Hostname: page-component-848d4c4894-5nwft Total loading time: 0 Render date: 2024-06-08T21:23:22.358Z Has data issue: false hasContentIssue false

The Exceptional Preservation of Interesting and Informative Biomolecules

Published online by Cambridge University Press:  21 July 2017

Roger E. Summons*
Affiliation:
Department of Earth, Atmospheric and Planetary Sciences, Massachusetts Institute of Technology, 77 Massachusetts Avenue E25-633, Cambridge, MA 02139 USA
Get access

Abstract

Organic carbon compounds record key aspects of the processes of their formation and mechanisms of preservation and are foci for research into the nature of life on the early Earth and the search for life beyond Earth. Prime in this respect are lipids preserved in sediments and sedimentary rocks that reveal much about the evolutionary trajectory of life on our planet. Lipids, predominantly derived from photoautotrophic microbes, dominate hydrocarbon records in the Precambrian. The rise to prominence of the Metazoa in the late Neoproterozoic, metaphytes in the Paleozoic, and modern plankton in the Mesozoic, can also be seen in the occurrences of distinctive molecular fossils. Organic matter of all types is optimally preserved in environments and sediments where radiation (solar and ionizing) and oxygen are excluded. In the marine realm, anoxic water bodies will often become sulfidic (euxinic) due to the activity of sulfate-reducing bacteria. Phototropic sulfur bacteria thrive in such environments, and the presence of their characteristic carotenoid pigments goes hand-in-hand with the enhanced preservation of all organic matter, driven by the reducing power of sulfide. The deleterious effects of radiation and oxygen on the preservation of organic matter are amply demonstrated by the results of ongoing searches for carbon compounds on the surface of Mars. The production of highly oxidizing substances through radical chemistry operating in the Martian atmosphere has resulted in environmental conditions that virtually assure destruction of much of the organic matter produced in situ or carried there on meteorites and interplanetary dust particles.

Type
Research Article
Copyright
Copyright © 2014 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

Adam, P., Schmid, J. C., Mycke, B., Strazielle, C., Connan, J., Huc, A., Riva, A., and Albrecht, P. 1993. Structural investigation of nonpolar sulfur cross-linked macromolecules in petroleum. Geochimica et Cosmochimica Acta, 57:33953419.Google Scholar
Aiello, A., Fattorusso, E., and Menna, M.M. 1999. Steroids from sponges: recent reports. Steroids, 64:687714.CrossRefGoogle ScholarPubMed
Antcliffe, J. B. 2013. Questioning the evidence of organic compounds called sponge biomarkers. Palaeontology, 56:917925.Google Scholar
Antcliffe, J. B., Callow, R. H. T., and Brasier, M. D. 2014. Giving the early fossil record of sponges a squeeze. Biological Reviews of the Cambridge Philosophical Society epublication, April 2014. doi: 10.1111/brv.12090 Google Scholar
Benner, S. A., Devine, K. G., Matveeva, L. N., and Powell, D. H. 2000. The missing organic molecules on Mars. Proceedings of the National Academy of Sciences of the United States of America, 97:24252430.Google Scholar
Bergquist, P. R., Karuso, P., Cambie, R. C., and Smith, D. J. 1991. Sterol composition and classification of the Porifera. Biochemical Systematics and Ecology, 19:1724.Google Scholar
Biddle, J., Lipp, J. S., Lever, M. A., Lloyd, K. G., Sorensen, K. B., Anderson, R., Fredricks, H., Elvert, M., Kelly, T. J., Schrag, D. P., Sogin, M. L., Brenchley, J. E., Teske, A., House, C. H., and Hinrichs, K.-U. 2006. Heterotrophic Archaea dominate sedimentary subsurface ecosystems off Peru. Proceedings of the National Academy of Sciences of the United States of America, 103:38463851.Google Scholar
Biemann, K., Oro, J., Toulmin, P., Orgel, L. E., Nier, A. O., Anderson, D. M., Simmonds, P. G., Flory, D., Diaz, A. V., Rushneck, D. R., Biller, J. E., and Lafleur, A. L. 1977. The search for organic substances and inorganic volatile compounds in the surface of Mars. Journal of Geophysical Research, 82:46414658.Google Scholar
Bird, C.W., Lynch, J. M., Pirt, F. J., Reid, W. W., Brooks, C. J. W., and Middleditch, B. S. 1971. Steroids and squalene in Methylococcus capsulatus grown on methane. Nature, 230:473474.Google Scholar
Birgel, D., Thiel, V., Hinrichs, K.-U., Elvert, M., Campbell, K. A., Reitner, J., Farmer, J. D., and Peckmann, J. 2006. Lipid biomarker patterns of methane-seep microbialites from the Mesozoic convergent margin of California. Organic Geochemistry, 37:12891302.Google Scholar
Bloch, K. 1979. Speculations on the evolution of sterol structure and function. CRC Critical Reviews of Biochemistry, 7:15.Google Scholar
Bloch, K. 1992. Sterol molecule: structure, biosynthesis, and function. Steroids, 57:378383.Google Scholar
Bohlmann, J., Meyer-Gauen, G., and Croteau, R. 1998. Plant terpenoid synthases: Molecular biology and phylogenetic analysis. Proceedings of the National Academy of Sciences of the United States of America, 95:41264133.CrossRefGoogle ScholarPubMed
Bradley, A. S., Fredricks, H., Hinrichs, K.-U., and Summons, R. E. 2009. Structural diversity of diether lipids in carbonate chimneys at the Lost City Hydrothermal Field. Organic Geochemistry, 40:11691178.Google Scholar
Brassell, S. C., Eglinton, G., and Maxwell, J. R. 1983. The geochemistry of terpenoids and steroids. Biochemical Society Transactions, 11:575586.Google Scholar
Brassell, S. C., McEvoy, J., Hoffmann, C. F., Lamb, N. A., Peakman, T. M., and Maxwell, J. R. 1984. Isomerisation, rearrangement and aromatisation of steroids in distinguishing early stages of diagenesis. Organic Geochemistry, 6:1123.Google Scholar
Briggs, D. E. G. 1999. Molecular taphonomy of animal and plant cuticles: selective preservation and diagenesis. Philosophical Transactions of the Royal Society of London Series B-Biological Sciences, 354:717.Google Scholar
Briggs, D. E. G., and Summons, R. E. 2014. Ancient biomolecules: their origin, fossilization and significance in revealing the history of life. BioEssays, 36:482490.CrossRefGoogle Scholar
