Hostname: page-component-8448b6f56d-jr42d Total loading time: 0 Render date: 2024-04-18T01:38:47.149Z Has data issue: false hasContentIssue false

Sedimentary facies analyses from nano- to millimetre scale exploring past microbial activity in a high-altitude lake (Lake Son Kul, Central Asia)

Published online by Cambridge University Press:  02 March 2015

MURIEL PACTON*
Affiliation:
Laboratoire de Géologie de Lyon: Terre, Planètes, Environnement (UMR 5276 CNRS), Université Claude Bernard–Lyon 1, Villeurbanne, France
PHILIPPE SORREL*
Affiliation:
Laboratoire de Géologie de Lyon: Terre, Planètes, Environnement (UMR 5276 CNRS), Université Claude Bernard–Lyon 1, Villeurbanne, France
BENOÎT BEVILLARD
Affiliation:
Laboratoire de Géologie de Lyon: Terre, Planètes, Environnement (UMR 5276 CNRS), Université Claude Bernard–Lyon 1, Villeurbanne, France
AXELLE ZACAÏ
Affiliation:
Laboratoire de Géologie de Lyon: Terre, Planètes, Environnement (UMR 5276 CNRS), Université Claude Bernard–Lyon 1, Villeurbanne, France
ARNAULD VINÇON-LAUGIER
Affiliation:
Laboratoire de Géologie de Lyon: Terre, Planètes, Environnement (UMR 5276 CNRS), Université Claude Bernard–Lyon 1, Villeurbanne, France
HEDI OBERHÄNSLI
Affiliation:
Helmholtz-Centre Potsdam, German Geoscience Research Centre (GFZ), Section 5.2, Telegrafenberg, D-14473 Potsdam, Germany Museum für Naturkunde, Leibnitz-Institute Berlin (Mineralogy), Invalidenstrasse 43, 10115 Berlin, Germany

Abstract

The fabric of sedimentary rocks in lacustrine archives usually contains long and continuous proxy records of biological, chemical and physical parameters that can be used to study past environmental and climatic variability. Here we propose an innovative approach to sedimentary facies analysis based on a coupled geomicrobiological and sedimentological study using high-resolution microscopic techniques in combination with mineralogical analyses. We test the applicability of this approach on Lake Son Kul, a high alpine lake in central Tien Shan (Kyrgyzstan) by looking at the mineral fabric and microbial communities observed down to the nanoscale. The characterization of microbe–mineral interactions allows the origin of four carbonate minerals (e.g. aragonite, dolomite, Mg-calcite, calcite) to be determined as primary or diagenetic phases in Lake Son Kul. Aragonite was mainly of primary origin and is driven by biological activity in the epilimnion, whereas diagenetic minerals such as Mg-calcite, calcite, dolomite and pyrite were triggered by bacterial sulphate reduction and possibly by methanotrophic archaea. A new morphotype of aragonite (i.e. spherulite-like precursor) occurs in Unit IV (c. 7100–5000 cal. BP) associated with microbial mat structures. The latter enhanced the preservation of viral relics, which have not yet been reported in Holocene lacustrine sediments. This study advocates that microbe–mineral interactions screened down to the nanoscale (e.g. virus-like particles) can be used successfully for a comprehensive description of the fabric of laminated lake sediments. In this sense, they complement traditional facies sedimentology tools and offer valuable new insights into: (1) the study of microbial and viral biosignatures in Quaternary sediments; and (2) palaeoenvironmental reconstructions.

Type
Original Articles
Copyright
Copyright © Cambridge University Press 2015 

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.)

