Hostname: page-component-76fb5796d-qxdb6 Total loading time: 0 Render date: 2024-04-28T21:04:23.665Z Has data issue: false hasContentIssue false

GDGT Thermometry: Lipid Tools for Reconstructing Paleotemperatures

Published online by Cambridge University Press:  21 July 2017

Jessica E. Tierney*
Affiliation:
Woods Hole Oceanographic Institution, Woods Hole, MA USA 02543. tierney@whoi.edu
Get access

Abstract

Microbial communities adjust the chemical structure of their cell membranes in response to environmental temperature. This enables the development of lipid-based paleothermometers such as the glycerol dialkyl glycerol tetraether (GDGT) proxies described here. Surface-sediment calibrations establish a strong empirical relationship between the relative distribution of GDGTs and temperature. GDGT proxies can be used in marine, lacustrine, and paleosol sequences as long as the organic material is not thermally mature. Thus far, GDGT proxies have been applied to sediments dating back to the middle Jurassic. Many of the key uncertainties of these proxies are related to our emerging understanding of archaeal (and for the branched GDGTs, bacterial) ecology.

Type
Research Article
Copyright
Copyright © 2012 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

Bechtel, A., Smittenberg, R., Bernasconi, S., and Schubert, C. 2010. Distribution of branched and isoprenoid tetraether lipids in an oligotrophic and a eutrophic Swiss lake: Insights into sources and GDGT-based proxies. Organic Geochemistry, 41:822832.Google Scholar
Blaga, C., Reichart, G., Schouten, S., Lotter, A., Werne, J., Kosten, S., Mazzeo, N.N., Lacerot, G., and Sinninghe Damsté, J. S. 2010. Branched glycerol dialkyl glycerol tetraethers in lake sediments: Can they be used as temperature and pH proxies? Organic Geochemistry, 41:12251234.CrossRefGoogle Scholar
Blaga, C. I., Reichart, G.-J., Heiri, O., and Sinninghe Damsté, J. S. 2009. Tetraether membrane lipid distributions in water-column particulate matter and sediments: a study of 47 European lakes along a north-south transect. Journal of Paleolimnology. 41 (3), 523540.Google Scholar
Blumenberg, M., Seifert, R., Reitner, J., Pape, T., and Michaelis, W. 2004. Membrane lipid patterns typify distinct anaerobic methanotrophic consortia. Proceedings of the National Academies of Sciences USA, 1011:1111111116 doi:10.1073/pnas.0401188101.Google Scholar
Brassell, S., Wardroper, A., Thomson, I., Maxwell, J., and Eglinton, G. 1981. Specific acyclic isoprenoids as biological markers of methanogenic bacteria in marine sediments. Nature, 290:693.Google Scholar
Brochier-Armanet, C., Boussau, B., Gribaldo, S., and Forterre, P. 2008. Mesophilic Crenarchaeota: proposal for a third archaeal phylum, the Thaumarchaeota. Nature Reviews Microbiology, 6:245252.CrossRefGoogle ScholarPubMed
Castañeda, I., Schefuß, E., Pätzold, J., Sinninghe Damsté, J. S., Weldeab, S., and Schouten, S. 2010. Millennial-scale sea surface temperature changes in the eastern Mediterranean (Nile River Delta region) over the last 27,000 years. Paleoceanography, 25:PA1208.CrossRefGoogle 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
Chazen, C., 2011. Holocene climate evolution of the eastern tropical Pacific told from high resolution climate records from the Peru margin and equatorial upwelling regions. , Brown University.Google Scholar