Brocks, J. J., Love, G. D., Summons, R. E., Knoll, A. H., Logan, G. A., and Bowden, S. A. 2005. Biomarker evidence for green and purple sulphur bacteria in a stratified Palaeoproterozoic sea. Nature, 437:866870.Google Scholar
Brocks, J. J., and Schaeffer, P. 2008. Okenane, a biomarker for purple sulfur bacteria (Chromatiaceae), and other new carotenoid derivatives from the 1640 Ma Barney Creek Formation. Geochimica et Cosmochimica Acta, 72:13961414.Google Scholar
Brocks, J. J., and Summons, R. E. 2003. Sedimentary hydrocarbons, biomarkers for early life, p. 64115. In Schlesinger, W. H. (ed.), Biogeochemistry. Elsevier.Google Scholar
Burhan, R. Y. P., Trendel, J. M., Adam, P., Wehrung, P., Albrecht, P., and Nissenbaum, A. 2002. Fossil bacterial ecosystem at methane seeps: Origin of organic matter from Be'eri sulfur deposit, Israel. Geochimica et Cosmochimica Acta, 66:40854101.Google Scholar
Chappe, B., Michaelis, W., and Albrecht, P. 1980. Molecular fossils of Archaebacteria as selective degradation products of kerogen. Physics and Chemistry of the Earth, 12:265274.Google Scholar
Cody, G. D., Gupta, N. S., Briggs, D. E. G., Kilcoyne, A. L. D., Summons, R. E., Kenig, F., Plotnick, R. E., and Scott, A. C. 2011. Molecular signature of chitin-protein complex in Paleozoic arthropods. Geology, 39:255258.CrossRefGoogle Scholar
Coolen, M. J. L., Boere, A., Abbas, B., Baas, M., Wakeham, S. G., and Sinninghe Damsté, J. S. 2006. Ancient DNA derived from alkenone-biosynthesizing haptophytes and other algae in Holocene sediments from the Black Sea. Paleoceanography, 21:PA1005. doi: 10.1029/2005PA001188 Google Scholar
Coolen, M. J. L., Orsi, W. D., Balkema, C., Quince, C., Harris, K., Sylva, S. P., Filipova-Marinova, M., and Giosan, L. 2013. Evolution of the plankton paleome in the Black Sea from the Deglacial to Anthropocene. Proceedings of the National Academy of Sciences of the United Sates of America, 110:86098614.Google Scholar
de la Torre, J. R., Walker, C. B., Ingalls, A. E., Könneke, M.M., and Stahl, D. A. 2008. Cultivation of a thermophilic ammonia oxidizing archaeon synthesizing crenarchaeol. Environmental Microbiology, 10:810818.Google Scholar
De Rosa, M., and Gambacorta, A. 1988. The lipids of archaebacteria. Progress in Lipid Research, 27:153175.CrossRefGoogle ScholarPubMed
Doughty, D. M., Coleman, M. L., Hunter, R. C., Sessions, A. L., Summons, R. E., and Newman, D. K. 2011. The RND-family transporter, HpnN, is required for hopanoid localization to the outer membrane of Rhodopseudomonas palustris TIE-1. Proceedings of the National Academy of Sciences of the United Sates of America, 108:E1045E1051.Google Scholar
Doughty, D. M., Hunter, R. C., Summons, R. E., and Newman, D. K. 2009. 2-Methylhopanoids are maximally produced in akinetes of Nostoc punctiforme: geobiological implications. Geobiology 7, 524532.CrossRefGoogle ScholarPubMed
Dutta, S., Tripathi, S. M., Mallick, M., Mathews, R. P., Greenwood, P. F., Rao, M. R., and Summons, R. E. 2011. Eocene out-of-India dispersal of Asian dipterocarps. Review of Paleobotany and Palynology, 166:6368.CrossRefGoogle Scholar
Ensminger, A., Van Dorsselaer, A., Spyckerelle, C., Albrecht, P., and Ourisson, G. 1974. Pentacyclic triterpanes of the hopane type as ubiquitous geochemical markers: origin and significance, p. 245260. In Tissot, B. and Bienner, F. (eds.), Advances in Organic Geochemistry 1973. Editions Technip, Paris.Google Scholar
Falkowski, P. G., Katz, M. E., Knoll, A. H., Quigg, A., Raven, J. A., Schofield, O., and Taylor, F. J. R. 2004. The evolution of modern eukaryotic phytoplankton. Science, 305:354360.Google Scholar
Falkowski, P., and Knoll, A. H., Eds. 2007. The Evolution of Primary Producers in the Sea. Elsevier, Boston.Google Scholar
Field, C. B., Behrenfeld, M. J., Randerson, J. T., and Falkowski, P. 1998. Primary production of the biosphere: Integrating terrestrial and oceanic components. Science, 281:237240.Google Scholar
Förster, H., Biemann, K., Haigh, W. G., Tattrie, N. H., and Colvin, J. R. 1973. The structure of novel C35 pentacyclic terpenes from Acetobacter xylinum . Biochemical Journal, 135:133143.Google Scholar
Freeman, K. H., Hayes, J. M., Trendel, J.-M., and Albrecht, P. 1990. Evidence from carbon isotope measurements for diverse origins of sedimentary hydrocarbons. Nature, 343:254256.CrossRefGoogle ScholarPubMed
French, K. L., Sepúlveda, J., Trabucho-Alexandre, J., Gröcke, D. R., and Summons, R. E. 2014. Organic geochemistry of the early Toarcian oceanic anoxic event in Hawsker Bottoms, Yorkshire, England. Earth and Planetary Science Letters, 390:116127.Google Scholar
Gaines, S. M., Eglinton, G., and Rullkötter, J. 2009. Echoes of Life: What Fossil Molecules Reveal About Earth History. Oxford University Press, New York.Google Scholar
Giner, J. L. 1993. Biosynthesis of marine sterol side chains. Chemical Reviews 93, 17351752.CrossRefGoogle Scholar
Giner, J. L., and Djerassi, C. 1991. Biosynthetic studies of marine lipids. 31. Evidence for a protonated cyclopropyl intermediate in the biosynthesis of 24-propylidenecholesterol. Journal of the American Chemical Society, 113:13861393.Google Scholar
Giner, J. L., Zhao, H., Boyer, G. L., Satchwell, M. F., and Andersen, R. A. 2009. Sterol chemotaxonomy of marine pelagophyte algae. Chemistry & Biodiversity, 6:11111130.Google Scholar
Glass, K.S., Ito, S., Wilby, P. R., Sota, T., Nakamura, A., Bowers, C. R., Miller, K. E., Dutta, S., Summons, R. E., Briggs, D. E. G., Wakamatsu, K., and Simon, J. D. 2013. Impact of diagenesis and maturation on the survival of eumelanin in the fossil record. Organic Geochemistry, 64:2937.CrossRefGoogle Scholar
Glass, K., Ito, S., Wilby, P. R., Sota, T., Nakamura, A., Bowers, C. R., Vinther, J., Dutta, S., Summons, R., Briggs, D. E. G., Wakamatsu, K., and Simon, J. D. 2012. Direct chemical evidence for eumelanin pigment from the Jurassic period. Proceedings of the National Academy of Sciences of the United States of America, published online ahead of print May 21, 2012. doi: 10.1073/pnas.1118448109 Google Scholar
