Hostname: page-component-7479d7b7d-c9gpj Total loading time: 0 Render date: 2024-07-14T14:16:34.496Z Has data issue: false hasContentIssue false

Microbial Sedimentology of Stromatolites in Neoproterozoic Cap Carbonates

Published online by Cambridge University Press:  21 July 2017

Tanja Bosak
Affiliation:
E25-649, Department of Earth, Atmospheric and Planetary Sciences, MIT, Cambridge, MA 02139 USA
Giulio Mariotti
Affiliation:
54-827, Department of Earth, Atmospheric and Planetary Sciences, MIT, Cambridge, MA 02139 USA
Francis A. MacDonald
Affiliation:
20 Oxford Street, Department of Earth and Planetary Science, Harvard University, Cambridge, MA 02138 USA
J. Taylor Perron
Affiliation:
54-1022, Department of Earth, Atmospheric and Planetary Sciences, MIT, Cambridge, MA 02139 USA
Sara B. Pruss
Affiliation:
The Department of Geosciences, Clark Science Center, Smith College, Northampton, MA 01063 USA
Get access

Abstract

Stromatolite shapes, sizes, and spacings are products of microbial processes and interactions with topography, sedimentation, and flow. Laboratory experiments and studies of modern microbial mats and sediments can help reconstruct processes that shaped some typical stromatolite forms and some atypical microbially influenced sediments from Neoproterozoic cap carbonates. Studies of modern, cohesive microbial mats indicate that microbialaminite facies in the lower Rasthof Formation (Cryogenian) formed in the presence of very low flow and were not deformed by strong waves or currents. Giant wave ripples, corrugated stromatolites, and tube-hosting stromatolites in basal Ediacaran cap carbonates record interactions between microbes, flow, and evolving bedforms. Preferential cementation in and close to the giant ripple crests is attributed to interactions between flow and local topography. These interactions pumped alkaline porewaters into ripple crests and helped nucleate elongated stromatolites. The similar textures of giant wave ripples and elongated, corrugated, and tube-hosting stromatolites suggest growth in the presence of organic-rich, rounded particles and microbial mats, and in flow regimes that permitted mat growth. These hypotheses can be tested by experiments and models that investigate lithification and the macroscopic morphology of microbial mats as a function of the flow regime, preexisting topography, redox-stratification in sediments, and delivery of organic-rich particles. The widespread microbially influenced textures in Cryogenian microbialaminites and basal Ediacaran cap dolostones record a strong reliance of carbonate deposition on the presence of organic nuclei, supporting carbonate accumulation rates comparable to those in modern reefs. Therefore, the unusual macroscopic morphologies of microbially influenced facies in Neoproterozoic cap carbonates may not reflect oceans that were greatly oversaturated with respect to carbonate minerals.

Type
Research Article
Copyright
Copyright © 2013 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

Adachi, N., Ezaki, Y., and Liu, J. 2004. The fabrics and origins of peloids immediately after the end-Permian extinction, Guizhou Province, South China. Sedimentary Geology, 164:161178.Google Scholar
Aitken, J. D. 1991. The Ice Brook Formation and post-Rapitan, Late Proterozoic glaciation, Mackenzie Mountains, Northwest Territories. Geological Survey of Canada Bulletin, 404:143.Google Scholar
Allen, P. A., and Hoffman, P. F. 2005. Extreme winds and waves in the aftermath of a Neoproterozoic glaciation. Nature, 433:123127.CrossRefGoogle ScholarPubMed
Altermann, W. 2008. Accretion, trapping and binding of sediment in Archean stromatolites—morphological expression of the antiquity of life. Space Science Reviews, 135:5579.CrossRefGoogle Scholar
Arp, G., Hofmann, J., and Reitner, J. 1998. Microbial fabric formation in spring mounds (“microbialites”) of alkaline salt lakes in the Badain Jaran sand sea, PR China. PALAIOS, 13:581592.CrossRefGoogle Scholar
Arp, G., Reimer, A., and Reitner, J. 1999. Calcification in cyanobacterial biofilms of alkaline salt lakes. European Journal of Phycology, 34:393403.Google Scholar
Arp, G., Reimer, A., and Reitner, J. 2003. Microbialite formation in seawater of increased alkalinity, Satonda Crater Lake, Indonesia. Journal of Sedimentary Research, 73:105127.CrossRefGoogle Scholar
Arp, G., Thiel, V., Reimer, A., Michaelis, W., and Reitner, J. 1999. Biofilm exopolymers control microbialite formation at thermal springs discharging into the alkaline Pyramid Lake, Nevada, USA. Sedimentary Geology, 126:159176.Google Scholar
