Hostname: page-component-848d4c4894-nmvwc Total loading time: 0 Render date: 2024-07-06T18:16:59.623Z Has data issue: false hasContentIssue false

Applying Oxygen Isotope Paleothermometry in Deep Time

Published online by Cambridge University Press:  21 July 2017

Ethan L. Grossman*
Affiliation:
Department of Geology and Geophysics, Texas A&M University, College Station, TX 77843-3115 USA. e-grossman@tamu.edu
Get access

Abstract

Oxygen isotope paleotemperature studies of the Mesozoic and Paleozoic are based mainly on conodonts, belemnite guards, and brachiopod shells—material resistant to diagenesis and generally precipitated in oxygen isotope equilibrium with ambient water. The greatest obstacle to accurate oxygen isotope paleothermometry in deep time is uncertainty in the oxygen isotopic composition of the ambient seawater. The second greatest obstacle is fossil diagenesis. Useful application of the oxygen isotope method to brachiopod shells requires extreme care in sample screening and analyses, and is best done with scanning-electron microscopy, and petrographic and cathodoluminescence microscopy, and trace-element analysis. Correct interpretation of oxygen isotope data is greatly aided by thorough understanding of the paleolatitude, paleoecology, and depositional environment of the samples. The oxygen isotope record for the Triassic, based on brachiopod shells, is too sparse to show any distinct isotopic features. Jurassic and Early Cretaceous δ18O records, based on belemnites, show a Toarcian (Jurassic) decline (warming), a Callovian-Oxfordian acme, and an Early Cretaceous increase (cooling) to a Valanginian-Hauterivian maximum, followed by a decline (warming) to a middle Barremian minimum. Deep-time applications to oxygen isotope thermometry provide evidence for cooling and glaciation in the Ordovician, Carboniferous, and Permian. The δ18O values from Silurian and Devonian brachiopod shells and conodonts average lower than those of the remaining Phanerozoic because of the absence of continental glaciers and possibly higher temperatures (~37°?), although slightly lower (≤2%o) seawater δ18O cannot be ruled out. The hypothesis of high temperatures in the early Paleozoic implies a relatively constant hydrospheric δ18O, which is supported by clumped isotope paleotemperatures. However, more research is needed to develop methods for evaluating clumped isotope reordering in fossils. Ongoing and future research in oxygen isotope and clumped isotope thermometry hold the promise of resolving deep-time temperatures, seawater δ18O, and salinity with heretofore unavailable accuracy (±2°, ±0.4%o, and ±2 psu), providing the environmental setting for the evolution of metazoan life on Earth.