Footnotes

*

These authors contributed equally to this work

References

Amann, R. I., Ludwig, W. & Schleifer, K-H. 1995. Phylogenetic identification and in situ detection of individual microbial cells without cultivation. Microbiological Reviews 59, 143–69.Google Scholar
Amann, R. I., Krumholz, L. & Stahl, D. A. 1990. Fluorescent-oligonucleotide probing of whole cells for determinative, phylogenetic, and environmental studies in microbiology. Journal of Bacteriology 172, 762–70.Google Scholar
Andreassen, J.-P., Beck, R. & Nergaard, M. 2012. Biomimetic type morphologies of calcium carbonate grown in absence of additives. Faraday Discussion 159, 247–61.Google Scholar
Arp, G., Helms, G., Karlinska, K., Schumann, G., Reimer, A., Reitner, J. & Trichet, J. 2012. Photosynthesis versus exopolymer degradation in the formation of microbialites on the atoll of Kiritimati, Republic of Kiribati, Central Pacific. Geomicrobiology Journal 29, 2965.Google Scholar
Banfield, J. F., Moreau, J. W., Chan, C. S., Welch, S. A. & Little, B. 2001. Mineralogical biosignatures and the search for life on Mars. Astrobiology 1, 447–63.Google Scholar
Beloglasova, V. N. & Smirnova, N. B. 1987. Altas Kirgizskoj SSR. GUGK SSSR, Bishkek (in Russian).Google Scholar
Berner, R. A. 1984. Sedimentary pyrite formation: An update. Geochimica et Cosmochimica Acta 48, 605–15.Google Scholar
Bettarel, Y., Bouvy, M., Dumont, C. & Sime-Ngando, T. 2006. Virus-bacterium interactions in water and sediment of West African inland aquatic systems. Applied and Environmental Microbiology 72, 5274–82.Google Scholar
Betts-Piper, A. M., Zeeb, B. A. & Smol, J. P. 2004. Distribution and autoecology of chrysophyte cysts from high Arctic Svalbard lakes: preliminary evidence of recent environmental change. Journal of Paleolimnology 31, 467–81.Google Scholar
Bird, D. F., Juniper, S. K., Ricciardi-Rigault, M., Martineu, P., Prairie, Y. T. & Calvert, S. E. 2001. Subsurface viruses and bacteria in Holocene/Late Pleistocene sediments of Saanich Inlet, BC: ODP holes 1033B and 1043B, Leg 169S. Marine Geology 174, 227–39.Google Scholar
Bird, M. I., Chivas, A. R., Radnell, C. J. & Burton, H. R. 1991. Sedimentological and stable-isotope evolution of lakes in the Vestfold Hills, Antarctica. Palaeogeography, Palaeoclimatology, Palaeoecology 84, 109–30.Google Scholar
Blowes, D. W. & Jambor, J. L. 1990. The pore-water geochemistry and the mineralogy of the vadose zone of sulfide tailings, Waite Amulet, Quebec, Canada. Applied Geochemistry 5, 327–46.CrossRefGoogle Scholar
Boetius, A., Ravenschlag, K., Schubert, C., Rickert, D., Widdel, F., Gieseke, A., Amann, R., Jørgensen, B. B., Witte, U. & Pfannkuche, O. 2000. A marine microbial consortium apparently mediating anaerobic oxidation of methane. Nature 407, 623–6.Google Scholar
Bontognali, T. R. R., McKenzie, J. A., Warthmann, R. J. & Vasconcelos, C. 2014. Microbially influenced formation of Mg-calcite and Ca-dolomite in the presence of exopolymeric substances produced by sulphate-reducing bacteria. Terra Nova 26, 72–7.Google Scholar
Boss, S. K. & Neumann, A. C. 1993. Physical versus chemical processes of whiting formation in the Bahamas. Carbonates Evaporites 8, 135–48.CrossRefGoogle Scholar
Brauer, A., Allen, J. R. M., Mingram, J., Dulski, P., Wulf, S. & Huntley, B. 2007. Evidence for last interglacial chronology and environmental change from southern Europe. PNAS 104 (2), 450–55.Google Scholar
Buckley, D. H., Baumgartner, L. K. & Visscher, P. T. 2008. Vertical distribution of methane metabolism in microbial mats of the Great Sippewissett Salt Marsh. Environmental Microbiology 10, 967–77.Google Scholar
Cabala, J. & Piatek, M. 2004. Chrysophycean stomatocysts from the Staw Toporowy Nizni lake (Tatra National Park, Poland). Annales de Limnologie – International Journal of Limnology 40, 149–65.Google Scholar