Awramik, S. M., and Riding, R. 1988. Role of algal eukaryotes in subtidal columnar stromatolite formation. Proceedings of the National Academy of Sciences of the United States of America. 85:13271329.CrossRefGoogle ScholarPubMed
Bao, H., Fairchild, I. J., Wynn, P. M., and Spötl, C. 2009. Stretching the envelope of past surface environments: Neoproterozoic glacial lakes from Svalbard. Science, 323:119122.CrossRefGoogle ScholarPubMed
Bao, H. M., Lyons, J. R., and Zhou, H. 2008. Triple oxygen isotope evidence for elevated CO2 levels after a Neoproterozoic glaciation. Nature, 453:504506.CrossRefGoogle ScholarPubMed
Battin, T. J., Kaplan, L. A., Newbold, J. D., Cheng, X., and Hansen, C. 2003. Effects of current velocity on the nascent architecture of stream microbial biofilms. Applied and Environmental Microbiology, 69:54435452.CrossRefGoogle ScholarPubMed
Baud, A., Richoz, S., and Marcoux, J. 2005. Calcimicrobial cap rocks from the basal Triassic units: western Taurus occurrences (SW Turkey). Comptes Rendus Palevol, 4:569582.CrossRefGoogle Scholar
Berelson, W., Corsetti, F., Pepe-Ranney, C., Hammond, D., Beaumont, W., and Spear, J. 2011. Hot spring siliceous stromatolites from Yellowstone National Park: assessing growth rate and laminae formation. Geobiology, 9:411424.Google ScholarPubMed
Besemer, K., Singer, G., Limberger, R., Chlup, A.-K., Hochedlinger, G., Hödl, I., Baranyi, C., and Battin, T. J. 2007. Biophysical controls on community succession in stream biofilms. Applied and Environmental Microbiology, 73:49664974.Google Scholar
Black, M. 1933. The algal sediments of Andros Island, Bahamas. Philosophical Transactions of the Royal Society of London Series B, 222:165192.Google Scholar
Bontognali, T. R., Vasconcelos, C., Warthmann, R. J., Bernasconi, S. M., Dupraz, C., Strohmenger, C. J., and McKenzie, J. A. 2010. Dolomite formation within microbial mats in the coastal sabkha of Abu Dhabi (United Arab Emirates). Sedimentology, 57:824844.CrossRefGoogle Scholar
Bosak, T., Bush, J. W. M., Flynn, M. R., Liang, B., Ono, S., Petroff, A. P., and Sim, M. S. 2010. Formation and stability of oxygen-rich bubbles that shape photosynthetic mats. Geobiology, 8:111.CrossRefGoogle ScholarPubMed
Bosak, T., Knoll, A. H., and Petroff, A. P. 2013. The meaning of stromatolites. Annual Review of Earth and Planetary Sciences, 41:2144.Google Scholar
Bosak, T., Lahr, D. J. G., Pruss, S. B., MacDonald, F. A., Dalton, L., and Matys, E. 2011. Agglutinated tests in post-Sturtian cap carbonates of Namibia and Mongolia. Earth and Planetary Science Letters, 308:2940.CrossRefGoogle Scholar
Bosak, T., Liang, B., Sim, M. S., and Petroff, A. P. 2009. Morphological record of oxygenic photosynthesis in conical stromatolites. Proceedings of the National Academy of Sciences, 106:1093910943.Google Scholar
Bosak, T., Liang, B., Wu, T. D., Templer, S., Evans, A., Vali, H., Guerquin-Kern, J. L., Klepac-Ceraj, V., Sim, M., Friedman, Y., and Mui, J. 2012. Cyanobacterial diversity and activity in modern conical microbialites. Geobiology, 10:384401.Google Scholar
Bosak, T., and Newman, D. K. 2005. Microbial kinetic controls on calcite morphology in supersaturated solutions. Journal of Sedimentary Research, 75:190199.Google Scholar
Bosak, T., Souza-Egipsy, V., Corsetti, F. A., and Newman, D. K. 2004. Micrometer-scale porosity as a biosignature in carbonate crusts. Geology, 32:781784.Google Scholar
Bosak, T., Souza-Egipsy, V., and Newman, D. K. 2004. An abiotic model for peloid formation. Geobiology, 2:189198.Google Scholar
Bristow, T. F., Bonifacie, M., Derkowski, A., Eiler, J. M., and Grotzinger, J. P. 2011. A hydrothermal origin for isotopically anomalous cap dolostone cements from south China, Nature, 474:6871.CrossRefGoogle ScholarPubMed
Brock, T. D. 1978. Stromatolites: Yellowstone analogues, p. 337385. In Brock, T. D. (ed.), Thermophilic Organisms and Life at High Temperatures. Springer Series in Microbiology, Springer, New York.CrossRefGoogle Scholar
Browne, K. M., Golubic, S., and Seong-Joo, L. 2000. Shallow marine microbial carbonate deposits, p. 233249. In Riding, R. E. and Awramik, S. M. (eds.), Microbial Sediments. Springer, New York.CrossRefGoogle Scholar
Buick, R. 1992. The antiquity of oxygenic photosynthesis: Evidence from stromatolites in sulphate-deficient Archaean lakes. Science, 255:7477.CrossRefGoogle ScholarPubMed
Buick, R., and Dunlop, J. S. R. 1990. Evaporitic sediments of Early Archaean age from the Warrawoona Group, North Pole, Western Australia. Sedimentology, 37:247277.Google Scholar