Church, M., Wai, B., Karl, D., and DeLong, E. 2010. Abundances of crenarchaeal amoA genes and transcripts in the Pacific Ocean. Environmental Microbiology, 12:679688.Google Scholar
de la Torre, J., Walker, C., Ingalls, A., Könneke, M., and Stahl, D. 2008. Cultivation of a thermophilic ammonia oxidizing archaeon synthesizing crenarchaeol. Environmental Microbiology, 10:810818.Google Scholar
de Rosa, M., Esposito, E., Gambacorta, A., Nicolaus, B., and Bu'Lock, J. 1980. Effects of temperature on ether lipid composition of Caldariella acidophila . Phytochemistry, 19:827831.CrossRefGoogle Scholar
Eder, W., Schmidt, M., Koch, M., Garbe-Schönberg, D., and Huber, R. 2002. Prokaryotic phylogenetic diversity and corresponding geochemical data of the brine–seawater interface of the Shaban Deep, Red Sea. Environmental Microbiology, 4:758763.Google Scholar
Fawcett, P.J., Werne, J. P., Anderson, R. S., Heikoop, J. M., Brown, E. T., Berke, M. A., Smith, S. J., Goff, F., Donohoo-Hurley, L., Cisneros-Dozal, L. M., Schouten, S., Sinninghe Damsté, J. S., Huang, Y., Toney, J., Fessenden, J., WoldeGabriel, G., Atudorei, V., Geissman, J. W., and Allen, C. D. 2011. Extended megadroughts in the southwestern United States during Pleistocene interglacials. Nature, 470:518521.Google Scholar
Galand, P., Gutierrez-Provecho, C., Massana, R., Gasol, J., and Casamayora, E. 2010. Interannual recurrence of archaeal assemblages in the coastal NW Mediterranean Sea (Blanes Bay Microbial Observatory). Limnology and Oceanography, 55:21172125.CrossRefGoogle Scholar
Gliozzi, A., Paoli, G., De Rosa, M., and Gambacorta, A. 1983. Effect of isoprenoid cyclization on the transition temperature of lipids in thermophilic archaebacteria. Biochimica et Biophysica Acta (BBA)—Biomembranes, 735:234242.Google Scholar
Herbert, T. D., 2003. Alkenone paleotemperature determinations. Treatise on Geochemistry, 6:391432.Google Scholar
Herndl, G., Reinthaler, T., Teira, E., Van Aken, H., Veth, C., Pernthaler, A., and Pernthaler, J. 2005. Contribution of Archaea to total prokaryotic production in the deep Atlantic Ocean. Applied Environmental Microbiology, 71:23032309.CrossRefGoogle ScholarPubMed
Ho, S., Yamamoto, M., Mollenhauer, G., and Minagawa, M. 2011. Core top TEX86 values in the south and equatorial Pacific. Organic Geochemistry, 42:9499.Google Scholar
Hoefs, M., Schouten, S., De Leeuw, J., King, L., Wakeham, S., and Sinninghe Damsté, J. S. 1997. Ether lipids of planktonic archaea in the marine water column. Applied Environmental Microbiology, 63:30903095.Google Scholar
Hollibaugh, J., Gifford, S., Sharma, S., Bano, N., and Moran, M. 2010. Metatranscriptomic analysis of ammonia-oxidizing organisms in an estuarine bacterioplankton assemblage. The ISME Journal, 5:866878.Google Scholar
Hopmans, E. C., Schouten, S., Pancost, R. D., van der Meer, M. T. J., and Sinninghe Damsté, J. S. 2000. Analysis of intact tetraether lipids in archaeal cell material and sediments by high performance liquid chromatography/atmospheric pressure chemical ionization mass spectrometry. Rapid Communications in Mass Spectrometry 14:585589.Google Scholar
Hopmans, E. C., Weijers, J. W. H., Schefuß, E., Herfort, L., Sinninghe Damsté, J. S., and Schouten, S. 2004. A novel proxy for terrestrial organic matter in sediments based on branched and isoprenoid tetraether lipids. Earth and Planetary Sciences Letters, 224:107116.CrossRefGoogle Scholar
