Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-x4r87 Total loading time: 0 Render date: 2024-04-26T01:37:16.231Z Has data issue: false hasContentIssue false

The TEX86 Paleotemperature Proxy

Published online by Cambridge University Press:  26 September 2020

Gordon N. Inglis
Affiliation:
University of Southampton
Jessica E. Tierney
Affiliation:
University of Arizona

Summary

The TEX86 paleothermometer is based upon the distribution of archaeal membrane lipids ('GDGTs') in marine sediments. GDGTs are ubiquitous, abundant and relatively resistant to degradation; as such, the TEX86 paleothermometer has been used to reconstruct sea surface temperature (SST) during the Cenozoic and early Mesozoic. We review the principles of the TEX86 proxy and developments made over the last two decades. We also discuss its application as a paleotemperature proxy and explore existing challenges and limitations.
Get access
Type
Element
Information
Online ISBN: 9781108846998
Publisher: Cambridge University Press
Print publication: 22 October 2020

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

Anagnostou, E., John, E. H., Edgar, K. M., et al. (2016) Changing atmospheric CO2 concentration was the primary driver of early Cenozoic climate. Nature 533, 380–4.CrossRefGoogle ScholarPubMed
Bijl, P. K., Schouten, S., Sluijs, A., Reichart, G.-J., Zachos, J. C. and Brinkhuis, H. (2009) Early Palaeogene temperature evolution of the southwest Pacific Ocean. Nature 461, 776–9.Google 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, 523–40.Google Scholar
Brassell, S., Eglinton, G., Marlowe, I., Pflaumann, U. and Sarnthein, M. (1986) Molecular stratigraphy: A new tool for climatic assessment. Nature 320, 129.CrossRefGoogle Scholar
Brassell, S. C. (2014) Climatic influences on the Paleogene evolution of alkenones. Paleoceanography 29(3), 255–72. https://doi.org/10.1002/2013PA002576Google Scholar
Church, M. J., Wai, B., Karl, D. M. and DeLong, E. F. (2010) Abundances of crenarchaeal amoA genes and transcripts in the Pacific Ocean. Environmental Microbiology 12, 679–88.Google Scholar
Cramwinckel, M. J., Huber, M., Kocken, I. J., et al. (2018) Synchronous tropical and polar temperature evolution in the Eocene. Nature 559, 382.Google Scholar
De la Torre, J. R., Walker, C. B., Ingalls, A. E., Könneke, M. and Stahl, D. A. (2008) Cultivation of a thermophilic ammonia oxidizing archaeon synthesizing crenarchaeol. Environmental Microbiology 10, 810–18.CrossRefGoogle ScholarPubMed
DeLong, E. F., King, L. L., Massana, R., et al. (1998) Dibiphytanyl ether lipids in nonthermophilic crenarchaeotes. Applied and Environmental Microbiology 64, 1133–8.Google Scholar
De Rosa, M., de Rosa, S., Gambacorta, A., Minale, L. and Bu’lock, J. D. (1977) Chemical structure of the ether lipids of thermophilic acidophilic bacteria of the Caldariella group. Phytochemistry 16, 1961–5.Google Scholar
De Rosa, M., Esposito, E., Gambacorta, A., Nicolaus, B. and Bu’Lock, J. D. (1980) Effects of temperature on ether lipid composition of Caldariella acidophila. Phytochemistry 19, 827–31.Google Scholar
dos Santos, R. A. L., Prange, M., Castañeda, I. S., et al.(2010) Glacial–interglacial variability in Atlantic meridional overturning circulation and thermocline adjustments in the tropical North Atlantic. Earth and Planetary Science Letters 300, 407–14.Google Scholar
Eley, Y.L., Thompson, W., Greene, S.E., Mandel, I., Edgar, K., Bendle, J.A. and Dunkley Jones, T., 2019. OPTiMAL: A new machine learning approach for GDGT-based palaeothermometry. Climate of the Past Discussions, pp.1–39. https://doi.org/10.5194/cp-2019-60CrossRefGoogle Scholar