Chanton, J. L., Chaser, P., Glasser, D. & Siegel, 2005. Carbon and hydrogen isotopic effects in microbial methane from terrestrial environments. Chapter 6. In Stable Isotopes and Biosphere - Atmosphere Interactions: Processes and Biological Controls (eds Flanagan, L. B., Ehleringer, J. R. & Pataki, D. E.), pp. 85112. Amsterdam: Elsevier.Google Scholar
Charpentier, D., Mosser-Ruck, R., Cathelineau, M. & Guillaume, D. 2004. Oxidation of mudstone in a tunnel (Tournemire, France): consequences on mineralogy and crystal chemistry of clay minerals. Clay Mineralogy 39, 135–49.Google Scholar
Charvet, S., Vincent, W. F. & Lovejoy, C. 2012. Chrysophytes and other protists in High Arctic lakes: molecular gene surveys, pigment signatures and microscopy. Polar Biology 35, 733–48.Google Scholar
Dean, W., Rosenbaun, J., Skipp, G., Colman, S., Forester, R., Liu, A., Simmons, K. & Bischoff, J. 2006. Unusual Holocene and late Pleistocene carbonate sedimentation in Bear Lake, Utah and Idaho, USA. Sedimentary Geology 185, 93112.Google Scholar
DeGroot, K. 1965. Inorganic precipitation of calcium carbonate production from seawater. Nature 207, 404–5.Google Scholar
Deng, S., Dong, H., Lv, G., Jiang, H., Yu, B. & Bishop, M. E. 2010. Microbial dolomite precipitation using sulfate reducing and halophilic bacteria: Results from Qinghai Lake, Tibetan Plateau, NW China. Chemical Geology 278, 151–9.Google Scholar
Duff, K. E. & Smol, J. P. 1991. Morphological descriptions and stratigraphic distributions of the chrysophycean stomatocysts from a recently acidified lake (Adirondack Park, N.Y.). Journal of Paleolimnology 5, 73113.Google Scholar
Duff, K. E., Zeeb, B. A. & Smol, J. P. 1995. Atlas of Chrysophycean Cysts. Dordrecht: Kluwer Academic Publishers, 189 pp.Google Scholar
Duhamel, S. & Jacquet, S. 2006. Flow cytometric analysis of bacteria- and virus-like particles in lake sediments. Journal of Microbiological Methods 64, 316–22.Google Scholar
Dupraz, C., Reid, R. P., Braissant, O., Decho, A. W., Norman, R. S. & Visscher, P. T. 2009. Processes of carbonate precipitation in modern microbial mats. Earth Science Reviews 6, 141–62.Google Scholar
Edwards, K. J., Bond, P. L., Druschel, G. K., McGuire, M. M., Hamers, R. J. & Banfield, J. F. 2000. Geochemical and biological aspects of sulfide mineral dissolution: lessons from Iron Mountain, California. Chemical Geology 169, 383–97.Google Scholar
Facher, E. & Schmidt, R. 1996. A siliceous chrysophycean cyst-based pH transfer function for Central European Lakes. Journal of Paleolimnology 16, 275321.Google Scholar
Fischer, U. R., Wieltschnig, C., Kirschner, A. K. T. & Velimirov, B. 2003. Does virus-induced lysis contribute significantly to bacterial mortality in the oxygenated sediment layer of shallow oxbow lakes? Applied and Environmental Microbiology 69, 5281–9.Google Scholar
Flügel, E. 2004. Microfacies of Carbonate Rocks. Analysis, Interpretation and Application. Berlin, Heidelberg, New York: Springer, 976 pp.Google Scholar
Francus, P., Suchodoletz, H., Dietze, M., Donner, R. V., Bouchard, F., Roy, A-J., Fagot, M., Verschuren, D. & Kröpelin, S. 2013. Varved sediments of Lake Yoa (Ounianga Kebir, Chad) reveal progressive drying of the Sahara during the last 6100 years. Sedimentology 60, 911–34.Google Scholar
Fuhrman, J. A. 1999. Marine viruses and their biogeochemical and ecological effects. Nature 399, 541–8.Google Scholar
Gaudin, A., Buatier, M. D., Beaufort, D., Petit, S., Grauby, O. & Decareau, A. 2005. Characterization and origin of Fe3+-Montmorillonite in deep water calcareous sediments (Pacific ocean, Costa Rica margin). Clays and Clay Minerals 53, 452–65.Google Scholar
Giralt, S., Julia, R. & Klerkx, J. 2001. Microbial biscuits of vaterite in Lake Issyk-Kul (Republic of Kyrgyzstan). Journal of Sedimentary Research 71, 430–5.CrossRefGoogle Scholar