Huber, M., 2010. Why improving molecular and isotopic proxy paleoclimate records is the most important problem in science today: A climate modeling perspective. In Gordon Research Conference. Holderness, NH, USA.Google Scholar
Huguet, C., De Lange, G. J., Gustafsson, O., Middelburg, J. J., Sinninghe Damsté, J. S., and Schouten, S. 2008. Selective preservation of soil organic matter in oxidized marine sediments (Madeira Abyssal Plain). Geochimica et Cosmochimica Acta, 72:60616068.Google Scholar
Huguet, C., Kim, J., de Lange, G., Sinninghe Damsté, J. S., and Schouten, S. 2009. Effects of long term oxic degradation on the TEX86 and BIT organic proxies. Organic Geochemistry, 40:11881194.CrossRefGoogle Scholar
Huguet, C., Schimmelmann, A., Thunell, R., Lourens, L., Sinninghe Damsté, J. S., and Schouten, S. 2007. A study of the TEX86 paleothermometer in the water column and sediments of the Santa Barbara Basin, California. Paleoceanography, 22:PA3203. doi:10.1029/2006PA001310.CrossRefGoogle Scholar
Ingalls, A., Shah, S., Hansman, R., Aluwihare, L., Santos, G., Druffel, E., and Pearson, A. 2006. Quantifying archaeal community autotrophy in the mesopelagic ocean using natural radiocarbon. Proceedings of the National Academies of Sciences USA 103:64426447.Google Scholar
Ionescu, D., Penno, S., Haimovich, M., Rihtman, B., Goodwin, A., Schwartz, D., Hazanov, L., Chernihovsky, M., Post, A., and Oren, A. 2009. Archaea in the Gulf of Aqaba. FEMS Microbiology Ecology, 69:425438.CrossRefGoogle ScholarPubMed
Karner, M., DeLong, E., and Karl, D. 2001. Archaeal dominance in the mesopelagic zone of the Pacific Ocean. Nature, 409:507510.Google Scholar
Kim, J., Van der Meer, J., Schouten, S., Helmke, P., Willmott, V., Sangiorgi, F., Koc, N., Hopmans, E., Sinninghe Damsté, J. S. 2010. New indices and calibrations derived from the distribution of crenarchaeal isoprenoid tetraether lipids: Implications for past sea surface temperature reconstructions. Geochimica et Cosmochimica Acta, 74:46394654.Google Scholar
Kim, J.-H., Huguet, C., Zonneveld, K. A. F., Versteegh, G. J. M., Roeder, W., Sinninghe Damsté, J. S., and Schouten, S. 2009. An experimental field study to test the stability of lipids used for the TEX86 and UK0 37 palaeothermometers. Geochimica et Cosmochimica Acta, 73:28882898.CrossRefGoogle Scholar
Kim, J.-H., Schouten, S., Hopmans, E. C., Donner, B., and Sinninghe Damsté, J. S. 2008. Global sediment core-top calibration of the TEX86 paleothermometer in the ocean. Geochimica et Cosmochimica Acta 72 (4), 11541173.Google Scholar
Kita, M., Aibara, S., Kato, M., Ishinaga, M., and Hata, T. 1973. Effect of changes in fatty acid composition of phospholipid species on the b-galactoside transport system of Escherichia coli K-12. Biochimica et Biophysica Acta (BBA)-Biomembranes 298:6974.Google Scholar
Könneke, M., Bernhard, A. E., De La Torre, J. R., Walker, C. B., Waterbury, J. B., and Stahl, D. A. 2005. Isolation of an autotrophic ammonia-oxidizing marine archaeon. Nature, 437:543546.Google Scholar
Lee, K., Kim, J., Wilke, I., Helmke, P., and Schouten, S. 2008. A study of the alkenone, TEX86, and planktonic foraminifera in the Benguela Up-welling System: Implications for past sea surface temperature estimates. Geochemistry Geophysics Geosystems, 9:Q10019.Google Scholar