Elling, F. J., Könneke, M., Lipp, J. S., Becker, K. W., Gagen, E. J. and Hinrichs, K.-U. (2014) Effects of growth phase on the membrane lipid composition of the thaumarchaeon Nitrosopumilus maritimus and their implications for archaeal lipid distributions in the marine environment. Geochimica et Cosmochimica Acta 141, 579–97.CrossRefGoogle Scholar
Elling, F. J., Könneke, M., Mußmann, M., Greve, A., and Hinrichs, K.-U. (2015) Influence of temperature, pH, and salinity on membrane lipid composition and TEX 86 of marine planktonic thaumarchaeal isolates. Geochimica et Cosmochimica Acta 171, 238–55.Google Scholar
Evans, T. W., Könneke, M., Lipp, J. S., et al. (2018) Lipid biosynthesis of Nitrosopumilus maritimus dissected by lipid specific radioisotope probing (lipid-RIP) under contrasting ammonium supply. Geochimica et Cosmochimica Acta 242, 5163.Google Scholar
Francis, C. A., Roberts, K. J., Beman, J. M., Santoro, A. E. and Oakley, B. B. (2005) Ubiquity and diversity of ammonia-oxidizing archaea in water columns and sediments of the ocean. Proceedings of the National Academy of Sciences of the USA 102, 14683–8.Google Scholar
Frieling, J., Gebhardt, H., Huber, M., et al. (2017) Extreme warmth and heat-stressed plankton in the tropics during the Paleocene-Eocene Thermal Maximum. Science Advances 3, e1600891.CrossRefGoogle ScholarPubMed
Herfort, L., Schouten, S., Boon, J. P. and Sinninghe Damsté, J. S. (2006) Application of the TEX 86 temperature proxy to the southern North Sea. Organic Geochemistry 37, 1715–26.CrossRefGoogle Scholar
Hoefs, M., Schouten, S., De Leeuw, J., King, L. L., Wakeham, S. G. and Sinninghe Damsté, J. S. (1997) Ether lipids of planktonic archaea in the marine water column. Applied and Environmental Microbiology 63, 3090–5.CrossRefGoogle ScholarPubMed
Hollis, C. J., Dunkley Jones, T., Anagnostou, E., et al. (2019) The DeepMIP contribution to PMIP4: Methodologies for selection, compilation and analysis of latest Paleocene and early Eocene climate proxy data, incorporating version 0.1 of the DeepMIP database. Geoscientific Model Development 12, 3149–206.Google Scholar
Hollis, C. J., Taylor, K. W. R., Handley, L., et al.(2012) Early Paleogene temperature history of the Southwest Pacific Ocean: Reconciling proxies and models. Earth and Planetary Science Letters 349 350, 5366.Google Scholar
Hopmans, E. C., Schouten, S., Pancost, R. D., van der Meer, M. T. 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, 585–9.Google Scholar
Hopmans, E. C., Schouten, S. and Sinninghe Damsté, J. S. (2016) The effect of improved chromatography on GDGT-based palaeoproxies. Organic Geochemistry 93, 16.Google Scholar
Hopmans, E. C., Weijers, J. W., 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 Science Letters 224, 107–16.CrossRefGoogle Scholar
Horak, R.E., Qin, W., Schauer, A.J., (2013) Ammonia oxidation kinetics and temperature sensitivity of a natural marine community dominated by Archaea. The ISME Journal, 7, 2023.Google Scholar
Huguet, C., Cartes, J. E., Sinninghe Damsté, J. S. and Schouten, S. (2006) Marine crenarchaeotal membrane lipids in decapods: Implications for the TEX86 paleothermometer. Geochemistry, Geophysics, Geosystems 7. https://doi.org/10.1029/2006GC001305Google Scholar
Huguet, C., Kim, J.-H., de Lange, G. J., Sinninghe Damsté, J. S. and Schouten, S. (2009) Effects of long term oxic degradation on the U37 K′, TEX86 and BIT organic proxies. Organic Geochemistry 40, 1188–94.Google Scholar
Huguet, C., Martrat, B., Grimalt, J. O., Sinninghe Damsté, J. S. and Schouten, S. (2011) Coherent millennial‐scale patterns in U37 k′ and TEX86 H temperature records during the penultimate interglacial‐to‐glacial cycle in the western Mediterranean. Paleoceanography 26. DOI: 10.1029/2010PA002048.Google Scholar