Glavin, D. P., Freissinet, C., Miller, K. E., Eigenbrode, J. L., Brunner, A. E., Buch, A., Sutter, B., Archer, P. D., Atreya, S. K., Brinckerhoff, W. B., Cabane, M., Coll, P., Conrad, P. G., Coscia, D., Dworkin, J. P., Franz, H. B., Grotzinger, J. P., Leshin, L. A., Martin, M. G., McKay, C., Ming, D. W., Navarro-González, R., Pavlov, A., Steele, A., Summons, R. E., Szopa, C., Teinturier, S., and Mahaffy, P. R. 2013. Evidence for perchlorates and the origin of chlorinated hydrocarbons detected by SAM at the Rocknest aeolian deposit in Gale Crater. Journal of Geophysical Research: Planets, 118:19551973.Google Scholar
Gliozzi, A., Rolandi, R., De Rosa, M., and Gambacorta, A. 1983. Monolayer black membranes from bipolar lipids of archaebacteria and their temperature-induced structural changes. Journal of Membrane Biology, 75:4556.CrossRefGoogle ScholarPubMed
Goldfine, H., and Langworthy, T. A. 1988. A growing interest in bacterial ether lipids. Trends in Biochemical Sciences, 13:217221.Google Scholar
Grantham, P. J. 1986. The occurrence of unusual C27 and C29 sterane predominances in two types of Oman crude oil. Organic Geochemistry, 9:110.Google Scholar
Grice, K., Cao, C., Love, G., Bottcher, M., Twitchett, R., Grosjean, E., Summons, R., Turgeon, S., Dunning, W., and Jin, Y. 2005. Photic zone euxinia during the Permian–Triassic superanoxic event. Science, 307:706709.Google Scholar
Grosjean, E., Love, G. D., Stalvies, C., Fike, D. A., and Summons, R. E. 2009. New oil-source rock correlations in the South Oman Salt Basin. Organic Geochemistry, 40:87110.Google Scholar
Grotzinger, J. P., Sumner, D. Y., Kah, L. C., Stack, K., Gupta, S., Edgar, L., Rubin, D., Lewis, K., Schieber, J., Mangold, N., Milliken, R., Conrad, P. G., DesMarais, D., Farmer, J., Siebach, K., Calef, F. III, Hurowitz, J., McLennan, S. M., Ming, D., Vaniman, D., Crisp, J., Vasavada, A., Edgett, K. S., Malin, M., Blake, D., Gellert, R., Mahaffy, P., Wiens, R. C., Maurice, S., Grant, J. A., Wilson, S., Anderson, R. C., Beegle, L., Arvidson, R., Hallet, B., Sletten, R. S., Rice, M., Bell, J. III, Griffes, J., Ehlmann, B., Anderson, R. B., Bristow, T. F., Dietrich, W. E., Dromart, G., Eigenbrode, J., Fraeman, A., Hardgrove, C., Herkenhoff, K., Jandura, L., Kocurek, G., Lee, S., Leshin, L. A., Leveille, R., Limonadi, D., Maki, J., McCloskey, S., Meyer, M., Minitti, M., Newsom, H., Oehler, D., Okon, A., Palucis, M., Parker, T., Rowland, S., Schmidt, M., Squyres, S., Steele, A., Stolper, E., Summons, R., Treiman, A., Williams, R., Yingst, A., MSL Science Team. 2014. A habitable fluviolacustrine environment at Yellowknife Bay, Gale Crater, Mars. Science, 343. doi: 10.1126/science.1242777 CrossRefGoogle Scholar
Gupta, N. S., Briggs, D. E. G., Collinson, M. E., Evershed, R. P., Michels, R., Jack, K. S., and Pancost, R. D. 2007a. Evidence for the in situ polymerisation of labile aliphatic organic compounds during the preservation of fossil leaves: Implications for organic matter preservation. Organic Geochemistry, 38:499522.Google Scholar
Gupta, N. S., Cody, G. D., Tetlie, O. E., Briggs, D. E. G., and Summons, R. E. 2009. Rapid incorporation of lipids into macromolecules during experimental decay of invertebrates: Initiation of geopolymer formation. Organic Geochemistry, 40:589594.Google Scholar
Gupta, N. S., Collinson, M. E., Briggs, D. E. G., Evershed, R. P., and Pancost, R. D. 2006. Reinvestigation of the occurrence of cutan in plants: implications for the leaf fossil record. Paleobiology, 32:432449.Google Scholar
Gupta, N. S., Michels, R., Briggs, D. E. G., Collinson, M. E., Evershed, R. P., and Pancost, R. D. 2007b. Experimental evidence for the formation of geomacromolecules from plant leaf lipids. Organic Geochemistry, 38:2836.Google Scholar
Gupta, N. S., and Summons, R. E. 2011. Fate of chitinous organisms in the geosphere, p. 133151. In Gupta, N. S. (ed.). Chiton. Topics in Geobiology 34, Springer, Netherlands.Google Scholar
Handley, L., Talbot, H. M., Cooke, M. P., Anderson, K. E., and Wagner, T. 2010. Bacteriohopanepolyols as tracers for continental and marine organic matter supply and phases of enhanced nitrogen cycling on the late Quaternary Congo deep sea fan. Organic Geochemistry, 41:910914.CrossRefGoogle Scholar
Hartgers, W. A., Sinninghe Damsté, J. S., Requejo, A. G., Allan, J., Hayes, J. M., Ling, Y., Xie, T.-M., Primack, J., and de Leeuw, J. W. 1994. A molecular and carbon isotopic study towards the origin and diagenetic fate of diaromatic carotenoids. Organic Geochemistry, 22:703725.Google Scholar
Hayes, J. M. 2001. Fractionation of carbon and hydrogen isotopes in biosynthetic processes. Reviews in Mineralogy & Geochemistry, 43:225277.Google Scholar
Hayes, J. M., Freeman, K. H., Popp, B. N., and Hoham, C. H. 1990. Compound-specific isotopic analyses: A novel tool for reconstruction of ancient biogeochemical processes. Organic Geochemistry, 16:11151128.Google Scholar
Hebting, Y., Schaeffer, P., Behrens, A., Adam, P., Schmitt, G., Schneckenburger, P., Bernasconi, S. M., and Albrecht, P. 2006. Biomarker evidence for a major preservation pathway of sedimentary organic carbon. Science, 312:16271631.Google Scholar
Hedges, J. I., and Keil, R. G. 1995. Sedimentary organic matter preservation: an assessment and speculative synthesis. Marine Chemistry, 49:81115.Google Scholar
Hernandez-Sanchez, M. T., Homoky, W. B., and Pancost, R. D. 2014. Occurrence of 1-O-monoalkyl glycerol ether lipids in ocean waters and sediments. Organic Geochemistry, 66:113.Google Scholar
Hills, I. R., and Whitehead, E. V. 1966. Triterpanes in optically active petroleum distillates. Nature, 209:977979.Google Scholar
Hinrichs, K.-U., Summons, R. E., Orphan, V., Sylva, S. P., and Hayes, J. M. 2000. Molecular and isotopic analysis of anaerobic methane-oxidizing communities in marine sediments. Organic Geochemistry, 31:16851701.Google Scholar
Hren, M. T., Pagani, M., Erwin, D. M., and Brandon, M. 2010. Biomarker reconstruction of the early Eocene paleotopography and paleoclimate of the northern Sierra Nevada. Geology, 38:710.Google Scholar
Jahnke, L. L., Eder, W., Huber, R., Hope, J. M., Hinrichs, K.-U., Hayes, J. M., Des Marais, D. J., Cady, S. L., and Summons, R. E. 2001. Signature lipids and stable carbon isotope analyses of Octopus Spring hyperthermophilic communities compared with those of aquificales representatives. Applied Environmental Microbiology, 67:51795189.Google Scholar