Burne, R. V., and Moore, L. S. 1987. Microbialites: organosedimentary deposits of benthic microbial communities. PALAIOS, 2:241254.Google Scholar
Calver, C. R., and Walter, M. R. 2000. The late Neoproterozoic Grassy Group of King Island, Tasmania: correlation and palaeogeographic significance. Precambrian Research, 100:299312.CrossRefGoogle Scholar
Campbell, K. A. 2006. Hydrocarbon seep and hydrothermal vent paleoenvironments and paleontology: Past developments and future research directions. Palaeogeography, Palaeoclimatology, Palaeoecology, 232:362407.CrossRefGoogle Scholar
Chafetz, H. S., and Guidry, S. A. 1999. Bacterial shrubs, crystal shrubs, and ray-crystal shrubs: bacterial vs. abiotic precipitation. Sedimentary Geology, 126:5774.Google Scholar
Cloud, P., and Semikhatov, M. 1969. Proterozoic stromatolite zonation. American Journal of Science, 267:10171061.CrossRefGoogle Scholar
Cloud, P. E. Jr., Wright, L. A., Williams, E. G., Diehl, P., and Walter, M. R. 1974. Giant stromatolites and associated vertical tubes from the Upper Proterozoic Noonday Dolomite, Death Valley region, eastern California. Geological Society of America Bulletin, 85:18691882.Google Scholar
Corsetti, F. A., and Grotzinger, J. P. 2005. Origin and significance of tube structures in Neoproterozoic post-glacial cap carbonates: Example from Noonday Dolomite, Death Valley, United States. PALAIOS, 20:348362.Google Scholar
Défarge, C., Trichet, J., Maurin, A., and Hucher, M. 1994. Kopara in Polynesian atolls: early stages of formation of calcareous stromatolites. Sedimentary Geology, 89:923.Google Scholar
Demicco, R. V., and Hardie, L. A. 1994. Sedimentary Structures and Early Diagenetic Features of Shallow Marine Carbonate Deposits. SEPM Atlas Series 1, SEPM, Tulsa, Oklahoma.Google Scholar
Dupraz, C., Reid, R. P., Braissant, O., Decho, A. W., Norman, R. S., and Visscher, P. T. 2009. Processes of carbonate precipitation in modern microbial mats. Earth-Science Reviews, 96:141162.CrossRefGoogle Scholar
Dupraz, C., and Visscher, P. T. 2005. Microbial lithification in marine stromatolites and hypersaline mats. Trends in Microbiology, 13:429438.Google Scholar
Eckman, J., Andres, M., Marinelli, R., Bowlin, E., Reid, R., Aspden, R., and Paterson, D. 2008. Wave and sediment dynamics along a shallow subtidal sandy beach inhabited by modern stromatolites. Geobiology, 6:2132.CrossRefGoogle Scholar
Fraiser, M. L., and Corsetti, F. A. 2003. Neoproterozoic carbonate shrubs: Interplay of microbial activity and unusual environmental conditions in post-Snowball Earth oceans. PALAIOS, 18:378387.2.0.CO;2>CrossRefGoogle Scholar
Franke, U., Polerecky, L., Precht, E., and Huettel, M. 2006. Wave tank study of particulate organic matter degradation in permeable sediments. Limnology and Oceanography, 51:10841096.Google Scholar
Gebelein, C. D. 1969. Distribution, morphology, and accretion rate of recent subtidal algal stromatolites, Bermuda. Journal of Sedimentary Petrology, 39:4969.Google Scholar
Gehling, J. 2000. Environmental interpretation and a sequence stratigraphic framework for the terminal Proterozoic Ediacara Member within the Rawnsley Quartzite, South Australia. Precambrian Research, 100:6595.Google Scholar
Gerdes, G., Klenke, T., and Noffke, N. 2000. Microbial signatures in peritidal siliciclastic sediments: a catalogue. Sedimentology, 47:279308.CrossRefGoogle Scholar
Giddings, J. A., and Wallace, M. W. 2009. Sedimentology and C-isotope geochemistry of the ‘Sturtian’ cap carbonate, South Australia. Sedimentary Geology, 216:114.CrossRefGoogle Scholar
Ginsburg, R. N., and Lowenstam, H. A. 1958. The influence of marine bottom communities on the depositional environment of sediments. The Journal of Geology, 66:310318.Google Scholar
Godillot, R., Caussade, B., Ameziane, T., and Capblancq, J. 2001. Interplay between turbulence and periphyton in rough open-channel flow. Journal of Hydraulic Research, 39:227239.Google Scholar
Golubic, S. 1982. Stromatolites, fossil and recent: a case history, p. 313326. In Westbroek, P. and De Jong, E. W. (eds.), Biomineralization and Biological Metal Accumulation. D. Reidel Publishing Company, Dordrecht.Google Scholar
Graba, M., Moulin, F. Y., Boulêtreau, S., Garabétian, F., Kettab, A., Eiff, O., Sanchez-Pérez, J. M., and Sauvage, S. 2010. Effect of near-bed turbulence on chronic detachment of epilithic biofilm: Experimental and modeling approaches. Water Resources Research, 46:W11531. doi: 10.1029/2009WR008679.CrossRefGoogle Scholar
Graba, M., Sauvage, S., Moulin, F. Y., Urrea, G., Sabater, S., and Sanchez-Pérez, J. M. 2013. Interaction between local hydrodynamics and algal community in epilithic biofilm. Water Research, 47:21532163.Google Scholar