Hurley, S. J., Elling, F. J., Könneke, M., et al. (2016) Influence of ammonia oxidation rate on thaumarchaeal lipid composition and the TEX86 temperature proxy. Proceedings of the National Academy of Sciences of the USA 113, 7762–7.Google Scholar
Inglis, G. N., Farnsworth, A., Lunt, D., et al. (2015) Descent toward the Icehouse: Eocene sea surface cooling inferred from GDGT distributions. Paleoceanography 30, 1000–20.Google Scholar
Karner, M. B., DeLong, E. F. and Karl, D. M. (2001) Archaeal dominance in the mesopelagic zone of the Pacific Ocean. Nature 409, 507–10.CrossRefGoogle ScholarPubMed
Kim, J. G., Park, S. J., Sinninghe Damsté, J. S., et al. (2016) Hydrogen peroxide detoxification is a key mechanism for growth of ammonia-oxidizing archaea. Proceedings of the National Academy of Sciences of the USA, 113, 7888–93.CrossRefGoogle ScholarPubMed
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, 1154–73.Google Scholar
Kim, J.-H., Schouten, S., Rodrigo-Gámiz, M., et al. (2015) Influence of deep-water derived isoprenoid tetraether lipids on the paleothermometer in the Mediterranean Sea. Geochimica et Cosmochimica Acta 150, 125–41.Google Scholar
Kim, J.-H., Van der Meer, J., Schouten, S., et al. (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, 4639–54.Google Scholar
Kitzinger, K., Padilla, C. C., Marchant, H. K., et al. (2019) Cyanate and urea are substrates for nitrification by Thaumarchaeota in the marine environment. Nature Microbiology 4, 234.Google Scholar
Könneke, M., Bernhard, A. E., José, R., Walker, C. B., Waterbury, J. B. and Stahl, D. A. (2005) Isolation of an autotrophic ammonia-oxidizing marine archaeon. Nature 437, 543–6.Google Scholar
Lengger, S. K., Hopmans, E.C., Sinninghe Damsté, J. S. and Schouten, S. (2014a) Fossilization and degradation of archaeal intact polar tetraether lipids in deeply buried marine sediments (Peru Margin). Geobiology 12, 212–20.CrossRefGoogle ScholarPubMed
Lengger, S. K., Hopmans, E. C., Sinninghe Damsté, J. S. and Schouten, S. J. G. (2014b) Fossilization and degradation of archaeal intact polar tetraether lipids in deeply buried marine sediments (Peru Margin). Geobiology 12, 212–20.Google Scholar
Lengger, S. K., Sutton, P. A., Rowland, S. J., et al. (2018) Archaeal and bacterial glycerol dialkyl glycerol tetraether (GDGT) lipids in environmental samples by high temperature-gas chromatography with flame ionisation and time-of-flight mass spectrometry detection. Organic Geochemistry 121, 1021.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 USA 111, 9858–63.Google Scholar
Liu, X. L., Lipp, J. S., Birgel, D., Summons, R. E. and Hinrichs, K. U. (2018) Predominance of parallel glycerol arrangement in archaeal tetraethers from marine sediments: Structural features revealed from degradation products. Organic Geochemistry 115, 1223.Google Scholar
Liu, Z., Pagani, M., Zinniker, D., et al. (2009) Global cooling during the Eocene-Oligocene climate transition. Science 323, 1187–90.Google Scholar
O’Brien, C. L., Robinson, S. A., Pancost, R. D., et al. (2017) Cretaceous sea-surface temperature evolution: Constraints from TEX 86 and planktonic foraminiferal oxygen isotopes. Earth Science Reviews 172, 224–47.Google Scholar
Pearson, A., Hurley, S. J., Walter, S. R. S., Kusch, S., Lichtin, S. and Zhang, Y. G. (2016) Stable carbon isotope ratios of intact GDGTs indicate heterogeneous sources to marine sediments. Geochimica et Cosmochimica Acta 181, 1835.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 Δ14 C analysis. Geochimica et Cosmochimica Acta 65, 3123–37.Google Scholar