Gorby, Y. A., Yanina, S., McLean, J. S., Rosso, K. M., Moyles, D., Dohnalkova, A., Beveridge, T. J., Chang, I. S., Kim, B. H., Kim, K. S., Culley, D. E., Reed, S. B., Romine, M. F., Saffarini, D. A., Hill, E. A., Shi, L., Elias, D. A., Kennedy, D. W., Pinchuk, G., Watanabe, K., Ishii, S., Logan, B., Nealson, K. H. & Fredrickson, J. K. 2006. Electrically conductive bacterial nanowires produced by Shewanella oneidensis strain MR-1 and other microorganisms. Proceedings of the National Academy of Sciences of the USA 103, 11358–63.Google Scholar
Habicht, K. S. & Canfield, D. E. 1997. Sulfur isotope fractionation during bacterial sulfate reduction in organic-rich sediments. Geochimica et Cosmochimica Acta 61, 5351–61.Google Scholar
Hansen, L. B., Finster, K., Fossing, H. & Iversen, N. 1998. Anaerobic methane oxidation in sulfate depleted sediments: effects of sulfate and molybdate additions. Aquatic Microbial Ecology 14, 195204.Google Scholar
Hendy, C. H. 2000. Late Quaternary lakes in the McMurdo Sound region of Antarctica. Geografiska Annaler 82A, 411–32.Google Scholar
Hendy, C. H., Healy, T. R., Rayner, E. M., Shaw, J. & Wilson, A. T. 1979. Late Pleistocene glacial chronology of the Taylor Valley, Antarctica, and the global climate. Quaternary Research 11, 172–84.Google Scholar
Hinrichs, K. U., Hayes, J. M., Sylva, S. P., Brewer, P. G. & DeLong, E. F. 1999. Methane consuming archaebacteria in marine sediments. Nature 398 (6730), 802–5.Google Scholar
Hodell, D. A., Schelske, C. L., Fahnenstiel, G. L. & Robbins, L. L. 1998. Biologically induced calcite and its isotopic composition in Lake Ontario. Limnology and Oceanography 43, 187–99.Google Scholar
Holmgren, S. K. 1984. Experimental lake fertilization in the Kuokkel area, Northern Sweden: Phytoplankton biomass and algal composition in natural and fertilized subarctic lakes. Internationale Revue der gesamten Hydrobiologie 69, 781817.Google Scholar
Huang, X., Oberhänsli, H., von Suchodoletz, H., Prasad, S., Sorrel, P., Plessen, B., Mathis, M. & Usubaliev, R. 2014. Hydrological changes in western Central Asia (Kyrgyzstan) during the Holocene: Results of a paleolimnological study from Son Kul. Quaternary Science Reviews 103, 134–52.Google Scholar
Jahren, A. H., LePage, B. A. & Werts, S. P. 2004. Methanogenesis in Eocene Arctic soils inferred from δ13C of tree fossil carbonates. Palaeogeography, Palaeoclimatology Palaeoecology 214, 347–58.Google Scholar
Kelts, K. & Hsü, K. J. 1978. Freshwater carbonate sedimentation. In: Lakes, Chemistry, Geology, Physics (ed. Lerman, A.), pp. 295323. New York: Springer-Verlag.Google Scholar
Kenward, P. A., Goldstein, R. H., Gonzalez, L. A. & Roberts, J. A. 2009. Precipitation of low-temperature dolomite from an anaerobic microbial consortium: the role of methanogenic Archaea. Geobiology 7 (5), 556–65.Google Scholar
Knittel, K., Lösekann, T., Boetius, A., Kort, R. & Amann, R. 2005. Diversity and distribution of methanotrophic archaea at cold seeps. Applied and Environmental Microbiology 71, 467–79.Google Scholar
Koschel, R. 1997. Structure and function of pelagic calcite precipitation in lake ecosystems. Verhandlungen des Internationalen Vereins für Limnologie 26, 343–9.Google Scholar
Koschel, R., Benndorf, J., Proft, G. & Recknagel, F. 1983. Calcite precipitation as a natural control mechanism of eutrophication. Archiv für Hydrobiologie 98 (3), 380408.Google Scholar
Kyle, J. E., Pedersen, K. & Ferris, F. G. 2008. Virus mineralization at low pH in the Rio Tinto, Spain. Geomicrobiology Journal 25, 338–45.Google Scholar
Land, L. S. 1998. Failure to precipitate dolomite at 25 °C from dilute solutions despite 1000-fold oversaturation after 32 years. Aquatic Geochemistry 4, 361–8.Google Scholar
Lauterbach, S., Witt, R., Plessen, B., Dulski, P., Prasad, S., Mingram, J., Gleixner, G., Hettler-Riedel, S., Stebich, M., Schnetger, B., Schwalb, A. & Schwarz, A. 2014. Climatic imprint of the mid-latitude Westerlies in the Central Tien Shan of Kyrgyzstan and teleconnections to North Atlantic climate variability during the last 6000 years. The Holocene 24 (8), 970–84.Google Scholar