Leider, A., Hinrichs, K., Mollenhauer, G., and Versteegh, G. 2010. Core-top calibration of the lipid-based UK0 37 and TEX86 temperature proxies on the southern Italian shelf (SW Adriatic Sea, Gulf of Taranto). Earth and Planetary Sciences Letters, 300:112124.Google Scholar
Leininger, S., Urich, T., Schloter, M., Schwark, L., Qi, J., Nicol, G., Prosser, J., Schuster, S., and Schleper, C. 2006. Archaea predominate among ammonia-oxidizing prokaryotes in soils. Nature, 442:806809.Google Scholar
Liu, X., Leider, A., Gillespie, A., Gröger, J., Versteegh, G., and Hinrichs, K. 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, Z., Pagani, M., Zinniker, D., DeConto, R., Huber, M., Brinkhuis, H., Shah, S., Leckie, R., and Pearson, A. 2009. Global cooling during the Eocene-Oligocene climate transition. Science, 323:11871190.Google Scholar
Llirós, M., Gich, F., Plasencia, A., Auguet, J., Darchambeau, F., Casamayor, E., Descy, J., and Borrego, C. 010. Vertical distribution of ammonia-oxidizing crenarchaeota and methanogens in the epipelagic waters of Lake Kivu (Rwanda-Democratic Republic of the Congo). Applied Environmental Microbiology, 76:8536863.Google Scholar
Lopes Dos Santos, R., Prange, M., Castañeda, I., Schefuß, E., Mulitza, S., Schulz, M., Niedermeyer, E., Sinninghe Damsté, J., and Schouten, S. 2010. Glacial-interglacial variability in Atlantic meridional overturning circulation and thermocline adjustments in the tropical North Atlantic. Earth and Planetary Sciences Letters, 300:407414.Google Scholar
Marlowe, I., Green, J., Neal, A., Brassell, S., Eglinton, G., and Course, P. 1984. Long chain (n-C37—C39) alkenones in the Prymnesiophyceae. Distribution of alkenones and other lipids and their taxonomic significance. British Phycological Journal, 19:203216.Google Scholar
Massana, R., Murray, A., Preston, C., and DeLong, E. 1997. Vertical distribution and phylogenetic characterization of marine planktonic Archaea in the Santa Barbara Channel. Applied Environmental Microbiology 63 (1), 5056.Google Scholar
Nozawa, Y., Iida, H., Fukushima, H., Ohki, K., and Ohnishi, S. 1974. Studies on Tetrahymena membranes: Temperature-induced alterations in fatty acid composition of various membrane fractions in Tetrahymena pyriformis and its effect on membrane fluidity as inferred by spin-label study. Biochimica et Biophysica Acta (BBA)-Biomembranes, 367:134147.Google Scholar
Ochsenreiter, T., Selezi, D., Quaiser, A., Bonch-Osmolovskaya, L., and Schleper, C. 2003. Diversity and abundance of Crenarchaeota in terrestrial habitats studied by 16S RNA surveys and real time PCR. Environmental Microbiology, 5:787797.Google Scholar
Pancost, R., Taylor, K., Handley, L., Huber, M.M., and Hollis, C. 2011. A Critical Evaluation of High TEX86-derived Sea Surface Temperatures from the Early Eocene. In: AGU Fall Meeting Abstracts, 1:6.Google Scholar
Pearson, A., Huang, Z., Ingalls, A., Romanek, C., Wiegel, J., Freeman, K., Smittenberg, R., and Zhang, C. 2004. Nonmarine crenarchaeol in Nevada hot springs. Applied Environmental Microbiology, 70:5229.Google Scholar
Pearson, A., McNichol, A. P., Benitez-Nelson, B. C., Hayes, J. M., and Eglinton, T. I. 2001. Origins of lipid biomarkers in Santa Monica Basin surface sediment: A case study using compound-specific 14C analysis. Geochimica et Cosmochimica Acta, 65:31233137.Google Scholar