Pearson, A., Pi, Y., Zhao, W., et al. (2008) Factors controlling the distribution of archaeal tetraethers in terrestrial hot springs. Applied and Environmental Microbiology 74, 3523–32.Google Scholar
Pearson, P. N., van Dongen, B. E., Nicholas, C. J., et al. (2007) Stable warm tropical climate through the Eocene Epoch. Geology 35, 211–14.Google Scholar
Pitcher, A., Rychlik, N., Hopmans, E. C., et al. (2010) Crenarchaeol dominates the membrane lipids of Candidatus Nitrososphaera gargensis, a thermophilic Group I. 1b Archaeon. The ISME Journal 4, 542.Google Scholar
Polik, C. A., Elling, F. J. and Pearson, A. (2018) Impacts of paleoecology on the TEX86 sea surface temperature proxy in the Pliocene‐Pleistocene Mediterranean Sea. Paleoceanography and Paleoclimatology 33, 1472–89.Google Scholar
Qin, W., Carlson, L. T., Armbrust, E. V., et al. (2015) Confounding effects of oxygen and temperature on the TEX86 signature of marine Thaumarchaeota. Proceedings of the National Academy of Sciences of the USA 112, 10979–84.Google Scholar
Richey, J. N. and Tierney, J. E. (2016) GDGT and alkenone flux in the northern Gulf of Mexico: Implications for the TEX86 and UK’37 paleothermometers. Paleoceanography 31, 1547–61.CrossRefGoogle Scholar
Robinson, S. A., Ruhl, M., Astley, D. L., (2017) Early Jurassic North Atlantic sea‐surface temperatures from TEX86 palaeothermometry. Sedimentology 64, 215–30.Google Scholar
Sagoo, N., Valdes, P., Flecker, R. and Gregoire, L. J. (2013) The Early Eocene equable climate problem: Can perturbations of climate model parameters identify possible solutions? Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 371.Google Scholar
Schouten, S., Forster, A., Panoto, F. E. and Sinninghe Damsté, J. S. (2007) Towards calibration of the TEX 86 palaeothermometer for tropical sea surface temperatures in ancient greenhouse worlds. Organic Geochemistry 38, 1537–46.CrossRefGoogle Scholar
Schouten, S., Hopmans, E. C., Baas, M., et al. (2008) Intact membrane lipids of “Candidatus Nitrosopumilus maritimus,” a cultivated representative of the cosmopolitan mesophilic Group I Crenarchaeota. Applied and Environmental Microbiology 74, 2433–40.Google Scholar
Schouten, S., Hopmans, E. C., Forster, A., van Breugel, Y., Kuypers, 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, 1069–72.Google Scholar
Schouten, S., Hopmans, E. C., Rosell-Melé, A., et al. (2013) An interlaboratory study of TEX86 and BIT analysis of sediments, extracts, and standard mixtures. Geochemistry, Geophysics, Geosystems 14, 5263–85.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, 265–74.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, 567–71.Google Scholar
Schouten, S., Pitcher, A., Hopmans, E. C., Villanueva, L., van Bleijswijk, J. and Sinninghe Damsté, J. S. (2012) Intact polar and core glycerol dibiphytanyl glycerol tetraether lipids in the Arabian Sea oxygen minimum zone: I. Selective preservation and degradation in the water column and consequences for the TEX86. Geochimica et Cosmochimica Acta 98, 228–43.Google Scholar
Schouten, S., Villanueva, L., Hopmans, E.C., van der Meer, M.T. and Damsté, J.S.S., 2014. Are Marine Group II Euryarchaeota significant contributors to tetraether lipids in the ocean?. Proceedings of the National Academy of Sciences, 111(41), pp.E4285–E4285, https://doi.org/10.1073/pnas.1416176111Google 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, 4577–94.Google Scholar
Sinninghe Damsté, J. S., Rijpstra, W. I. C., Hopmans, E. C., den Uijl, M. J., Weijers, J. W. and Schouten, S. (2018) The enigmatic structure of the crenarchaeol isomer. Organic Geochemistry, 124, 22–8.CrossRefGoogle Scholar