Lawrence, M. J. F. & Hendy, C. H. 1985. Water column and sediment characteristics of Lake Fryxell, Taylor Valley, Antarctica. New Zealand Journal of Geology and Geophysics 28, 543–52.Google Scholar
Lemke, M., Wickstrom, C. & Leff, L. 1997. Preliminary study on the distribution of viruses and bacteria in lotic environments. Archiv für Hydrobiologie 141, 6774.Google Scholar
Lippmann, F. 1973. Sedimentary Carbonate Minerals. Berlin: Springer-Verlag.Google Scholar
MacLean, L. C., Tyliszczak, T., Gilbert, P. U., Zhou, D., Pray, T. J., Onstott, T. C. & Southam, G. 2008. A high-resolution chemical and structural study of framboidal pyrite formed within a low-temperature bacterial biofilm. Geobiology 6, 471–80.Google Scholar
Maranger, R. & Bird, D. F. 1996. High concentrations of viruses in the sediments of Lac Gilbert, Québec. Microbial Ecology 31, 141–51.Google Scholar
Mathis, M., Sorrel, P., Klotz, S., Huang, X. & Oberhänsli, H. 2014. Regional vegetation patterns at Lake Son Kul reveal Holocene climatic variability in central Tien Shan (Kyrgyzstan, Central Asia). Quaternary Science Reviews 89, 169–85.Google Scholar
McKenzie, J. A. & Vasconcelos, C. 2009. Dolomite Mountains and the origin of the dolomite rock of which they mainly consist: historical developments and new perspectives. Sedimentology 56, 205–19.Google Scholar
Meister, P. 2013. Two opposing effects of sulfate reduction on carbonate precipitation in normal marine, hypersaline, and alkaline environments. Geology 41, 499502.Google Scholar
Meulepas, R. J. W., Jagersma, C. G., Khadem, A. F., Stams, A. J. M. & Lens, P. N. L. 2010. Effect of methanogenic substrates on anaerobic oxidation of methane and sulfate reduction by an anaerobic methanotrophic enrichment. Applied Microbiology and Biotechnology 87, 1499–506.Google Scholar
Middelboe, M., Glud, R. N. & Filippini, M. 2011. Viral abundance and activity in the deep sub-seafloor biosphere. Aquatic Microbial Ecology 63, 18.Google Scholar
Middelboe, M. & Jørgensen, N. O. G. 2006. Viral lysis of bacteria: an important source of dissolved amino acids and cell wall components. Journal of the Marine Biological Association of the UK 86, 605–12.Google Scholar
Milliman, J. D., Freile, D., Steinen, R. P. & Wilber, R. J. 1993. Great Bahama Bank aragonitic muds: mostly inorganically precipitated, mostly exported. Journal of Sedimentary Petrology 63, 589–95.Google Scholar
Moreira, N. F., Walter, L. M., Vasconcelos, C., McKenzie, J. A. & McCall, P. J. 2004. Role of sulfide oxidation in dolomitization: sediment and pore-water geochemistry of a modern hypersaline lagoon system. Geology 32 (8), 701–4.Google Scholar
Morse, J. W., Gledhill, D. K. & Millero, F. J. 2003. CaCO3 precipitation kinetics in waters from the great Bahama Bank: implications for the relationshipbetweenbankhydrochemistry and whitings. Geochimica et Cosmochimica Acta 67, 2819–26.Google Scholar
Müller, G., Irion, G. & Förstner, U. 1972. Formation and diagenesis of inorganic Ca-Mg carbonates in the lacustrine environment. Naturwissenschaften 59, 158–64.CrossRefGoogle Scholar
Nauhaus, K., Treude, T., Boetius, A. & Krüger, M. 2005. Environmental regulation of the anaerobic oxidation of methane: a comparison of ANME-I and ANME-II communities. Environmental Microbiology 7, 98106.Google Scholar
Neugebauer, I., Brauer, A., Dräger, N., Dulski, P., Wulf, S., Plessen, B., Mingram, J., Herzschuh, U. & Brande, A. 2012. A Younger Dryas varve chronology from the Rehwiese palaeolake record in NE Germany. Quaternary Science Reviews 36, 91102.Google Scholar
Niemann, H., Lösekann, T., de Beer, D., Elvert, M., Nadalig, T., Knittel, K., Amann, R., Sauter, E. J., Schlüter, M., Klages, M., Foucher, J. P. & Boetius, A. 2006. Novel microbial communities of the Haakon Mosby mud volcano and their role as methane sink. Nature 443, 854–8.Google Scholar