Pearson, E., Juggins, S., Talbot, H., Weckstörm, J., Rosen, P., Ryves, D., Roberts, S., and Schmidt, R. 2011. A lacustrine GDGT-temperature calibration from the Scandinavian Arctic to Antarctic: Renewed potential for the application of GDGTpaleothermometry in lakes. Geochimica et Cosmochimica Acta, 75:62256238.Google Scholar
Peterse, F., Hopmans, E., Schouten, S., Mets, A., Rijpstra, W., and Sinninghe Damsté, J. 2011. Identification and distribution of intact polar branched tetraether lipids in peat and soil. Organic Geochemistry, 42:10071015.Google Scholar
Pitcher, A., Rychlik, N., Hopmans, E., Spieck, E., Rijpstra, W., Ossebaar, J., Schouten, S., Wagner, M., and Sinninghe Damsté, J. 2010. Crenarchaeol dominates the membrane lipids of Candidatus Nitrososphaera gargensis, a thermophilic Group I. lb Archaeon. The ISME Journal, 4:542552.Google Scholar
Pitcher, A., Wuchter, C., Siedenberg, K., Schouten, S., and Sinninghe Damsté, J. S. 2011. Crenarchaeol tracks winter blooms of ammonia-oxidizing Thaumarchaeota in the coastal North Sea. Limnology and Oceanography, 56:23082318.Google Scholar
Powers, L. A., Johnson, T. C., Werne, J. P., Castañeda, I. S., Hopmans, E. C., Sinninghe Damsté, J. S., and Schouten, S. 2005. Large temperature variability in the southern African tropics since the Last Glacial Maximum. Geophysical Research Letters 32: L08706. doi:10.1029/2004GL022014.Google Scholar
Powers, L. A., Werne, J. P., Johnson, T. C., Hopmans, E. C., Sinninghe Damsté, J. S., and Schouten, S. 2004. Crenarchaeotal membrane lipids in lake sediments: A new paleotemperature proxy for continental paleoclimate reconstruction? Geology, 32:613616.Google Scholar
Powers, L. A., Werne, J. P., Vanderwoude, A. J., Sinninghe Damsté, J. S., Hopmans, E. C., and Schouten, S. 2010. Applicability and calibration of the TEX86 paleothermometer in lakes. Organic Geochemistry, 41:404413.Google Scholar
Ray, P., White, D., and Brock, T. 1971. Effect of temperature on the fatty acid composition of Thermus aquaticus. Journal of Bacteriology 106:25.Google Scholar
Rueda, G., Rosell-Melé, A., Escala, M., Gyllencreutz, R., and Backman, J. 2009. Comparison of instrumental and GDGT-based estimates of sea surface and air temperatures from the Skagerrak. Organic Geochemistry, 40:287291.Google Scholar
Schouten, S., Forster, A., Panoto, E., and Sinninghe Damsté, J. S. 2007a. Towards calibration of the TEX86 palaeothermometer for tropical sea surface temperatures in ancient greenhouse worlds. Organic Geochemistry, 38:15371546.Google Scholar
Schouten, S., Hopmans, E., Baas, M., Boumann, H., Standfest, S., Könneke, M., Stahl, D., and Sinninghe Damsté, J. 2008a. Intact membrane lipids of “Candidatus Nitrosopumilus maritimus,” a cultivated representative of the cosmopolitan mesophilic group I crenarchaeota. Applied Environmental Microbiology, 74:2433.Google Scholar
Schouten, S., Hopmans, E. C., Forster, A., van Breugel, Y., Kuypers, M. M. M., and Sinninghe Damsté, J. S. 2003. Extremely high sea-surface temperatures at low latitudes during the middle Cretaceous as revealed by archaeal membrane lipids. Geology, 31:10691072.CrossRefGoogle Scholar