Jørgensen, B. B., and Boetius, A. 2007. Feast and famine—microbial life in the deep-sea bed. Nature Reviews Microbiology, 5:770781.Google Scholar
Kannenberg, E. L., and Poralla, K. 1999. Hopanoid biosynthesis and function in bacteria. Naturwissenschaften 86, 168176.Google Scholar
Kelly, A. E., Love, G. D., Zumberge, J. E., and Summons, R. E. 2011. Hydrocarbon biomarkers of Neoproterozoic to Lower Cambrian oils from eastern Siberia. Organic Geochemistry, 42:640654.Google Scholar
Kimble, B. J., Maxwell, J. R., Philp, R. P., Eglinton, G., Albrecht, P., Ensminger, A., Arpino, P., and Ourisson, G. 1974. Tri- and tetraterpenoid hydrocarbons in the Messel oil shale. Geochimica et Cosmochimica Acta, 38:11651181.Google Scholar
Knappy, C., Nunn, C. M., Morgan, H., and Keely, B. 2011. The major lipid cores of the archaeon Ignisphaera aggregans: implications for the phylogeny and biosynthesis of glycerol monoalkyl glycerol tetraether isoprenoid lipids. Extremophiles, 15:517528.Google Scholar
Knoll, A. H., Summons, R. E., Waldbauer, J. R., and Zumberge, J. 2007. The geological succession of primary producers in the oceans, p. 133163. In Falkwoski, P. and Knoll, A. H. (eds.), The Evolution of Primary Producers in the Sea. Elsevier, Boston.Google Scholar
Kodner, R. B., Pearson, A., Summons, R. E., and Knoll, A. H. 2008. Sterols in red and green algae: quantification, phylogeny, and relevance for the interpretation of geologic steranes. Geobiology, 6:411420.Google Scholar
Koga, Y. 2014. From promiscuity to the lipid divide: On the evolution of distinct membranes in Archaea and Bacteria. Journal of Molecular Evolution, 78:234242.Google Scholar
Koga, Y., and Morii, H. 2005. Recent advances in structural research on ether lipids from Archaea including comparative and physiological aspects. Bioscience, Biotechnology, and Biochemistry, 69:20192034.Google Scholar
Kohnen, M. E. L., Damsté, J. S. S., ten Haven, H. L., and de Leeuw, J. W. 1989. Early incorporation of polysulphides in sedimentary organic matter. Nature, 341:640.CrossRefGoogle Scholar
Kohnen, M. E. L., Schouten, S., Damsté, J. S. S., de Leeuw, J. W., Merritt, D. A., and Hayes, J. M. 1992. Recognition of paleobiochemicals by a combined molecular sulfur and isotope geochemical approach. Science, 256:358362.Google Scholar
Koopmans, M. P., de Leeuw, J. W., and Damsté, J. S. S. 1997. Novel cyclised and aromatised diagenetic products of β-carotene in the Green River Shale. Organic Geochemistry, 26:451466.Google Scholar
Langworthy, T. A., Holzer, G., Zeikus, J. G., and Tornabene, T. G. 1983. Iso- and anteiso-branched glycerol diethers of the thermophilic anaerobe Thermodesulfobacterium commune. Systematic and Applied Microbiology, 4:117.Google Scholar
Leshin, L. A., Mahaffy, P. R., Webster, C. R., Cabane, M., Coll, P., Conrad, P. G., Archer, P. D. Jr., Atreya, S. K., Brunner, A. E., Buch, A., Eigenbrode, J. L., Flesch, G. J., Franz, H. B., Freissinet, C., Glavin, D. P., McAdam, A. C., Miller, K. E., Ming, D. W., Morris, R. V., Navarro-González, R., Niles, P. B., Owen, T., Pepin, R. O., Squyres, S., Steele, A., Stern, J. C., Summons, R. E., Sumner, D. Y., Sutter, B., Szopa, C., Teinturier, S., Trainer, M. G., Wray, J. J., J. P. Grotzinger, and MSL Science Team. 2013. Volatile, isotope, and organic analysis of Martian fines with the Mars Curiosity Rover. Science, 341. doi: 10.1126/science.1238937 Google Scholar
Lewan, M. D., Spiro, B., Illich, H., Raiswell, R., Mackenzie, A. S., Durand, B., Manning, D. A. C., Comet, P. A., Berner, R. A., and Leeuw, J. W. D. 1985. Evaluation of petroleum generation by hydrous pyrolysis experimentation [and discussion]. Philosophical Transactions of the Royal Society of London Series A-Mathematical and Physical Sciences, 315:123134.Google Scholar
Lincoln, S. A., Bradley, A. S., Newman, S. A., and Summons, R. E. 2013. Archaeal and bacterial glycerol dialkyl glycerol tetraether lipids in chimneys of the Lost City Hydrothermal Field. Organic Geochemistry, 60:4553.Google Scholar
Lincoln, S. A., Wai, B., Eppley, J. M., Church, M. J., Summons, R. E., and DeLong, E. F. 2014. Planktonic Euryarchaeota are a significant source of archaeal tetraether lipids in the ocean. Proceedings of the National Academy of Sciences of the United States of America, 111:98589863.Google Scholar
Lipp, J., Morono, Y., Inagaki, F., and Hinrichs, K. 2008. Significant contribution of Archaea to extant biomass in marine subsurface sediments. Nature, 454:991994.Google Scholar
Lipp, J. S., and Hinrichs, K.-U. 2009. Structural diversity and fate of intact polar lipids in marine sediments. Geochimica et Cosmochimica Acta, 73:68166833.Google Scholar
Littler, K., Robinson, S. A., Bown, P. R., Nederbragt, A. J., and Pancost, R. D. 2011. High sea-surface temperatures during the Early Cretaceous Epoch. Nature Geosciences, 4:169172.Google Scholar
Liu, X.-L., Leider, A., Gillespie, A., Gröger, J., Versteegh, G. J. M., and Hinrichs, K.-U. 2010. Identification of polar lipid precursors of the ubiquitous branched GDGT orphan lipids in a peat bog in Northern Germany. Organic Geochemistry, 41:653660.Google Scholar
Liu, X.-L., Summons, R. E., and Hinrichs, K.-U. 2012. Extending the known range of glycerol ether lipids in the environment: structural assignments based on MS/MS fragmentation patterns. Rapid Communications in Mass Spectrometry, 19:22952302.Google Scholar
Love, G. D., Grosjean, E., Stalvies, C., Fike, D. A., Grotzinger, J. P., Bradley, A. S., Kelly, A. E., Bhatia, M., Meredith, W., Snape, C. E., Bowring, S. A., Condon, D. J., and Summons, R. E. 2009. Fossil steroids record the appearance of Demospongiae during the Cryogenian Period. Nature, 457:718721.Google Scholar
Love, G. D., Snape, C. E., Carr, A. D., and Houghton, R. C. 1995. Release of covalently-bound alkane biomarkers in high yields from kerogen via catalytic hydropyrolysis. Organic Geochemistry, 23:981986.Google Scholar
Lutnaes, B. F., Brandal, Ø., Sjöblom, J., and Krane, J. 2006. Archaeal C80 isoprenoid tetraacids responsible for naphthenate deposition in crude oil processing. Organic and Biomolecular Chemistry, 4:616620.Google Scholar