Ohfuji, H. & Rickard, D. 2005. Experimental syntheses of framboids – a review. Earth-Science Reviews 71, 147–70.Google Scholar
Orange, F., Chabin, A., Gorlas, A., Lucas-Staat, S., Geslin, C., Le Romancer, M., Prangishvili, D., Forterre, P. & Westall, F. 2011. Experimental fossilisation of viruses from extremophilic Archaea. Biogeosciences 8, 1465–75.Google Scholar
Orphan, V. J., House, C. H., Hinrichs, K. U., McKeegan, K. D. & DeLong, E. F. 2001. Methane consuming archaea revealed by directly coupled isotopic and phylogenetic analysis. Science 293 (5529), 484–7.Google Scholar
Pacton, M., Fiet, N. & Gorin, G. 2006. Revisiting amorphous organic matter in Kimmeridgian laminites: what is the role of the vulcanization process in the amorphization of organic matter? Terra Nova 18, 380–7.Google Scholar
Pacton, M., Fiet, N. & Gorin, G. 2007. Bacterial activity and preservation of sedimentary organic matter: the role of exopolymeric substances. Geomicrobiology Journal 24, 571–81.Google Scholar
Pacton, M., Fiet, N. & Gorin, G. 2008. Unravelling the origin of ultralaminae in sedimentary organic matter: the contribution of bacteria and photosynthetic organisms. Journal of Sedimentary Research 78, 654–67.CrossRefGoogle Scholar
Pacton, M., Gorin, G. & Vasconcelos, C. 2011. Amorphous organic matter – experimental data on formation and the role of microbes. Review of Palaeobotany and Palynology 166, 253–67.Google Scholar
Pacton, M., Wacey, D., Corinaldesi, C., Tangherlini, M., Kilburn, M. R., Gorin, G., Danovaro, R. & Vasconcelos, C. 2014. Viruses as new agents of organomineralization in the geological record. Nature Communications 5, 4298.Google Scholar
Paulo, C. & Dittrich, M. 2013. 2D Raman spectroscopy study of dolomite and cyanobacterial extracellular polymeric substances from Khor Al-Adaid sabkha (Qatar). Journal of Raman Spectroscopy 44, 1563–9.Google Scholar
Peng, X., Xu, H., Jones, B., Chen, S. & Zhou, H. 2013. Silicified virus-like nanoparticles in an extreme thermal environment: implications for the preservation of viruses in the geological record. Geobiology 11, 511–26.Google Scholar
Pienitz, R., Walker, I. R., Zeeb, B. A., Smol, J. P. & Leavitt, P. R. 1992. Biomonitoring past salinity changes in an athalassic sub-Arctic lake. International Journal of Salt Lake Research 1, 91123.Google Scholar
Pirbadian, S., Barchinger, S. E., Leung, K. M., Byun, H. S., Jangir, Y., Bouhenni, R. A., Reed, S. B., Romine, M. F., Saffarini, D. A., Shi, L., Gorby, Y. A., Golbeck, J. H. & El-Naggar, M. Y. 2014. Shewanella oneidensis MR-1 nanowires are outer membrane and periplasmic extensions of the extracellular electron transport components. Proceedings of the National Academy of Sciences of the United States of America 111, 12883–8.Google Scholar
Popa, R., Badescu, A. & Kinkle, B. K. 2004. Pyrite framboids as biomarkers for iron-sulfur systems. Geomicrobiology Journal 21, 114.Google Scholar
Prasad, V., Garg, R., Singh, V. & Thakur, B. 2007. Organic matter distribution pattern in Arabian Sea: Palynofacies analysis from the surface sediments off Karwar coast (west coast of India). Indian Journal of Marine Sciences 36, 399406.Google Scholar
Reguera, G., McCarthy, K. D., Metha, T., Nicoll, J. S., Tuominen, M. T. & Lovley, D. R. 2005. Extracellular electron transfer via microbial nanowires. Nature 435, 1098–101.Google Scholar
Roberts, J. A., Bennett, P. C., Gonzalez, L. A., Macpherson, G. L. & Milliken, K. L. 2004. Microbial precipitation of dolomite in methanogenic groundwater. Geology 32 (4), 277–80.Google Scholar
Sanchez-Navas, A., Martìn-Algarra, A., Rivadeneyra, M. A., Melchor, S., Martìn-Ramos, J. D. 2009. Crystal-growth behaviour in Ca–Mg carbonate bacterial spherulites. Crystal Growth and Design 9, 2690–9.Google Scholar