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

Affek, H. 2012. Clumped isotopes paleothermometry. In Ivany, L. C. and Huber, B. T. (eds.), Reconstructing Earth's Deep-Time Climate—The State of the Art in 2012. Paleontological Society Papers 18. Paleontological Society.Google Scholar
Algeo, T. J., and Heckel, P. H. 2008. The Late Pennsylvanian Midcontinent Sea of North America: A review. Palaeogeography, Palaeoclimatology, Palaeoecology, 268(3–4): 205221.Google Scholar
Al-Rousan, S., Al-Moghrabi, S., Patzold, J., and Wefer, G. 2003. Stable oxygen isotopes in Pontes corals monitor weekly temperature variations in the northern Gulf of Aqaba, Red Sea. Coral Reefs, 22(4):346356.CrossRefGoogle Scholar
Anderson, T. F., and Arthur, M. A. 1983. Stable isotopes of oxygen and carbon and their applicaiton to sedimentologic and paleoenvironmental problems, p. 11 to 1–151, Stable Isotopes in Sedimentary Geology. SEPM Short Course No. 10. SEPM.Google Scholar
Anderson, T. F., Popp, B. N., Williams, A. C., Ho, L. Z., and Hudson, J. D. 1994. The stable isotopic records of fossils from the Peterborough member, Oxford Clay formation (Jurassic), UK – Paleoenvironmental implications. Journal of the Geological Society, 151:125138.CrossRefGoogle Scholar
Angiolini, L., Jadoul, F., Leng, M. J., Stephenson, M. H., Rushton, J., Chenery, S., and Crippa, G. 2009. How cold were the Early Permian glacial tropics? Testing sea-surface temperature using the oxygen isotope composition of rigorously screened brachiopod shells. Journal of the Geological Society, 166:933945.Google Scholar
Azmy, K., Veizer, J., Bassett, M. G., and Copper, P. 1998. Oxygen and carbon isotopic composition of Silurian brachiopods: Implications for coeval seawater and glaciations. Geological Society of America Bulletin, 110(11):14991512.Google Scholar
Banner, J. L., and Kaufman, J. 1994. The isotopic record of ocean chemistry and diagenesis preserved in nonluminescent brachiopods from Mississippian carbonate rocks, Illinois and Missouri. Geological Society of America Bulletin, 106(8):10741082.Google Scholar
Bassett, D., Macleod, K. G., Miller, J. E., and Ethington, R. L. 2007. Oxygen isotopic composition of biogenic phosphate and the temperature of Early Ordovician seawater. Palaios, 22(1):98103.CrossRefGoogle Scholar
Beck, W. C., Grossman, E. L., and Morse, J. W. 2005. Experimental studies of oxygen isotope fractionation in the carbonic acid system at 15°, 25°, and 40°C. Geochimica et Cosmochimica Acta, 69(14):34933503.CrossRefGoogle Scholar
Bemis, B. E., Spero, H. J., Bijma, J., and Lea, D. W. 1998. Reevaluation of the oxygen isotopic composition of planktonic foraminifera: Experimental results and revised paleotemperature equations. Paleoceanography, 13(2):150160.Google Scholar
Bickert, , Patzold, J., Samtleben, C., and Munnecke, A. 1997. Paleoenvironmental changes in the Silurian indicated by stable isotopes in brachiopod shells from Gotland, Sweden. Geochimica et Cosmochimica Acta, 61(13):27172730.CrossRefGoogle Scholar
Bigg, G. R., and Rohling, E. J. 2000. An oxygen isotope data set for marine waters. Journal of Geophysical Research-Oceans, 105(C4):85278535.Google Scholar
Brand, U. 1982. The oxygen and carbon isotope composition of Carboniferous fossil components —sea-water effects. Sedimentology, 29(1):139147.CrossRefGoogle Scholar
Brand, U. 1989. Biogeochemistry of late Paleozoic North America brachiopods and secular variation of seawater composition. Biogeochemistry. 7:159193.Google Scholar
Brand, U., Azmy, K., and Veizer, J. 2006. Evaluation of the Salinic I tectonic, Cancaniri glacial and Ireviken biotic events: Biochemostratigraphy of the lower Silurian succession in the Niagara Gorge area, Canada and USA. Palaeogeography Palaeoclimatology Palaeoecology, 241(2):192213.CrossRefGoogle Scholar
Brand, U., and Legrand-Blain, M. 1992. Paleoecology and biogeochemistry of brachiopods from the Devonian-Carboniferous boundary interval of the Griotte formation, La Serre, Montagne Noire, France: Annales de la Société géologique de Belgique, T. 115, fasc. 2:497505.Google Scholar
Brand, U., Logan, A., Hiller, N., and Richardson, J. 2003. Geochemistry of modern brachiopods: applications and implications for oceanography and paleoceanography. Chemical Geology, 198(3–4): 305334.Google Scholar
Brand, U., Tazawa, J., Sano, H., Azmy, K., and Lee, X. Q. 2009. Is mid-late Paleozoic ocean-water chemistry coupled with epeiric seawater isotope records? Geology, 37(9):823826.CrossRefGoogle Scholar
Brenchley, P. J., Marshall, J. D., Carden, G. A. F., Robertson, D. B. R., Long, D. G. F., Meidla, T., Hints, L., and Anderson, T. F. 1994. Bathymetric and isotopic evidence for a short-lived Late Ordovician glaciation in a greenhouse period. Geology, 22(4):295298.Google Scholar