Mackenzie, A. S., Brassell, S. C., Eglinton, G., and Maxwell, J. R. 1982. Chemical fossils: The geological fate of steroids. Science, 217:491504.Google Scholar
Mackenzie, A. S., Patience, R. L., Maxwell, J. R., Vandenbroucke, M., and Durand, B. 1980. Molecular parameters of maturation in the Toarcian shales, Paris Basin, France—I. Changes in the configurations of acyclic isoprenoid alkanes, steranes and triterpanes. Geochimica et Cosmochimica Acta, 44:17091721.CrossRefGoogle Scholar
Mahaffy, P. R., Webster, C. R., Cabane, M., Conrad, P. G., Coll, P., Atreya, S. K., Arvey, R., Barciniak, M., Benna, M., Bleacher, L, Brinckerhoff, W. B., Eigenbrode, J. L., Carignan, D., Cascia, M., Chalmers, R. A., Dworkin, J. P., Errigo, T., Everson, P., Franz, H., Farley, R., Feng, S., Frazier, G., Freissinet, C., Glavin, D. P., Harpold, D. N., Hawk, D., Holmes, V., Johnson, C. S., Jones, A., Jordan, P., Kellogg, J., Lewis, J., Lyness, E, Malespin, C. A., Martin, D. K., Maurer, J., McAdam, A. C., McLennan, D., Nolan, T. J., Noriega, M., Pavlov, A. A., Prats, B., Raaen, E., Sheinman, O., Sheppard, D., Smith, J., Stern, J. C., Tan, F., Trainer, M., Ming, D. W., Morris, R. V., Jones, J., Gundersen, C., Steele, A., Wray, J., Botta, O., Leshin, L. A., Owen, T., Battel, S., Jakosky, B. M., Manning, H., Squyres, S., Navarro-González, R., McKay, C. P., Raulin, F., Sternberg, R., Buch, A., Sorensen, P., Kline-Schoder, R., Coscia, D., Szopa, C., Teinturier, S., Baffes, C., Feldman, J., Flesch, G., Forouhar, S., Garcia, R., Keymeulen, D., Woodward, S., Block, B. P., Arnett, K., Miller, R., Edmonson, C., Gorevan, S., and Mumm, E. 2012. The sample analysis at Mars investigation and instrument suite. Space Science Reviews, 170:401478.Google Scholar
Mallol, C., Hernández, C. M., Cabanes, D., Sistiaga, A., Machado, J., Rodríguez, A., Pérez, L., and Galván, B. 2013. The black layer of Middle Palaeolithic combustion structures. Interpretation and archaeostratigraphic implications. Journal of Archaeological Science, 40:25152537.Google Scholar
Maresca, J., Graham, J., and Bryant, D. 2008. The biochemical basis for structural diversity in the carotenoids of chlorophototrophic bacteria. Photosynthesis Research, 97:121140.Google Scholar
Martiny, A. C., Tai, A. P. K., Veneziano, D., Primeau, F., and Chisholm, S. W. 2009. Taxonomic resolution, ecotypes and the biogeography of Prochlorococcus . Environmental Microbiology, 11:823832.Google Scholar
McCaffrey, M. A., Moldowan, M. J., Lipton, P. A., Summons, R. E., Peters, K. E., Jeganathan, A., and Watt, D. S. 1994. Paleoenvironmental implications of novel C30 steranes in Precambrian to Cenozoic Age petroleum and bitumen. Geochimica et Cosmochimica Acta, 58:529532.Google Scholar
Michaelis, W., and Albrecht, P. 1979. Molecular fossils of archaebacteria in kerogen. Naturwissenschaften, 66:420422.Google Scholar
Ming, D. W. Archer, P. D. Jr., Glavin, D. P., Eigenbrode, J. L., Franz, H. B., Sutter, B., Brunner, A. E., Stern, J. C., Freissinet, C., McAdam, A. C., Mahaffy, P. R., Cabane, M., Coll, P., Campbell, J. L., Atreya, S. K., Niles, P. B., Bell, J. F. III, Bish, D. L., Brinckerhoff, W. B., Buch, A., Conrad, P. G., Des Marais, D. J., Ehlmann, B. L., Fairén, A. G., Farley, K., Flesch, G. J., Francois, P., Gellert, R., Grant, J. A., Grotzinger, J. P., Gupta, S., Herkenhoff, K. E., Hurowitz, J. A., Leshin, L. A., Lewis, K. W., McLennan, S. M., Miller, K. E., Moersch, J., Morris, R. V., Navarro-González, R., Pavlov, A. A., Perrett, G. M., Pradler, I., Squyres, S. W., Summons, R. E., Steele, A., Stolper, E. M., Sumner, D. Y., Szopa, C., Teinturier, S., Trainer, M. G., Treiman, A. H., Vaniman, D. T., Vasavada, A. R., Webster, C. R., Wray, J. J., Yingst, R. A., MSL Science Team. Volatile and Organic Compositions of Sedimentary Rocks in Yellowknife Bay, Gale Crater, Mars. Science, 343. doi: 10.1126/science.1245267 Google Scholar
Moldowan, J. M. 1984. C30-steranes, novel markers for marine petroleums and sedimentary rocks. Geochimica et Cosmochimica Acta, 48:27672768.Google Scholar
Moldowan, J. M., Dahl, J., Huizinga, B. J., Fago, F. J., Hickey, L. J., Peakman, T. M., and Taylor, D. W. 1994. The molecular fossil record of oleanane and its relation to angiosperms. Science, 265:768771.Google Scholar
Morii, H., Eguchi, T., Nishihara, M., Kakinuma, K., König, H., and Koga, Y. 1998. A novel ether core lipid with H-shaped C80-isoprenoid hydrocarbon chain from the hyperthermophilic methanogen Methanothermus fervidus . Biochimica et Biophysica Acta (BBA)-Lipids and Lipid Metabolism, 1390:339345.Google Scholar
Navarro-Gonzalez, R., Vargas, E., de la Rosa, J., Raga, A. C., and McKay, C. P. 2010. Reanalysis of the Viking results suggests perchlorate and organics at midlatitudes on Mars. Journal Of Geophysical Research-Planets, 115:E12. doi:10.1029/2010JE003599 Google Scholar
Nes, W. R., Joseph, J. M., Landrey, J. R., and Conner, R. L. 1978. The steric requirements for the metabolism of sterols by Tetrahymena pyriformis . Journal of Biological Chemistry 253, 23612367.Google Scholar
Neyens, E., and Baeyens, J. 2003. A review of classic Fenton's peroxidation as an advanced oxidation technique. Journal of Hazardous Materials, 98:3350.Google Scholar
Nichols, P. D., Leeming, R., Rayner, M. S., and Latham, V. 1996. Use of capillary gas chromatography for measuring fecal-derived sterols application to stormwater, the sea-surface microlayer, beach greases, regional studies, and distinguishing algal blooms and human and non-human sources of sewage pollution. Journal of Chromatography A, 733:497509.Google Scholar
Ourisson, G., and Albrecht, P. 1992. Geohopanoids—the most abundant natural-products on Earth. Accounts of Chemical Research, 25:398402.Google Scholar
Ourisson, G., Albrecht, P., and Rohmer, M. 1979. The hopanoids. Palaeochemistry and biochemistry of a group of natural products. Pure and Applied Chemistry, 51:709729.Google Scholar
Ourisson, G., Rohmer, M., and Poralla, K. 1987. Prokaryotic hopanoids and other polyterpenoid sterol surrogates. Annual Review of Microbiology, 41:301333.Google Scholar