Sanchez-Roman, M., Vasconcelos, C., Schmid, T., Dittrich, M., McKenzie, J. A., Zenobi, R. & Rivadeneyra, M. A. 2008. Aerobic microbial dolomite at the nanometer scale: implications for the geologic record. Geology 36, 879–82.Google Scholar
Sassen, R., Roberts, H. H., Carney, R., Milkov, A. V., DeFreitas, D. A., Lanoil, B. & Zhang, C. 2004. Free hydrocarbon gas, gas hydrate, and authigenic minerals in chemosynthetic communities of the northern Gulf of Mexico continental slope: relation to microbial processes. Chemical Geology 205 (3–4), 195217.Google Scholar
Sawlowicz, Z. 2000. Framboids: from their Origin to Application. Poland: Prace Mineralogiczne, 88 pp.Google Scholar
Schink, B. 2002. Synergistic interactions in the microbial world. Antonie Van Leeuwenhoek 81, 257–61.Google Scholar
Schubert, C. J., Vazquez, F., Lösekann-Behrens, T., Knittel, K., Tonolla, M. & Boetius, A. 2011. Evidence for anaerobic oxidation of methane in sediments of a freshwater system (Lago di Cadagno). FEMS Microbiology Ecology 76, 2638.Google Scholar
Shinn, E. A., Steinen, R. P., Lidz, B. H. & Swart, P. K. 1989. Whitings, a sedimentologic dilemma. Journal of Sedimentary Petrology 59, 147–61.Google Scholar
Smol, J. 1988. Chrysophycean microfossils in paleolimnological studies. Palaeogeography, Palaeoclimatology, Palaeoecology 62, 287–97.Google Scholar
Sondi, I. & Juracic, M. 2010. Whiting events and the formation of aragonite in Mediterranean Karstic Marine Lakes: new evidence on its biologically induced inorganic origin. Sedimentology 57, 8595.Google Scholar
Sorrel, P., Oberhänsli, H., Boroffka, N. G. O., Nourgaliev, D., Dulski, P. & Röhl, U. 2007. Control of wind strength and frequency in the Aral Sea basin during the late Holocene. Quaternary Research 67, 371–82.Google Scholar
Spadafora, A., Perri, E., McKenzie, J. A. & Vasconcelos, C. 2010. Microbial biomineralization processes forming modern Ca:Mg carbonate stromatolites. Sedimentology 57, 2740.Google Scholar
Stams, A. J. M. & Plugge, C. M. 2009. Electron transfer in syntrophic communities of anaerobic bacteria and archaea. Nature Reviews Microbiology 7, 568–77.Google Scholar
Suttle, C. A. 2007. Marine viruses – major players in the global ecosystem. Nature Reviews in Microbiology 5, 801–12.Google Scholar
Swierczynski, T., Lauterbach, S., Dulski, P., Delgado, J., Merz, B. & Brauer, A. 2013. Mid- to late Holocene flood frequency changes in the northeastern Alps as recorded in varved sediments of Lake Mondsee (Upper Austria). Quaternary Science Reviews 80, 7890.Google Scholar
Talbot, M. R. & Kelts, K. 1986. Primary and diagenetic carbonates in the anoxic sediments of Lake Bosumtwi, Ghana. Geology 14, 912–6.Google Scholar
Thauer, R. K. & Shima, S. 2008. Methane as fuel for anaerobic organisms. Annals of the NY Academy of Sciences 1125, 158–70.Google Scholar
Turcq, B., Albuquerque, A. L. S., Cordeiro, R. C., Sifeddine, A., Simoes Filho, F. F. L., Souza, A. G., Abrão, J. J., Oliveira, F. B. L., Silva, A. O. & Capitâneo, J. 2002. Accumulation of organic carbon in five Brazilian lakes during the Holocene. Sedimentary Geology 148, 319–42.Google Scholar
Valentine, D. L. & Reeburgh, W. S. 2000. New perspectives on anaerobic methane oxidation. Environmental Microbiology 2, 477–84.Google 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. A., Bernasconi, S., Grujic, D. & Tiens, A. J. 1995. Microbial mediation as a possible mechanism for natural dolomite formation at low temperatures. Nature 377, 220–2.Google Scholar
Vigneron, A., Cruaud, P., Pignet, P., Caprais, J. C., Gayet, N., Cambon-Bonavita, M. A., Godfroy, A. & Toffin, L. 2013. Bacterial communities and syntrophic associations involved in anaerobic oxidation of methane process of the Sonora Margin cold seeps, Guaymas Basin. Environmental Microbiology, published online 13 December 2013. doi: 10.1111/1462-2920.12324.Google Scholar
