, 469–74.CrossRefGoogle Scholar
Sánchez-Román, M., McKenzie, J. A., de Luca Rebello Wagener, A., Rivadeneyra, M. A. & Vasconcelos, C. 2009. Presence of sulfate does not inhibit low-temperature dolomite formation. Earth and Planetary Science Letters 285, 131–9.Google Scholar
Sanz de Galdeano, C. 1990. Geologic evolution of the Betic Cordilleras in the Western Mediterranean, Miocene to the present. Tectonophysics 172, 107–19.Google Scholar
Sanz-Montero, M. E., Rodríguez-Aranda, J. P. & García del Cura, M. A. 2009. Bioinduced precipitation of barite and celestite in dolomite microbialites. Examples from Miocene lacustrine sequences in the Madrid and Duero Basins, Spain. Sedimentary Geology 222, 138–48.Google Scholar
Schulz, H. N. 2002. Thiomargarita namibiensis: giant microbe holding its breath. ASM News 68, 122–7.Google Scholar
Schulz, H. N. & Jørgensen, B. B. 2001. Big bacteria. Annual Review of Microbiology 55, 105–37.CrossRefGoogle ScholarPubMed
Schulz, H. N. & Schulz, H. D. 2005. Large sulfur bacteria and the formation of phosphorite. Science 307, 416–18.Google Scholar
Servant-Vildary, S., Rouchy, J. M., Pierre, C. & Foucault, A. 1990. Marine and continental water contributions to a hypersaline basin using diatom ecology, sedimentology and stable isotopes: an example in the Late Miocene of the Mediterranean (Hellin Basin, southern Spain). Palaeogeography, Palaeoclimatology, Palaeoecology 79, 189204.Google Scholar
Sharma, T. & Clayton, R. N. 1965. Measurement of O18/O16 ratios of total oxygen of carbonates. Geochimica et Cosmochimica Acta 29, 1347–53.CrossRefGoogle Scholar
Slaughter, M. & Hill, R. J. 1991. The influence of organic matter in organogenic dolomitization. Journal of Sedimentary Petrology 61, 296303.Google Scholar
Smith, B. N. & Epstein, S. 1971. Two categories of 13C/12C ratios for higher plants. Plant Physiology 47, 380–4.Google Scholar
Stabel, H. H. 1986. Calcite precipitation in Lake Constance: chemical equilibrium, sedimentation, and nucleation by algae. Limnology and Oceanography 31, 1081–94.Google Scholar
Talbot, M. R. 1990. A review of the palaeohydrological interpretation of carbon and oxygen isotopic ratios in primary lacustrine carbonates. Chemical Geology 80, 261–79.Google Scholar
Talbot, M. R. & Kelts, K. 1990. Paleolimnological signatures from carbon and oxygen isotopic ratios in carbonates from organic carbon-rich lacustrine sediments. In Lacustrine Basin Exploration – Case studies and modern analogs (ed. Katz, B. J.), pp. 99112. Tulsa: American Association of Petroleum Geologists Memoir.Google Scholar
Taylor, B. E. & Wheeler, M. C. 1984. Stable isotope geochemistry of acid mine drainage: experimental oxidation of pyrite. Geochimica et Cosmochimica Acta 48, 2669–78.Google Scholar
Tenzer, G., Meyers, P. A. & Knoop, P. 1997. Sources and distribution of organic and carbonate carbon in surface sediments of Pyramid Lake, Nevada. Journal of Sedimentary Research 67, 884–90.Google Scholar
Thode, H. G. & Monster, J. 1973 (reprint of 1965). Sulfur-isotope geochemistry of petroleum, evaporites, and ancient seas. In Marine Evaporites: Origin, diagenesis and geochemistry. Benchmark Papers in Geology (eds Kirkland, D. W. & Evans, R.), pp. 363–73. Stroudsburg, Pennsylvania: Dowden, Hutchinson & Ross, Inc.Google Scholar
Utrilla, R., Pierre, C., Ortí, F. & Pueyo, J. J. 1992. Oxygen and sulphur isotope compositions as indicators of the origin of Mesozoic and Cenozoic evaporites from Spain. Chemical Geology 102, 229–44.Google Scholar
van Dam, J. A. & Weltje, G. J. 1999. Reconstruction of the Late Miocene climate of Spain using rodent palaeocommunity successions: an application of end-member modelling. Palaeogeography, Palaeoclimatology, Palaeoecology 151, 267305.CrossRefGoogle Scholar
van Lith, Y., Warthmann, R., Vasconcelos, C. & McKenzie, J. 2003. Sulphate reducing bacteria induce low-temperature Ca-dolomite and high Mg-calcite formation. Geobiology 1, 71–9.CrossRefGoogle Scholar
Vasconcelos, C. & McKenzie, J. A. 1997. Microbial mediation of modern dolomite precipitation and diagenesis under anoxic conditions (Lagoa Vermelha, Rio de Janeiro, Brazil). Journal of Sedimentary Research 67, 378–90.Google Scholar
Vasconcelos, C., McKenzie, J., Bernasconi, S. M., Crujic, D. & Tien, A. J. 1995. Microbial mediation as possible mechanism for natural dolomite formation at low temperatures. Nature 377, 220–2.Google Scholar
Vasconcelos, C., McKenzie, J. A., Warthmann, R. & Bernasconi, S. M. 2005. Calibration of the δ18O paleothermometer for dolomite precipitated in microbial cultures and natural environments. Geology 33, 317–20.Google Scholar
Walter, L. M., Bischof, S. A., Patterson, W. P. & Lyons, T. W. 1993. Dissolution and recrystallization in modern shelf carbonates: evidence from pore water and solid phase chemistry. Philosophical Transactions of the Royal Society of London Series A–Mathematical Physical and Engineering Sciences 344, 2736.Google Scholar
Warren, J. 2000. Dolomite: occurrence, evolution and economically important associations. Earth-Science Reviews 52, 181.CrossRefGoogle Scholar
Weaver, C. E. & Beck, K. C. 1971. Clay Water Diagenesis During Burial: How mud becomes gneiss. Geological Society of America, Special Paper 134. Boulder: Geological Society of America, 96 pp.Google Scholar
Wright, D. T. 1999. The role of sulphate-reducing bacteria and cyanobacteria in dolomite formation in distal ephemeral lakes of the Coorong region, South Australia. Sedimentary Geology 126, 147–57.Google Scholar
Wright, D. T. & Wacey, D. 2005. Precipitation of dolomite using sulphate-reducing bacteria from the Coorong Region, South Australia: significance and implications. Sedimentology 52, 9871008.Google Scholar
Youssef, E. S. A. A. 1989. Geology and genesis of sulfur deposits at Ras Gemsa area, Red Sea coast, Egypt. Geology 17, 797801.Google Scholar
Ziegenbalg, S. B., Brunner, B., Rouchy, J. M., Birgel, D., Pierre, C., Böttcher, M. E., Caruso, A., Immenhauser, A. & Peckmann, J. 2010. Formation of secondary carbonates and native sulphur in sulphate-rich Messinian strata, Sicily. Sedimentary Geology 227, 3750.CrossRefGoogle Scholar
Supplementary material: File

Lindtke Supplementary Material

Lindtke Supplementary Appendix

Download Lindtke Supplementary Material(File)
File 869.9 KB