Gregg, J. M., and Shelton, K. L. 1990. Dolomitization and dolomite neomorphism in the back reef facies of the Bonneterre and Davis formations (Cambrian), southeastern Missouri. Journal of Sedimentary Research, 60:549562.Google Scholar
Grey, K., and Thorne, A. M. 1985. Biostratigraphic significance of stromatolites in upward shallowing sequences of the early Proterozoic Duck Creek Dolomite, Western Australia. Precambrian Research, 29:183206.CrossRefGoogle Scholar
Grotzinger, J. P., and Knoll, A. H. 1999. Stromatolites in Precambrian carbonates: Evolutionary mileposts or environmental dipsticks? Annual Review of Earth and Planetary Sciences, 27:313358.CrossRefGoogle ScholarPubMed
Halverson, G. P., Maloof, A. C., and Hoffman, P. F. 2004. The Marinoan glaciation (Neoproterozoic) in northeast Svalbard. Basin Research, 16:297324.Google Scholar
Harwood, C. L., and Sumner, D. Y. 2011. Microbialites of the Neoproterozoic Beck Spring Dolomite, Southern California. Sedimentology, 58:16481673.Google Scholar
Hegenberger, W. 1987. Gas escape structures in Precambrian peritidal carbonate rocks. Communications of the Geological Survey of South West Africa/Namibia, 3:4955.Google Scholar
Higgins, J. A., Fischer, W. W., and Schrag, D. P. 2009. Oxygenation of the ocean and sediments: Consequences for the seafloor carbonate factory. Earth and Planetary Science Letters, 284:2533.Google Scholar
Higgins, J. A., and Schrag, D. P. 2003. Aftermath of a Snowball Earth. Geochemistry, Geophysics, Geosystems, 4:1028. doi: 10.1029/2002GC000403.Google Scholar
Hoffman, P. 1974. Shallow and deepwater stromatolites in lower Proterozoic platform-to-basin facies change, Great Slave Lake, Canada. American Association of Petroleum Geologists Bulletin, 58:865867.Google Scholar
Hoffman, P. F. 1976. Stromatolite morphogenesis in Shark Bay, Western Australia, p. 261272. In Walter, M. R. (ed.), Stromatolites. Elsevier, Amsterdam.CrossRefGoogle Scholar
Hoffman, P. F. 2011. Strange bedfellows: glacial diamictite and cap carbonate from the Marinoan (635 Ma) glaciation in Namibia. Sedimentology, 58:57119.Google Scholar
Hoffman, P. F., and Halverson, G. P. 2008. Otavi Group of the western Northern Platform, the Eastern Kaoko Zone and the western Northern Margin Zone, p. 13.69–13.136. In Miller, R. M. (ed.), The Geology of Namibia, Volume 2. Handbook of the Geological Survey of Namibia, Windhoek.Google Scholar
Hoffman, P. F., and Halverson, G. P. 2011. Neoproterozoic glacial record in the Mackenzie Mountains, northern Canadian Cordillera, p. 397412. In Arnaud, E., Halverson, G. P., and Shields-Zhou, G. (eds.), The Geological Record of Neoproterozoic Glaciations. The Geological Society of London Memoirs 36, London.CrossRefGoogle Scholar
Hoffman, P. F., Halverson, G. P., Domack, E. W., Husson, J. M., Higgins, J. A., and Schrag, D. P. 2007. Are basal Ediacaran (635 Ma) post-glacial “cap dolostones” diachronous? Earth and Planetary Science Letters, 258:114131.CrossRefGoogle 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 MacDonald, F. A. 2010. Sheet-crack cements and early regression in Marinoan (635 Ma) cap dolostones: Regional benchmarks of vanishing ice-sheets? Earth and Planetary Science Letters, 300:374384.CrossRefGoogle Scholar
Hoffman, P. F., and Schrag, D. P. 2002. The Snowball Earth hypothesis: testing the limits of global change. Terra Nova, 14:129155.CrossRefGoogle Scholar
Hofmann, H. 1975. Stratiform Precambrian stromatolites, Belcher Islands, Canada; relations between silicified microfossils and microstructure. American Journal of Science, 275:11211132.CrossRefGoogle Scholar
Hofmann, H. J. 2000. Archean stromatolites as microbial archives, p. 315327. In Riding, R. E. and Awramik, S. M. (eds.), Microbial Sediments. Springer, Berlin Heidelberg.Google Scholar
Horodyski, R. J. 1975. Stromatolites of the lower Missoula Group (Middle Proterozoic), Belt Supergroup, Glacier National Park, Montana. Precambrian Research, 2:215254.CrossRefGoogle Scholar
Horodyski, R. J., Bloeser, B., and Vonder Haar, S. 1977. Laminated algal mats from a coastal lagoon, Laguna Mormona, Baja California, Mexico. Journal of Sedimentary Research, 47:680696.Google Scholar
Huettel, M., and Gust, G. 1992. Impact of bioroughness on interfacial solute exchange in permeable sediments. Marine Ecology Progress Series, 89:253267.Google Scholar
Huettel, M., Ziebis, W., Forster, S., and Luther, G. 1998. Advective transport affecting metal and nutrient distributions and interfacial fluxes in permeable sediments. Geochimica Et Cosmochimica Acta, 62:613631.Google Scholar