Schouten, S., Hopmans, E. C., Schefuß, E.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 Sciences Letters. 204:265274.Google Scholar
Schouten, S., Hopmans, E. C., and Sinninghe Damsté, J. S. 2004. The effect of maturity and depositional redox conditions on archaeal tetraether lipid palaeothermometry. Organic Geochemistry, 35:567571.Google Scholar
Schouten, S., Huguet, C., Hopmans, E. C., Kienhuis, M. V. M., and Sinninghe Damsté, J. S. 2007b. Analytical methodology for TEX86 paleothermometry by high-performance liquid chromatography/atmospheric pressure chemical ionization-mass spectrometry. Analytical Chemistry 79:29402944.CrossRefGoogle ScholarPubMed
Schouten, S., Rijpstra, W., Durisch-Kaiser, E., Schubert, C., and Sinninghe Damsté, J. 2012. Distribution of glycerol dialkyl glycerol tetraether lipids in the water column of Lake Tanganyika. Organic Geochemistry, in press.Google Scholar
Schouten, S., van der Meer, M., Hopmans, E., and Sinninghe Damsté, J. 2008b. Comment on “Lipids of marine Archaea: Patterns and provenance in the water column and sediments” by Turich et al. (2007). Geochimica et Cosmochimica Acta, 72:53425346.Google Scholar
Seki, O., Sakamoto, T., Sakai, S., Schouten, S., Hopmans, E., Sinninghe Damsté, J., and Pancost, R. 2009. Large changes in seasonal sea ice distribution and productivity in the Sea of Okhotsk during the deglaciations. Geochemistry Geophysics Geosystems, 10:Q10007.Google Scholar
Shah, S. R., Mollenhauer, G., Ohkouchi, N., Eglinton, T. I., and Pearson, A. 2008. Origins of archaeal tetraether lipids in sediments: Insights from radiocarbon analysis. Geochimica et Cosmochimica Acta, 72:45774594.Google Scholar
Shevenell, A., Ingalls, A., Domack, E., and Kelly, C. 2011. Holocene Southern Ocean surface temperature variability west of the Antarctic Peninsula. Nature, 470:250254.Google Scholar
Shintani, T., Yamamoto, M., and Chen, M. 2010. Paleoenvironmental changes in the northern South China Sea over the past 28,000 years: A study of TEX86-derived sea surface temperatures and terrestrial biomarkers. Journal of Asian Earth Sciences, 40:12211229.Google Scholar
Sinninghe Damsté, J., Rijpstra, W., Hopmans, E., Weijers, J., Foesel, B., Overmann, J., and Dedysh, S. 2011. 13, 16-Dimethyl octacosanedioic acid (iso-diabolic acid), a common membrane-spanning lipid of Acidobacteria subdivisions 1 and 3. Applied Environmental Microbiology, 77:41474154.Google Scholar
Sinninghe Damsté, J., Rijpstra, W., and Reichart, G. 2002a. The influence of oxic degradation on the sedimentary biomarker record II. Evidence from Arabian Sea sediments. Geochimica et Cosmochimica Acta, 66:27372754.Google Scholar
Sinninghe Damsté, J. S., Hopmans, E. C., Schouten, S., Van Duin, A. C. T., and Geenevasen, J. A. J. 2002b. Crenarchaeol: the characteristic core glycerol dibiphytanyl glycerol tetraether membrane lipid of cosmopolitan pelagic crenarchaeota. Journal of Lipid Research, 43:16411651.Google Scholar
Sinninghe Damsté, J. S., Ossebaar, J., Abbas, B., Schouten, S., and Verschuren, D. 2009. Fluxes and distribution of tetraether lipids in an equatorial African lake: Constraints on the application of the TEX86 palaeothermometer and BIT index in lacustrine settings. Geochimica et Cosmochimica Acta, 73:42324249.Google Scholar