Sinninghe Damsté, J. S., Schouten, S., Hopmans, E. C., Van Duin, A. C. and Geenevasen, J. A. (2002) Crenarchaeol the characteristic core glycerol dibiphytanyl glycerol tetraether membrane lipid of cosmopolitan pelagic crenarchaeota. Journal of Lipid Research 43, 1641–51.Google Scholar
Taylor, K. W., 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, 158–74.Google Scholar
Tierney, J.E. and Tingley, M. P. (2014) A Bayesian, spatially-varying calibration model for the TEX86 proxy. Geochimica et Cosmochimica Acta 127, 83106.Google Scholar
Tierney, J. E. and Tingley, M. P. (2015) A TEX86 surface sediment database and extended Bayesian calibration. Scientific Data 2.CrossRefGoogle ScholarPubMed
Turich, C., Freeman, K. H., Bruns, M. A., Conte, M., 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, 3272–91.Google Scholar
Uda, I., Sugai, A., Itoh, Y. H. and Itoh, T. (2001) Variation in molecular species of polar lipids from Thermoplasma acidophilum depends on growth temperature. Lipids 36, 103–5.CrossRefGoogle ScholarPubMed
Villanueva, L., Schouten, S. and Sinninghe Damsté, J. S. (2015) Depth‐related distribution of a key gene of the tetraether lipid biosynthetic pathway in marine Thaumarchaeota. Environmental Microbiology 17, 3527–39.Google Scholar
Weijers, J. W., 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, 1680–93.Google Scholar
Wörmer, L., Elvert, M., Fuchser, J., et al. (2014) Ultra-high-resolution paleoenvironmental records via direct laser-based analysis of lipid biomarkers in sediment core samples. Proceedings of the National Academy of Sciences of the USA 111, 15669–74.Google Scholar
Wuchter, C., Schouten, S., Boschker, H. T. and Sinninghe Damsté, J. S. (2003) Bicarbonate uptake by marine Crenarchaeota. FEMS Microbiology Letters 219, 203–7.CrossRefGoogle ScholarPubMed
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.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.Google Scholar
Yamamoto, M., Shimamoto, A., Fukuhara, T., Tanaka, Y. and Ishizaka, J. (2012) Glycerol dialkyl glycerol tetraethers and TEX86 index in sinking particles in the western North Pacific. Organic Geochemistry 53, 5262.Google Scholar
Zachos, J. C., Schouten, S., Bohaty, S., et al.(2006) Extreme warming of mid-latitude coastal ocean during the Paleocene-Eocene Thermal Maximum: Inferences from TEX86 and isotope data. Geology 34, 737–40.Google Scholar
Zeng, Z., Liu, X. L., Farley, K. R., et al. (2019) GDGT cyclization proteins identify the dominant archaeal sources of tetraether lipids in the ocean. Proceedings of the National Academy of Sciences of the USA 116, 22505–11.Google Scholar
Zhang, Y. G. and Liu, X. (2018) Export depth of the TEX86 signal. Paleoceanography 33, 666–71.Google Scholar
Zhang, Y. G., Pagani, M. and Wang, Z. (2016) Ring Index: A new strategy to evaluate the integrity of TEX86 paleothermometry. Paleoceanography 31, 220–32.Google Scholar
Zhang, Y. G., Zhang, C. L., Liu, X.-L., Li, L., Hinrichs, K.-U. and Noakes, J. E. (2011) Methane Index: a tetraether archaeal lipid biomarker indicator for detecting the instability of marine gas hydrates. Earth and Planetary Science Letters 307, 525–34.Google Scholar
Zhou, A., Weber, Y., Chui, B. K., et al. (2019) Energy flux controls tetraether lipid cyclization in Sulfolobus acidocaldarius. Environmental Microbiology 22, 343–53.Google Scholar

Save element to Kindle

To save this element to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

The TEX86 Paleotemperature Proxy
Available formats
×

Save element to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

The TEX86 Paleotemperature Proxy
Available formats
×

Save element to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

The TEX86 Paleotemperature Proxy
Available formats
×