Jahnert, R. J., and Collins, L. B. 2012. Characteristics, distribution and morphogenesis of subtidal microbial systems in Shark Bay, Australia. Marine Geology, 303:115136.CrossRefGoogle Scholar
James, N. P., Narbonne, G. M., and Kyser, T. K. 2001. Late Neoproterozoic cap carbonates: Mackenzie Mountains, northwestern Canada: precipitation and global glacial meltdown. Canadian Journal of Earth Sciences, 38:12291262.Google Scholar
Jerolmack, D. J., and Mohrig, D. 2005. Formation of Precambrian sediment ripples: Arising from P.A. Allen, P.F. Hoffman. Nature, 433:123127.Google Scholar
Jiang, G., Kennedy, M. J., and Christie-Blick, N. 2003. Stable isotopic evidence for methane seeps in Neoproterozoic postglacial cap carbonates. Nature, 426:822826.Google Scholar
Jiang, G., Kennedy, M. J., Christie-Blick, N., Wu, H., and Zhang, S. 2006. Stratigraphy, sedimentary structures, and textures of the late Neoproterozoic Doushantuo cap carbonate in South China. Journal of Sedimentary Research, 76:978995.CrossRefGoogle Scholar
Johnson, D. B., and Hallberg, K. B. 2003. The microbiology of acidic mine waters. Research in Microbiology, 154:466473.CrossRefGoogle ScholarPubMed
Jones, B., and Renaut, R. W. 1996. Morphology and growth of aragonite crystals in hot-spring travertines at Lake Bogoria, Kenya Rift Valley. Sedimentology, 43:323340.Google Scholar
Jones, B., and Renaut, R. W. 2008. Cyclic development of large, complex, calcite dendrite crystals in the Clinton travertine, Interior British Columbia, Canada. Sedimentary Geology, 203:1735.Google Scholar
Kah, L. C., and Knoll, A. H. 1996. Microbenthic distribution of Proterozoic tidal flats: Environmental and taphonomic considerations. Geology, 24:7982.2.3.CO;2>CrossRefGoogle ScholarPubMed
Kasemann, S. A., Hawkesworth, C. J., Prave, A. R., Fallick, A. E., and Pearson, P. N. 2005. Boron and calcium isotope composition in Neoproterozoic carbonate rocks from Namibia: evidence for extreme environmental change. Earth and Planetary Science Letters, 231:7386.Google Scholar
Kasemann, S. A., Prave, A. R., Fallick, A. E., Hawkesworth, C. J., and Hoffmann, K.-H. 2010. Neoproterozoic ice ages, boron isotopes, and ocean acidification: Implications for a snowball Earth. Geology, 38:775778.CrossRefGoogle Scholar
Kaźmierczak, J., Coleman, M. L., Gruszczynski, M., and Kempe, S. 1996. Cyanobacterial key to the genesis of micritic and peloidal limestones in ancient seas. Acta Palaeontologica Polonica, 41:319338.Google Scholar
Kennedy, M. J. 1996. Stratigraphy, sedimentology, and isotope geochemistry of Australian Neoproterozoic postglacial cap dolostones: deglaciation, δ13C excursions, and carbonate precipitation. Journal of Sedimentary Research, 66:10501064.Google Scholar
Kennedy, M. J., and Christie-Blick, N. 2011. Condensation origin for Neoproterozoic cap carbonates during deglaciation. Geology, 39:319322.Google Scholar
Kennedy, M. J., Christie-Blick, N., and Sohl, L. E. 2001. Are Proterozoic cap carbonates and isotopic excursions a record of gas hydrate destabilization following Earth's coldest intervals? Geology, 29:443446.2.0.CO;2>CrossRefGoogle Scholar
Kilner, B., Niocaill, C., and Brasier, M. 2005. Low-latitude glaciation in the Neoproterozoic of Oman. Geology, 33:413416.CrossRefGoogle Scholar
Knoll, A. H., and Semikhatov, M. A. 1998. The genesis and time distribution of two distinctive Proterozoic stromatolite microstructures. PALAIOS, 13:408422.Google Scholar
Labiod, C., Godillot, R., and Caussade, B. 2007. The relationship between stream periphyton dynamics and near-bed turbulence in rough open-channel flow. Ecological Modelling, 209:7896.Google Scholar
Lamb, M. P., Fischer, W. W., Raub, T. D., Perron, J. T., and Myrow, P. M. 2012. Origin of giant wave ripples in snowball Earth cap carbonate. Geology, 40:827830.Google Scholar
Le Ber, E., Le Heron, D., Winterleitner, G., Bosence, D., Vining, B., and Kamona, F. 2013. Microbialite recovery in the aftermath of the Sturtian glaciation: Insights from the Rasthof Formation, Namibia. Sedimentary Geology, 294:112.CrossRefGoogle Scholar
Logan, B. W. 1961. Cryptozoon and associate stromatolites from the Recent, Shark Bay, Western Australia. Journal of Geology, 69:517533.CrossRefGoogle Scholar
Logan, B. W., Rezak, R., and Ginsburg, R. N. 1964. Classification and environmental significance of algal stromatolites. Journal of Geology, 72:6883.Google Scholar
MacDonald, F. A., Jones, D. S., and Schrag, D. P. 2009a. Stratigraphic and tectonic implications of a new glacial diamictite-cap carbonate couplet in southwestern Mongolia. Geology, 37:123126.Google Scholar