Pancost, R. D., Bouloubassi, I., Aloisi, G., Sinninghe Damsté, J. S., and Party, M. S. S. 2001. Three series of non-isoprenoidal dialkyl glycerol diethers in cold-seep carbonate crusts. Organic Geochemistry, 32:695707.Google Scholar
Pancost, R. D., Taylor, K. W. R., Inglis, G. N., Kennedy, E. M., Handley, L., Hollis, C. J., Crouch, E. M., Pross, J., Huber, M., Schouten, S., Pearson, P. N., Morgans, H. E. G., and Raine, J. I. 2013. Early Paleogene evolution of terrestrial climate in the SW Pacific, Southern New Zealand. Geochemistry, Geophysics, Geosystems, 14:54135429.Google Scholar
Partensky, F., Hess, W. R., and Vaulot, D. 1999. Prochlorococcus, a marine photosynthetic prokaryote of global significance. Microbiology and Molecular Biology Reviews, 63:106127.Google Scholar
Pavlov, A. A., Vasilyev, G., Ostryakov, V. M., Pavlov, A. K., and Mahaffy, P. 2012. Degradation of the organic molecules in the shallow subsurface of Mars due to irradiation by cosmic rays. Geophysical Research Letters, 39. doi: 10.1029/2012GL052166 Google Scholar
Pearson, A., and Ingalls, A. E. 2013. Assessing the use of archaeal lipids as marine environmental proxies. Annual Review of Earth and Planetary Sciences, 41:359384.Google Scholar
Peckmann, J., Senowbari-Daryan, B., Birgel, D., and Goedert, J. L. 2007. The crustacean ichnofossil Palaxius associated with callianassid body fossils in an Eocene methane-seep limestone, Humptulips Formation, Olympic Peninsula, Washington. Lethaia, 40:273280.Google Scholar
Peters, K. E., Clark, M. E., Das Gupta, U., McCaffrey, A. M., and Lee, C. Y. 1995. Recognition of an Infracambrian source rock based on biomarkers in the Baghewala-1 oil, India. AAPG Bulletin, 79:14811494.Google Scholar
Peters, K. E., Walters, C. C., and Moldowan, J. M. 2005. The Biomarker Guide (2nd Edition). Cambridge University Press, Cambridge, UK.Google Scholar
Peterse, F., Nicol, G. W., Schouten, S., and Sinninghe Damsté, J. S. 2010. Influence of soil pH on the abundance and distribution of core and intact polar lipid-derived branched GDGTs in soil. Organic Geochemistry, 41:11711175.CrossRefGoogle Scholar
Peterson, K. J., and Butterfield, N. J. 2005. Origin of the Eumetazoa: Testing ecological predictions of molecular clocks against the Proterozoic fossil record. Proceedings of the National Academy of Sciences of the United States of America, 102:95479552.Google Scholar
Peterson, K. J., Cotton, J. A., Gehling, J. G., and Pisani, D. 2008. The Ediacaran emergence of bilaterians: Congruence between the genetic and the geological fossil records. Philosophical Transactions of the Royal Society B, 363:14351443.Google Scholar
Peterson, K. J., McPeek, M. A., and Evans, D. A. D. 2005. Tempo and mode of early animal evolution: inferences from rocks, Hox, and molecular clocks. Paleobiology, 31:3655.Google Scholar
Pichersky, E., and Gang, D. R. 2000. Genetics and biochemistry of secondary metabolites in plants: an evolutionary perspective. Trends in Plant Science, 5:439445.Google Scholar
Rohmer, M. 2010. Hopanoids, p. 133142. In Timmis, K. N. (ed.), Handbook of Hydrocarbon and Lipid Microbiology. Springer, Heidelberg.Google Scholar
Rohmer, M., Bouvier, P., and Ourisson, G. 1979. Molecular evolution of biomembranes—structual equivalents and phylogenetic precursors of sterols. Proceedings of the National Academy of Sciences of the United States of America, 76:847851.Google Scholar
Rohmer, M., Bouvier-Nave, P., and Ourisson, G. 1984. Distribution of hopanoid triterpanes in prokaryotes. Journal of General Microbiology, 130:11371150.Google Scholar
Rohmer, M., and Ourisson, G. 1976. Dérivés du bactériohopane: variations structurales et répartition. Tetrahedron Letters, 17:36373640.Google Scholar
Rostek, F., and Bard, E. 2013. Hydrological changes in eastern Europe during the last 40,000 yr inferred from biomarkers in Black Sea sediments. Quaternary Research, 80:502509.Google Scholar
Rust, J., Singh, H., Rana, R. S., McCann, T., Singh, L., Anderson, K., Sarkar, N., Nascimbene, P. C., Stebner, F., Thomas, J. C., Solórzano Kraemer, M., Williams, C. J., Engel, M. S., Sahni, A., and Grimaldi, D. 2010. Biogeographic and evolutionary implications of a diverse paleobiota in amber from the early Eocene of India. Proceedings of the National Academy of Sciences of the United States of America, 107:1836018365.Google Scholar
Rutters, H., Sass, H., Cypionka, H., and Rullkotter, J. 2001. Mono alky 1 ether phospholipids in the sulfate-reducing bacteria Desulfosarcina variabilis and Desulforhabdus amnigenus . Archives of Microbiology, 176:435442.Google Scholar
Rutters, H., Sass, H., Cypionka, H., and Rullkotter, J. 2002. Phospholipid analysis as a tool to study complex microbial communities in marine sediments. Journal of Microbiological Methods, 48:149160.Google Scholar
Saenz, J. P., Eglinton, T. I., and Summons, R. E. 2011a. Abundance and structural diversity of bacteriohopanepolyols in suspended particulate matter along a river to ocean transect. Organic Geochemistry, 42:774780.Google Scholar
Saenz, J. P. Wakeham, S. G., Eglinton, T. I., and Summons, R. E. 2011b. New constraints on the provenance of hopanoids in the marine geologic record: Bacteriohopanepolyols in marine suboxic and anoxic environments. Organic Geochemistry, 42:13511362.Google Scholar
Schmerk, C. L., Welander, P. V., Hamad, M. A., Bain, K. L., Bernards, M. A., Summons, R. E., and Valvano, M. A. 2014. Elucidation of the Burkholderia cenocepacia hopanoid biosynthesis pathway uncovers functions for conserved proteins in hopanoid-producing bacteria. Environmental Microbiology [epublication ahead of print]. doi: 10.1111/1462-2920.12509.Google Scholar
Schouten, S., Baas, M., Hopmans, E. C., Reysenbach, A. L., and Sinninghe Damsté, J. S. 2008a. Tetraether membrane lipids of Candidatus “Aciduliprofundum boonei”, a cultivated obligate thermoacidophilic euryarchaeote from deep-sea hydrothermal vents. Extremophiles, 12:119124.Google Scholar
Schouten, S., Baas, M., Hopmans, E. C., and Sinninghe Damsté, J. S. 2008b. An unusual isoprenoid tetraether lipid in marine and lacustrine sediments. Organic Geochemistry, 39:10331038.Google Scholar
Schouten, S., Eldrett, J., Greenwood, D. R., Harding, I., Baas, M., and Damsté, J. S. S. 2008c. Onset of long-term cooling of Greenland near the Eocene–Oligocene boundary as revealed by branched tetraether lipids. Geology, 36:147150.Google Scholar