Sluijs, A., Bijl, P. K., Schouten, S., Röhl, U., Reichart, G.-J., and Brinkhuis, H. 2011. Southern ocean warming, sea level and hydrological change during the Paleocene-Eocene thermal maximum. Climate of the Past 7:4761.CrossRefGoogle Scholar
Sluijs, A., Schouten, S., Pagani, M., Woltering, M., Brinkhuis, H., Sinninghe Damsté, J. S., Dickens, G. R., Huber, M., Reichart, G.-J., Stein, R., Matthiessen, J., Lourens, L. J., Pedentchouk, N., Backman, J., and Moran, K. 2006. Subtropical arctic ocean temperatures during the Palaeocene/Eocene thermal maximum. Nature, 441:610613.Google Scholar
Spang, A., Hatzenpichler, R., Brochier-Armanet, C., Rattei, T., Tischler, P., Spieck, E., Streit, W., Stahl, D., Wagner, M., and Schleper, C. 2010. Distinct gene set in two different lineages of ammonia-oxidizing archaea supports the phylum Thaumarchaeota. Trends In Microbiology, 18:331340.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
Sun, Q., Chu, G., Liu, M., Xie, M., Li, S., Ling, Y., Wang, X., Shi, L., Jia, G., and , H. 2011. Distributions and temperature dependence of branched glycerol dialkyl glycerol tetraethers in recent lacustrine sediments from China and Nepal. Journal of Geophysical Research–Biogeoscience, 116:G01008.Google Scholar
Thierstein, H., Geitzenauer, K., and Molfino, B. Shackleton, N., 1977. Global synchroneity of late Quaternary coccolith datum levels: Validation by oxygen isotopes. Geology, 5:400.Google Scholar
Tierney, J., Mayes, M., Meyer, N., Johnson, C., Swarzenski, P., Cohen, A., and Russell, J. 2010a. Late-twentieth-century warming in Lake Tanganyika unprecedented since AD 500. Nature Geoscience, 3:422425.Google Scholar
Tierney, J., Schouten, S., Pitcher, A., Hopmans, E., and Sinninghe Damsté, J. 2012. Core and intact polar glycerol dialkyl glycerol tetraethers (GDGTs) in Sand Pond, Warwick, Rhode Island (USA): Insights into the origin of lacustrine GDGTs. Geochimica et Cosmochimica Acta, 77:561581.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
Tierney, J. E., Russell, J. M., Eggermont, H., Hopmans, E. C., Verschuren, D., and Sinninghe Damsté, J. S. 2010b. Environmental controls on branched tetraether lipid distributions in tropical East African lake sediments. Geochimica et Cosmochimica Acta, 74:49024918.CrossRefGoogle Scholar
Tierney, J. E., Russell, J. M., Huang, Y., Sinninghe Damsté, J. S., Hopmans, E. C., and Cohen, A. S. 2008. Northern Hemisphere controls on tropical southeast African climate during the past 60,000 years. Science, 322:252255.Google Scholar
Trommer, G., Siccha, M., van der Meer, M., Schouten, S., Sinninghe Damsté, J. S., Schulz, H., Hemleben, C.C. and Kucera, M. 2009. Distribution of Crenarchaeota tetraether membrane lipids in surface sediments from the Red Sea. Organic Geochemistry, 40:724731.CrossRefGoogle Scholar
Turich, C., Freeman, K. H., Bruns, M. A., Conte, M. H., Jones, A. D., and Wakeham, S. G. 2007. Lipids of marine archaea: Patterns and provenance in the water-column and sediments. Geochimica et Cosmochimica Acta, 71:32723291.Google Scholar
Tyler, J., Nederbragt, A., Jones, V., and Thurow, J. 2010. Assessing past temperature and soil pH estimates from bacterial tetraether membrane lipids: Evidence from the recent lake sediments of Lochnagar, Scotland. Journal of Geophysical Research–Biogeoscience, 115:G01015.Google Scholar