MacDonald, F. A., McClelland, W. C., Schrag, D. P., and MacDonald, W. P. 2009b. Neoproterozoic glaciation on a carbonate platform margin in Arctic Alaska and the origin of the North Slope subterrane. Geological Society of America Bulletin 121:448473.Google Scholar
MacDonald, F. A., Strauss, J. V., Rose, C. V., Dudás, F. O., and Schrag, D. P. 2010. Stratigraphy of the Port Nolloth Group of Namibia and South Africa and implications for the age of Neoproterozoic iron formations. American Journal of Science, 310:862888.CrossRefGoogle Scholar
Mariotti, G., and Fagherazzi, S. 2012. Modeling the effect of tides and waves on benthic biofilms. Journal of Geophysical Research: Biogeosciences (2005–2012). 117:G04010.CrossRefGoogle Scholar
Martin, J. M., Braga, J. C., and Riding, R. 1997. Late Miocene Halimeda alga-microbial segment reefs in the marginal Mediterranean Sorbas Basin, Spain. Sedimentology, 44:441456.CrossRefGoogle Scholar
Mata, S. A., Harwood, C. L., Corsetti, F. A., Stork, N. J., Eilers, K., Berelson, W. M., and Spear, J. R. 2012. Influence of gas production and filament orientation on stromatolite microfabric. PALAIOS, 27:206219.CrossRefGoogle Scholar
Milliman, J. D. 1993. Production and accumulation of calcium carbonate in the ocean: Budget of a nonsteady state. Global Biogeochemical Cycles, 7:927957.Google Scholar
Monty, C. L. V. 1976. The origin and development of cryptalgal fabrics, p. 193250. In Walter, M. R. (ed.), Stromatolites. Elsevier, Amsterdam.Google Scholar
Morse, J. W., Arvidson, R. S., and Luttge, A. 2007. Calcium carbonate formation and dissolution. Chemical Reviews-Columbus, 107:342381.CrossRefGoogle ScholarPubMed
Neumann, A. C., Gebelein, C. D., and Scoffin, T. P. 1970. The composition, structure and erodability of subtidal mats, Abaco, Bahamas. Journal of Sedimentary Petrology, 40:274297.Google Scholar
Niederberger, T. D., Perreault, N., Lawrence, J. R., Nadeau, L. J., Mielke, R. E., Greer, C. W., Andersen, D. T., and Whyte, L. G. 2009. Novel sulfur-oxidizing streamers thriving in perennial cold saline springs of the Canadian high Arctic. Environmental Microbiology, 11:616629.Google Scholar
Nikora, V., Goring, D., and Biggs, B. 1997. On stream periphyton-turbulence interactions. New Zealand Journal of Marine and Freshwater Research, 31:435448.Google Scholar
Noffke, N. 2010. Geobiology: Microbial Mats in Sandy Deposits from the Archean Era to Today. Springer, Heidelberg.Google Scholar
Noffke, N., Gerdes, G., Klenke, T., and Krumbein, W. E. 2001. Microbially induced sedimentary structures—A new category within the classification of primary sedimentary structures: perspectives. Journal of Sedimentary Research, 71:649656.Google Scholar
Okumura, T., Takashima, C., Shiraishi, F., Akmaluddin, , and Kano, A. 2012. Textural transition in an aragonite travertine formed under various flow conditions at Pancuran Pitu, Central Java, Indonesia. Sedimentary Geology, 265–266:195209.CrossRefGoogle Scholar
Paterson, D. M. 1997. Biological mediation of sediment erodibility: ecology and physical dynamics, p. 215229. In Burt, N., Parker, R. and Watts, J. (eds.), Cohesive Sediments. John Wiley and Sons, Hoboken, New Jersey.Google Scholar
Payne, J. L., Turchyn, A. V., Paytan, A., DePaolo, D. J., Lehrmann, D. J., Yu, M., and Wei, J. 2010. Calcium isotope constraints on the end-Permian mass extinction. Proceedings of the National Academy of Sciences, 107:85438548.CrossRefGoogle ScholarPubMed
Pentecost, A. 1990. The formation of travertine shrubs: Mammoth Hot Springs, Wyoming. Geological Magazine, 127:159168.CrossRefGoogle Scholar
Perri, E., Tucker, M. E., and Spadafora, A. 2012. Carbonate organo-mineral micro- and ultrastructures in sub-fossil stromatolites: Marion Lake, South Australia. Geobiology, 10:105117.CrossRefGoogle ScholarPubMed
Petroff, A. P., Rothman, D. H., Beukes, N. J., and Bosak, T. in press. Biofilm growth and fossil form. Physical Review. Google Scholar
Petroff, A. P., Sim, M. S., Maslov, A., Krupenin, M., Róthman, D. H., and Bosak, T. 2010. Biophysical basis for the geometry of conical stromatolites. Proceedings of the National Academy of Sciences, 107:99569961.Google Scholar
Petryshyn, V. A., and Corsetti, F. A. 2011. Analysis of growth directions of columnar stromatolites from Walker Lake, western Nevada. Geobiology, 9:425435.Google Scholar
Playford, P. E. 1980. Environmental controls on the morphology of modern stromatolites at Hamelin Pool, Western Australia. Western Australia Geological Survey Annual Report for 1979. Benbow Publishing Co., Perth, Western Australia, p. 7377.Google Scholar