Schouten, S., Hopmans, E. C., Schefuß, E., and Sinninghe Damsté, J. S. 2002. Distributional variations in marine crenarchaeotal membrane lipids: a new tool for reconstructing ancient sea water temperatures? Earth and Planetary Science Letters, 204:265274.Google Scholar
Schouten, S., Hopmans, E. C., Baas, M., Boumann, H., Standfest, S., Konneke, M., Stahl, D. A., and Sinninghe Damsté, J. S. 2008d. Intact membrane lipids of “Candidatus Nitrosopumilus maritimus,” a cultivated representative of the cosmopolitan mesophilic Group I Crenarchaeota. Applied Environmental Microbiology, 74:24332440.Google Scholar
Schouten, S., Hopmans, E. C., Forster, A., van Breugel, Y., Kuypers, M. M. M., and Damsté, J. S. S. 2003. Extremely high sea-surface temperatures at low latitudes during the middle Cretaceous as revealed by archaeal membrane lipids. Geology, 31:10691072.Google Scholar
Schouten, S., Hopmans, E. C., Pancost, R. D., and Damsté, J. S. S. 2000. Widespread occurrence of structurally diverse tetraether membrane lipids: Evidence for the ubiquitous presence of low-temperature relatives of hyperthermophiles. Proceedings of the National Academy of Sciences of the United States of America, 97:1442114426.Google Scholar
Schouten, S., Hopmans, E. C., and Sinninghe Damsté, J. S. 2013. The organic geochemistry of glycerol dialkyl glycerol tetraether lipids: A review. Organic Geochemistry, 54:1961.Google Scholar
Schouten, S., Forster, A., Panoto, F. E., and Sinninghe Damsté, J. S. 2007. Towards calibration of the TEX86 palaeothermometer for tropical sea surface temperatures in ancient greenhouse worlds. Organic Geochemistry, 38:15371546.Google Scholar
Schouten, S., Rijpstra, W. I. C., Kok, M., Hopmans, E. C., Summons, R. E., Volkman, J. K., and Sinninghe Damsté, J. S. 2001. Molecular organic tracers of biogeochemical processes in a saline meromictic lake (Ace Lake). Geochimica et Cosmochimica Acta, 65:16291640.Google Scholar
Schubotz, F., Wakeham, S. G., Lipp, J. S., Fredricks, H. F., and Hinrichs, K.-U. 2009. Detection of microbial biomass by intact polar membrane lipid analysis in the water column and surface sediments of the Black Sea. Environmental Microbiology, 11:27202734.Google Scholar
Seifert, W., and Moldowan, J. 1980. The effect of thermal stress on source-rock quality as measured by hopane stereochemistry. Physics and Chemistry of the Earth, 12:229237.Google Scholar
Seifert, W., and Moldowan, M. 1978. Applications of steranes, terpanes and monoaromatics to the maturation, migration and source of crude oils. Geochimica et Cosmochimica Acta, 42:7795.Google Scholar
Sinninghe Damsté, J. S., and Koopmans, M. P. 1997. The fate of carotenoids in sediments: an overview. Pure and Applied Chemistry, 69:20672074.Google Scholar
Sinninghe Damsté, J. S., Rijpstra, W. I. C., Hopmans, E. C., Jung, M.-Y., Kim, J.-G., Rhee, S.-K., Stieglmeier, M., and Schleper, C. 2012. Intact polar and core glycerol dibiphytanyl glycerol tetraether lipids of Group I.1a and I.1b Thaumarchaeota in soil. Applied and Environmental Microbiology, 78:68666874.Google Scholar
Sinninghe Damsté, J. S., Rijpstra, W. I. C., Hopmans, E. C., Weijers, J. W. H., Foesel, B. U., Overmann, J., and Dedysh, S. N. 2011. 13,16-Dimethyl octacosanedioic acid (iso-diabolic Acid), a common membrane-spanning lipid of acidobacteria Subdivisions 1 and 3. Applied and Environmental Microbiology, 77:41474154.Google Scholar
Sistiaga, A., Mallol, C., Galván, B., and Summons, R. E. 2014. The Neanderthal meal: a new perspective using faecal biomarkers. PLOS ONE, 9(6):e101045. doi:10.1371/journal.pone.0101045 Google Scholar
Smittenberg, R. H., Pancost, R. D., Hopmans, E. C., Paetzel, M., and Sinninghe Damsté, J. S. 2004. A 400-year record of environmental change in an euxinic fjord as revealed by the sedimentary biomarker record. Palaeogeography, Palaeoclimatology, Palaeoecology, 202:331351.Google Scholar
Sperling, E. A., Robinson, J. M., Pisani, D., and Peterson, K. J. 2010. Where's the glass? Biomarkers, molecular clocks, and microRNAs suggest a 200-Myr missing Precambrian fossil record of siliceous sponge spicules. Geobiology, 8:2436.Google Scholar
Stankiewicz, B. A., Briggs, D. E. G., and Evershed, R. O. 1997a. Chemical composition of Paleozoic and Mesozoic fossil invertebrate cuticles as revealed by pyrolysis–gas chromatography/mass spectrometry. Energy & Fuels, 11:515521.Google Scholar
Stankiewicz, B. A., Briggs, D. E. G., Evershed, R. P., Flannery, M. B., and Wuttke, M. 1997b. Preservation of chitin in 25-million-year-old fossils. Science, 276:15411543.Google Scholar
Sturt, H. F., Summons, R. E., Smith, K., Elvert, M., and Hinrichs, K.-U. 2004. Intact polar membrane lipids in prokaryotes and sediments deciphered by high-performance liquid chromatography/electrospray ionization multistage mass spectrometry — new biomarkers for biogeochemistry and microbial ecology. Rapid Communications in Mass Spectrometry, 18:617628.Google Scholar
Summons, R. E., Bradley, A. S., Jahnke, L. L., and Waldbauer, J. R. 2006. Steroids, triterpenoids and molecular oxygen. Philosophical Transactions of the Royal Society of London B-Biological Sciences, 361:951968.Google Scholar
Summons, R. E., and Lincoln, S. A. 2012. Biomarkers: Informative molecules for studies in geobiology, p. 269296. In Knoll, A. H., Canfield, D. E., and Konhauser, K. O. (eds.), Fundamentals of Geobiology, First Edition. Blackwell Publishing Ltd. Google Scholar
Summons, R. E., and Powell, T. G. 1986. Chlorobiaceae in Paleozoic seas revealed by biological markers, isotopes and geology. Nature, 319:763765.Google Scholar
Summons, R. E., and Powell, T. G. 1987. Identification of aryl isoprenoids in source rocks and crude oils: Biological markers for the green sulphur bacteria. Geochimica et Cosmochimica Acta, 51:557566.Google Scholar
Summons, R. E., Powell, T. G., and Boreham, C. J. 1988. Petroleum geology and geochemistry of the Middle Proterozoic McArthur Basin, Northern Australia: III. Composition of extractable hydrocarbons. Geochimica et Cosmochimica Acta, 52:17471763.Google Scholar