Uda, I., Sugai, A., Itoh, Y., and Itoh, T.T. 2001. Variation in molecular species of polar lipids from Thermoplasma acidophilum depends on growth temperature. Lipids, 36:103105.Google Scholar
Volkman, J., Eglinton, G., Corner, E., and Forsberg, T. 1980. Long-chain alkenes and alkenones in the marine coccolithophorid Emiliania huxleyi . Phytochemistry, 19:26192622.Google Scholar
Wakeham, S. G., Lewis, C., Hopmans, E. C., Schouten, S., and Sinninghe Damsté, J. S. 2003. Archaea mediate anaerobic oxidation of methane in deep euxinic waters of the Black Sea. Geochimica et Cosmochimica Acta 67:13591374.Google Scholar
Weijers, J. W. H., Panoto, E., van Bleijswijk, J., Schouten, S., Rijpstra, W. I. C., Balk, M., Stams, A. J. M., and Sinninghe Damsté, J. 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., Schouten, S., and Sinninghe Damsté, J. S. 2007a. Coupled thermal and hydrological evolution of tropical Africa over the last deglaciation. Science, 315:17011704.Google Scholar
Weijers, J. W H., Schouten, S., Hopmans, E. C., Geenevasen, J. A. J., David, O. R. P., Coleman, J. M., Pancost, R. D., and Sinninghe Damsté, J. S. 2006a. Membrane lipids of mesophilic anaerobic bacteria thriving in peats have typical archaeal traits. Environmental Microbiology, 8:648657.Google Scholar
Weijers, J. W H., Schouten, S., Sluijs, A., Brinkhuis, H., and Sinninghe Damsté, J. S. 2007b. Warm arctic continents during the Palaeocene-Eocene thermal maximum. Earth and Planetary Sciences Letters, 261:230238.Google Scholar
Weijers, J. W. H., Schouten, S., Spaargaren, O. C., and Sinninghe Damsté, J. S. 2006b. 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. 2007c. Environmental controls on bacterial tetraether membrane lipid distribution in soils. Geochimica et Cosmochimica Acta, 71:703713.Google Scholar
Wuchter, C., Schouten, S., Coolen, M. J. L., and Sinninghe Damsté, J. S. 2004. Temperature-dependent variation in the distribution of tetraether membrane lipids of marine Crenarchaeota: Implications for TEX86 paleothermometry. Paleoceanography, 19:PA4028.Google Scholar
Wuchter, C., Schouten, S., Wakeham, S. G., and Sinninghe Damsté, J. S. 2005. Temporal and spatial variation in tetraether membrane lipids of marine crenarchaeota in particulate organic matter: Implications for TEX86 paleothermometry. Paleoceanography, 20:PA3013.Google Scholar
Wuchter, C., Schouten, S., Wakeham, S. G., and Sinninghe Damsté, J. S. 2006. Archaeal tetraether membrane lipid fluxes in the northeastern Pacific and the Arabian Sea: Implications for TEX86 paleothermometry. Paleoceanography, 21:PA4208.Google Scholar
Yang, H., Ding, W., Zhang, C., Wu, X., Ma, X., He, G., Huang, J., and Xie, S. 2011. Occurrence of tetraether lipids in stalagmites: Implications for sources and GDGT-based proxies. Organic Geochemistry, 42:108115.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, Y., Zhang, C., Liu, X., Li, L., Hinrichs, K., and Noakes, J. 2011. Methane Index: A tetraether archaeal lipid biomarker indicator for detecting the instability of marine gas hydrates. Earth and Planetary Science Letters, 307:525534.Google Scholar
Zink, K., Vandergoes, M., Mangelsdorf, K., Dieffenbacher-Krall, A., and Schwark, L.L. 2010. Application of bacterial glycerol dialkyl glycerol tetraethers (GDGTs) to develop modern and past temperature estimates from New Zealand lakes. Organic Geochemistry 41:10601066.Google Scholar