Playford, P. E., and Cockbain, A. E. 1976. Modern algal stromatolites at Hamelin Pool, a hypersaline barred basin in Shark Bay, Western Australia, p. 389411. In Walter, M. R. (ed.), Stromatolites. Elsevier, Amsterdam.CrossRefGoogle Scholar
Precht, E., Franke, U., Polerecky, L., and Huettel, M. 2004. Oxygen dynamics in permeable sediments with wave-driven pore water exchange. Limnology and Oceanography, 49:693705.CrossRefGoogle Scholar
Precht, E., and Huettel, M. 2003. Advective porewater exchange driven by surface gravity waves and its ecological implications. Limnology and Oceanography, 48:16741684.Google Scholar
Pruss, S. B., Bosak, T.T., MacDonald, F. A., McLane, M., and Hoffman, P. F. 2010. Microbial facies in a Sturtian cap carbonate, the Rasthof Formation, Otavi Group, northern Namibia. Precambrian Research, 181:187198.Google Scholar
Raaben, M. 2006. Dimensional parameters of columnar stromatolites as a result of stromatolite ecosystem evolution. Stratigraphy and Geological Correlation. 14:150163.CrossRefGoogle Scholar
Radu, A., Vrouwenvelder, J., van Loosdrecht, M., and Picioreanu, C. 2012. Effect of flow velocity, substrate concentration and hydraulic cleaning on biofouling of reverse osmosis feed channels. Chemical Engineering Journal, 188:3039.CrossRefGoogle Scholar
Reid, R. P. 1987. Nonskeletal peloidal precipitates in upper Triassic reefs, Yukon Territory (Canada). Journal of Sedimentary Research, 57:893900.Google Scholar
Reid, R. P., James, N. P., MacIntyre, I. G., Dupraz, C. P., and Burne, R. V. 2003. Shark Bay stromatolites: Microfabrics and reinterpretation of origins. Facies, 49:299324.CrossRefGoogle Scholar
Reid, R. P., Visscher, P. T., Decho, A. W., Stolz, J. F., Bebout, B. M., Dupraz, C., MacIntyre, L. G., Paerl, H. W., Pinckney, J. L., Prufert-Bebout, L., Steppe, T. F., and DesMarais, D. J. 2000. The role of microbes in accretion, lamination and early lithification of modern marine stromatolites. Nature, 406:989992.CrossRefGoogle ScholarPubMed
Riding, R. 2000. Microbial carbonates: the geological record of calcified bacterial-algal mats and biofilms. Sedimentology, 47:179214.Google Scholar
Riding, R., and Tomás, S. 2006. Stromatolite reef crusts, Early Cretaceous, Spain: bacterial origin of in situ-precipitated peloid microspar? Sedimentology, 53:2334.CrossRefGoogle Scholar
Rooney, A. D., MacDonald, F. A., Dudás, F. O., Hallmann, C., Strauss, J. V., and Selby, D. in review. Weathering the Snowball.Google Scholar
Sami, T. T., and James, N. P. 1994. Peritidal carbonate platform growth and cyclicity in an early Proterozoic foreland basin, upper Pethei Group, northwest Canada. Journal of Sedimentary Research, 64:111131.Google Scholar
Santos, I. R., Eyre, B. D., and Huettel, M. 2012. The driving forces of porewater and groundwater flow in permeable coastal sediments: A review. Estuarine, Coastal and Shelf Science. 98:115.CrossRefGoogle Scholar
Schieber, J., Bose, P. K., Eriksson, P. G., Banerjee, S., Sarkar, S., Altermann, W., and Catuneanu, O. 2007. Atlas of Microbial Mat Features Preserved Within The Siliciclastic Rock Record. Atlases in Geoscience Volume 2, Elsevier, Amsterdam.Google Scholar
Schmidt, P. W., Williams, G. E., and McWilliams, M. O. 2009. Palaeomagnetism and magnetic anisotropy of late Neoproterozoic strata, South Australia: Implications for the palaeolatitude of late Cryogenian glaciation, cap carbonate and the Ediacaran System. Precambrian Research, 174:3552.Google Scholar
Schopf, J. W. 2006. Fossil evidence of Archaean life. Philosophical Transactions of the Royal Society of London, Series B, 361:869885.Google Scholar
Schröder, S., Beukes, N. J., and Sumner, D. Y. 2009. Microbialite-sediment interactions on the slope of the Campbellrand carbonate platform (Neoarchean, South Africa). Precambrian Research, 169:6879.CrossRefGoogle Scholar
Scoffin, T. P. 1970. The trapping and binding of subtidal carbonate sediments by marine vegetation in Bimini Lagoon, Bahamas. Journal of Sedimentary Research, 40:249273.Google Scholar
Semikhatov, M., Gebelein, C., Cloud, P., Awramik, S., and Benmore, W. 1979. Stromatolite morphogenesis-progress and problems. Canadian Journal of Earth Sciences, 16:9921015.CrossRefGoogle Scholar
Semikhatov, M., and Raaben, M. 1996. Dynamics of the global diversity of Proterozoic stromatolites. Article II: Africa, Australia, North America, and general synthesis. Stratigraphy and Geological Correlation, 4:2450.Google Scholar