Talbot, H. M., Squier, A. H., Keely, B. J., and Farrimond, P. 2003. Atmospheric pressure chemical ionisation reversed-phase liquid chromatography/ion trap mass spectrometry of intact bacteriohopanepolyols. Rapid Communications in Mass Spectrometry, 17:728737.Google Scholar
Talbot, H. M., Summons, R. E., Jahnke, L. L., Cockell, C. S., Rohmer, M., and Farrimond, P. 2008. Cyanobacterial bacteriohopanepolyol signatures from cultures and natural environmental settings. Organic Geochemistry, 39:232263.Google Scholar
Talbot, H. M., Watson, D. F., Murrell, J. C., Carter, J. F., and Farrimond, P. 2001. Analysis of intact bacteriohopanepolyols from methanotrophic bacteria by reversed-phase high-performance liquid chromatography–atmospheric pressure chemical ionisation mass spectrometry. Journal of Chromatography A, 921:175185.Google Scholar
Taylor, D. W., Li, H., Dahl, J., Fago, F. J., Zinniker, D., and Moldowan, J. M. 2006. Biogeochemical evidence for the presence of the angiosperm molecular fossil oleanane in Paleozoic and Mesozoic non-angiospermous fossils. Paleobiology, 32:179190.Google Scholar
Taylor, K. W. R., Huber, M., Hollis, C. J., Hernandez-Sanchez, M. T., and Pancost, R. D. 2013. Re-evaluating modern and Palaeogene GDGT distributions: Implications for SST reconstructions. Global and Planetary Change, 108:158174.Google Scholar
Tierney, J. E., and Russell, J. M. 2009. Distributions of branched GDGTs in a tropical lake system: Implications for lacustrine application of the MBT/CBT paleoproxy. Organic Geochemistry, 40:10321036.Google Scholar
van Helmond, N. A. G. M., Sluijs, A., Reichart, G.-J., Sinninghe Damsté, J. S., Slomp, C. P., and Brinkhuis, H. 2014. A perturbed hydrological cycle during Oceanic Anoxic Event 2. Geology, 42:123126.CrossRefGoogle Scholar
van Meer, G., Voelker, D. R., and Feigenson, G. W. 2008. Membrane lipids: where they are and how they behave. Nature Reviews Molecular Cell Biology, 9:112124.CrossRefGoogle ScholarPubMed
Volkman, J. K. 1986. A review of sterol markers for marine and terrigenous organic matter. Organic Geochemistry, 9:8399.Google Scholar
Volkman, J. K., Barrett, S. M., Blackburn, S. I., Mansour, M. P., Sikes, E. L., and Gelin, F. 1998. Microalgal biomarkers: a review of recent research developments. Organic Geochemistry, 29:11631179.Google Scholar
Volkman, J. K., Burton, H. R., Everitt, D. A., and Allen, D. I. 1988. Pigment and lipid compositions of algal and bacterial communities in Ace Lake, Vestfold Hills, Antarctica. Hydrobiologia, 165:4157.Google Scholar
Wagner, T., Herrle, J. O., Damsté, J. S. S., Schouten, S., Stüsser, I., and Hofmann, P. 2008. Rapid warming and salinity changes of Cretaceous surface waters in the subtropical North Atlantic. Geology, 36:203206.Google Scholar
Wakeham, S. G., Amann, R., Freeman, K. H., Hopmans, E. C., Jorgensen, B. B., Putnam, I. F., Schouten, S., Damsté, J. S. S., Talbot, H. M., and Woebken, D. 2007. Microbial ecology of the stratified water column of the Black Sea as revealed by a comprehensive biomarker study. Organic Geochemistry, 38:20702097.Google Scholar
Wakeham, S. G., Sinninghe Damsté, J. S., Kohnen, M. E. L., and De Leeuw, J. W. 1995. Organic sulfur compounds formed during early diagenesis in Black Sea sediments. Geochimica et Cosmochimica Acta, 59:521533.Google Scholar
Weijers, J. W. H., Panoto, E., van Bleijswijk, J., Schouten, S., Rijpstra, W. I. C., Balk, M., Stams, A. J. M., and Damsté, J. S. S. 2009. Constraints on the biological source(s) of the orphan branched tetraether membrane lipids. Geomicrobiology Journal, 26:402414.Google Scholar
Weijers, J. W. H., Schefuß, E., Kim, J.-H., Sinninghe Damsté, J. S., and Schouten, S. 2014. Constraints on the sources of branched tetraether membrane lipids in distal marine sediments. Organic Geochemistry, 72:1422.Google Scholar
Weijers, J. W. H., Schouten, S., Sluijs, A., Brinkhuis, H., and Sinninghe Damsté, J. S. 2007a. Warm arctic continents during the Palaeocene–Eocene thermal maximum. Earth and Planetary Science Letters, 261:230238.Google Scholar
Weijers, J. W. H., Schouten, S., Spaargaren, O. C., and Sinninghe Damsté, J. S. 2006. Occurrence and distribution of tetraether membrane lipids in soils: Implications for the use of the TEX86 proxy and the BIT index. Organic Geochemistry, 37:16801693.Google Scholar
Weijers, J. W. H., Schouten, S., van den Donker, J. C., Hopmans, E. C., and Sinninghe Damsté, J. S. 2007b. Environmental controls on bacterial tetraether membrane lipid distribution in soils. Geochimica et Cosmochimica Acta, 71:703713.Google Scholar
Welander, P. V., Doughty, D. M., Wu, C. H., Mehay, S., Summons, R. E., and Newman, D. K. 2012. Identification and characterization of Rhodopseudomonas palustris TIE-1 hopanoid biosynthesis mutants. Geobiology, 10:163177.Google Scholar
Welander, P. V., and Summons, R. E. 2012. Discovery, taxonomic distribution, and phenotypic characterization of a gene required for 3-methylhopanoid production. Proceedings of the National Academy of Sciences of the United States of America, 109:1290512910.Google Scholar
Welander, P. V., and Summons, R. E. 2013. Identification of hopanoid, sterol, and tetrahymanol production in the aerobic methanotroph Methylomicrobium alcaliphilum 20Z. American Geophysical Union Abstracts Fall Meeting, 51F:1789.Google Scholar
White, D. C., and Ringelberg, D. B. 1998. Signature lipid biomarker analysis, p. 255272. In Burlage, R. S., Atlas, R., Stahl, D. A., Geesey, G., and Sayler, G. (eds.), Techniques in Microbial Ecology. Oxford University Press, New York.Google Scholar
Zachos, J. C., Schouten, S., Bohaty, S., Quattlebaum, T., Sluijs, A., Brinkhuis, H., Gibbs, S. J., and Bralower, T. J. 2006. Extreme warming of mid-latitude coastal ocean during the Paleocene–Eocene Thermal Maximum: Inferences from TEX86 and isotope data. Geology, 34:737740.Google Scholar
Zhang, X., Gillespie, A. L., and Sessions, A. L. 2009. Large D/H variations in bacterial lipids reflect central metabolic pathways. Proceedings of the National Academy of Sciences of the United States of America, 106:1258012586.Google Scholar
Zundel, M., and Rohmer, M. 1985. Prokaryotic triterpenoids. European Journal of Biochemistry, 150:2327.Google Scholar