Vincent, W. F., Whyte, L. G., Lovejoy, C., Greer, C. W., Laurion, I., Suttle, C. A., Corbeil, J. & Mueller, D. R. 2009. Arctic microbial ecosystems and impacts of extreme warming during the International Polar Year. Polar Science 3, 171–80.Google Scholar
Vuillemin, A., Ariztegui, D., De Coninck, A. S., Lücke, A., Mayr, C., Schubert, C. J. & The Pasado Scientific Team. 2013. Origin and significance of diagenetic concretions in sediments of Laguna Potrok Aike, southern Argentina. Journal of Paleolimnology 50, 275–91.Google Scholar
Waldron, S., Hall, A. J. & Fallick, A. E. 1999. Enigmatic stable isotope dynamics of deep peat methane. Global Biogeochemical Cycles 13 (1), 93100.Google Scholar
Warren, J. 2000. Dolomite: occurrence, evolution and economically important associations. Earth-Science Reviews 52, 181.Google Scholar
Warthmann, R., Lith, Y. V., Vasconcelos, C., McKenzie, J. A. & Karpoff, A. M. 2000. Bacterially induced dolomite precipitation in anoxic culture experiments. Geology 28, 1091–4.Google Scholar
Wilhelm, S. W. & Suttle, C. A. 1999. Viruses and nutrient cycles in the sea. Bioscience 49, 781–8.Google Scholar
Wilkin, R. T. & Barnes, H. L. 1997. Formation processes of framboidal pyrite. Geochimica et Cosmochimica Acta 61, 323–39.Google Scholar
Wilken, L. R., Kristiansen, J. & Jürgensen, T. 1995. Silica-scaled chrysophytes from the peninsula of Nuusuaq/Nûugssuaq. Nova Hedwigia 61, 355–66.Google Scholar
Wilkinson, A. N., Hall, R. I. & Smol, J. P. 1999. Chrysophyte cysts as paleolimnological indicators of environmental change due to cottage development and acidic deposition in the Muskoka-Haliburton region, Ontario, Canada. Journal of Paleolimnology 22, 1739.Google Scholar
Wright, D. 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. & Oren, A. 2005. Non-photosynthetic bacteria and the formation of carbonates and evaporites through time. Geomicrobiology Journal 22, 2753.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
Yau, S., Lauro, F. M., DeMaere, M. Z., Brown, M. V., Thomas, T., Raftery, M. J., Andrews-Pfannkoch, C., Lewis, M., Hoffman, J. M., Gibson, J. A. & Cavicchioli, R. 2011. Virophage control of antarctic algal host–virus dynamics. PNAS 108, 6163–8.Google Scholar
Zeeb, B. A., Christie, C. E., Smol, J. P., Findlay, D., Kling, H. & Birks, H. J. B. 1994. Responses of diatom and chrysophyte assemblages in Lake 227 to experimental eutrophication. Canadian Journal of Fisheries and Aquatic Sciences 51, 2300–11.Google Scholar
Zeeb, B. A., Duff, K. E. & Smol, J. P. 1990. Morphological descriptions and stratigraphic profiles of chrysophycean stomatocysts from the recent sediments of Little Round Lake, Ontario. Nova Hedwigia 51, 361–80.Google Scholar
Zeeb, B. A. & Smol, J. P. 1993. Postglacial chrysophycean cyst record from Elk Lake, Minnesota. In Elk Lake, Minnesota: Evidence for Rapid Climate Change in the North-Central United States, Geological Society of America Special Paper 276 (eds Bradbury, J. P. & Dean, W. E.), pp. 239–49. Boulder, Colorado: Geological Society of America.Google Scholar
Zeeb, B. A. & Smol, J. P. 1995. A weighted-averaging regression and calibration model for inferring lakewater salinity using chrysophycean cysts from lakes in western Canada. International Journal of Salt Lake Research 4, 123.Google Scholar
Supplementary material: File

Pacton supplementary material

Pacton supplementary material 1

Download Pacton supplementary material(File)
File 16.2 KB
Supplementary material: Image

Pacton supplementary material

Figure S1

Download Pacton supplementary material(Image)
Image 452.1 KB
Supplementary material: Image

Pacton supplementary material

Figure S2

Download Pacton supplementary material(Image)
Image 449.2 KB
Supplementary material: Image

Pacton supplementary material

Figure S3

Download Pacton supplementary material(Image)
Image 395.7 KB
Supplementary material: Image

Pacton supplementary material

Figure S4

Download Pacton supplementary material(Image)
Image 102.6 KB