Seong-Joo, L., Browne, K. M., and Golubic, S. 2000. On stromatolite lamination, p. 1624. In Riding, R. E. and Awramik, S. M. (eds.), Microbial Sediments. Springer, Heidelberg.Google Scholar
Seong-Joo, L., and Golubic, S. 2000. Biological and mineral components of an ancient stromatolite: Gaoyuzhuang Formation, Mesoproterozoic of China, p. 91102. In Grotzinger, J. P. and James, N. P. (eds.), Carbonate Sedimentation and Diagenesis in the Evolving Precambrian World. SEPM Special Publication Volume 67, SEPM, Tulsa, Oklahoma.CrossRefGoogle Scholar
Shields, G. A. 2005. Neoproterozoic cap carbonates: a critical appraisal of existing models and the plumeworld hypothesis. Terra Nova, 17:299310.CrossRefGoogle Scholar
Sim, M. S., Liang, B., Petroff, A. P., Evans, A., Klepac-Ceraj, V., Flannery, D. T., Walter, M. R., and Bosak, T. 2012. Oxygen-dependent morphogenesis of modern clumped photosynthetic mats and implications for the Archean stromatolite record. Geosciences, 2:235259.Google Scholar
Simonson, B. M., and Carney, K. E. 1999. Roll-up structures; evidence of in situ microbial mats in late Archean deep shelf environments. PALAIOS, 14:1324.CrossRefGoogle Scholar
Spadafora, A., Perri, E., McKenzie, J. A., and Vasconcelos, C. 2010. Microbial biomineralization processes forming modern Ca:Mg carbonate stromatolites. Sedimentology, 57:2740.CrossRefGoogle Scholar
Sprachta, S., Camoin, G., Golubic, S., and Le Campion, T. 2001. Microbialites in a modern lagoonal environment: nature and distribution, Tikehau atoll (French Polynesia). Palaeogeography, Palaeoclimatology, Palaeoecology, 175:103124.Google Scholar
Sun, S., and Wright, V. 1989. Peloidal fabrics in Upper Jurassic reefal limestones, Weald Basin, southern England. Sedimentary Geology, 65:165181.Google Scholar
Tice, M. M., Thornton, D. C., Pope, M. C., Olszewski, T. D., and Gong, J. 2011. Archean microbial mat communities. Annual Review of Earth and Planetary Sciences, 39:297319.Google Scholar
Tribovillard, N., Trentesaux, A., Trichet, J., and Défarge, C. 2000. A Jurassic counterpart for modern kopara of the Pacific atolls: lagoonal, organic matter-rich, laminated carbonate of Orbagnoux (Jura Mountains, France). Palaeogeography, Palaeoclimatology, Palaeoecology, 156:277288.Google Scholar
Trichet, J., and Défarge, C. 1995. Non-biologically supported organomineralization. Bulletin of the Oceanographic Institute, Monaco. 14:203236.Google Scholar
Trichet, J., Défarge, C., Tribble, J., Tribble, G., and Sansone, F. 2001. Christmas Island lagoonal lakes, models for the deposition of carbonateevaporite-organic laminated sediments. Sedimentary Geology, 140:177189.CrossRefGoogle Scholar
Trindade, R., Font, E., D'Agrella-Filho, M., Nogueira, A., and Riccomini, C. 2003. Low-latitude and multiple geomagnetic reversals in the Neoproterozoic Puga cap carbonate, Amazon craton. Terra Nova, 15:441446.Google Scholar
Turner, E. C., James, N. P., and Narbonne, G. M. 2000. Taphonomic control on microstructure in Early Neoproterozoic reefal stromatolites and thrombolites. PALAIOS, 15:87111.2.0.CO;2>CrossRefGoogle Scholar
Uehlinger, U., Bührer, H., and Reichert, P. 1996. Periphyton dynamics in a floodprone prealpine river: evaluation of significant processes by modelling. Freshwater Biology, 36:249263.CrossRefGoogle Scholar
Vasconcelos, C., Warthmann, R., McKenzie, J. A., Visscher, P. T., Bittermann, A. G., and van Lith, Y. 2006. Lithifying microbial mats in Lagoa Vermelha, Brazil: Modern Precambrian relics? Sedimentary Geology, 185:175183.Google Scholar
Visscher, P. T., Reid, R. P., and Bebout, B. M. 2000. Microscale observations of sulfate reduction: Correlation of microbial activity with lithified micritic laminae in modern marine stromatolites. Geology, 28:919922.Google Scholar
Walter, M., and Heys, G. 1985. Links between the rise of the Metazoa and the decline of stromatolites. Precambrian Research, 29:149174.CrossRefGoogle Scholar
Walter, M. R., Bauld, J., and Brock, T. D. 1976. Microbiology and morphogenesis of columnar stromatolites (Conophyton, Vacerrilla) from hot springs in Yellowstone National Park, p. 273310. In Walter, M. R. (ed.), Stromatolites. Elsevier, Amsterdam.Google Scholar
Wright, L. A., Williams, E. G., and Cloud, P. 1978. Algal and cryptalgal structures and platform environments of the late pre-Phanerozoic Noonday Dolomite, eastern California. Geological Society of America Bulletin, 89:321333.Google Scholar
Xiao, S., Bao, H., Wang, H., Kaufman, A. J., Zhou, C., Li, G., Yuan, X., and Ling, H. 2004. The Neoproterozoic Quruqtagh Group in eastern Chinese Tianshan: evidence for a post-Marinoan glaciation. Precambrian Research, 130:126.Google Scholar