Brenchley, P. J., Carden, G.A.F., and Marshall, J.D. 1995. Environmental changes associated with the “first strike” of the late Ordovician mass extinction. Modern Geology, 20:6982.Google Scholar
Brock, T. D. 1985. Life at high-temperatures. Science, 230(4722):132138.CrossRefGoogle ScholarPubMed
Broecker, W. S. 1989. The salinity contrast between the Atlantic and Pacific oceans during glacial time. Paleoceanography, 4:207212.CrossRefGoogle Scholar
Bruckschen, P., Oesmann, S., and Veizer, J. 1999. Isotope stratigraphy of the European Carboniferous: proxy signals for ocean chemistry, climate and tectonics. Chemical Geology, 161(1–3): 127163.Google Scholar
Bruckschen, P., and Veizer, J. 1997. Oxygen and carbon isotopic composition of Dinantian brachiopods: Paleoenvironmental implications for the Lower Carboniferous of western Europe. Palaeogeography Palaeoclimatology Palaeoecology, 132(1–4): 243264.Google Scholar
Bruckschen, P., Veizer, J., Schwark, L., and Leythaeuser, D. 2001. Isotope stratigraphy for the transition from the late Palaeozoic greenhouse in the Permo—Carboniferous icehouse—new results. Terra Nostra 2001/4:711.Google Scholar
Buggisch, W., Joachimski, M. M., Sevastopulo, G., and Morrow, J. R. 2008. Mississippian δ13Ccarb and conodont apatite δ18O records — Their relation to the Late Palaeozoic glaciation. Palaeogeography Palaeoclimatology Palaeoecology, 268:273292.CrossRefGoogle Scholar
Came, R. E., Eiler, J. M., Veizer, J., Azmy, K., Brand, U., and Weidman, C. R. 2007. Coupling of surface temperatures and atmospheric CO2 concentrations during the Palaeozoic era. Nature, 449:198201, doi:10.1038/nature06085.Google Scholar
Carpenter, S. J., and Lohmann, K.C. 1995. δ18O and δ13C values of modern brachiopod shells. Geochimica et Cosmochimica Acta, 59:37493764.Google Scholar
Carpenter, S. J., Lohmann, K. C., Holden, P., Walter, L. M., Huston, T. J., and Halliday, A. N. 1991. δ18O values, 87Sr/86Sr and Sr/Mg ratios of Late Devonian abiotic marine calcite: Implications for the composition of ancient seawater. Geochimica et Cosmochimica Acta, 55:19912010.Google Scholar
Cecil, C. B., Dulong, F. T., West, R. R., Stamm, R., Wardlaw, B., and Edgar, N. T. 2003. Climate controls on the stratigraphy of a middle Pennsylvanian cyclothem in North America. In Cecil, C. B., and Edgar, N. T. (eds.), Climate Controls on Stratigraphy, SEPM Spec. Publ. no. 77, 151180.Google Scholar
Chafetz, H. S., Wu, Z., Lapen, T. J., and Milliken, K. L. 2008. Geochemistry of preserved Permian aragonitic cements in the tepees of the Guadalupe mountains, west texas and New Mexico, USA. Journal of Sedimentary Research, 78(3–4): 187198.CrossRefGoogle Scholar
Chen, B., Joachimski, M. M., Shen, S.-Z., Lambert, L. L., Lai, X.-L., Wang, X.-D., Chen, J., and Yuan, D.-X. in press. Permian ice volume and palaeoclimate history: oxygen isotope proxies revisited. Gondwana Research.Google Scholar
Cochran, J. K., Kallenberg, K., Landman, N. H., Harries, P. J., Weinreb, D., Turekian, K. K., Beck, A. J., and Cobban, W. A. 2010. Effect of diagenesis on the Sr, O, and C isotope composition of Late Cretaceous mollusks from the Western Interior Seaway of North America. American Journal of Science, 310(2):6988.Google Scholar
Cochran, J. K., Landman, N. H., Turekian, K. K., Michard, A., and Schrag, D. P. 2003. Paleoceanography of the Late Cretaceous (Maastrichtian) Western Interior Seaway of North America: evidence from Sr and O isotopes. Palaeogeography Palaeoclimatology Palaeoecology, 191(1):4564.Google Scholar
Compston, W. 1960. The carbon isotopic composition of certain marine invertebrates and coals from the Australian Permian. Geochimica et Cosmochimica Acta, 18:122.CrossRefGoogle Scholar
Coplen, T. B. 1988. Normalization of oxygen and hydrogen isotope data. Chemical Geology. (Isotope Geoscience Section), 72:293297.CrossRefGoogle Scholar
Coplen, T. B., De Bièvre, P., Krouse, H.R., Vocke, R.D. Jr., Gröning, M., and Rozanski, K. 1996. Ratios for light-element isotopes standardized for better interlaboratory comparison. Eos Trans. AGU, 77(27):255.Google Scholar
Corfield, R. M. 1995. An introduction to the techniques, limitation and landmarks of carbonate oxygen isotope palaeothermometry, p. 2742 In Bosence, D. W. and Allison, P. A. (eds.), Marine Palaeoenvironmental Analysis from Fossils. Geological Society Special Publication 83.Google Scholar
Craig, H. 1961. Standard for reporting concentrations of deuterium and oxygen-18 in natural waters. Science, 133:18331834.Google Scholar
Craig, H. 1965. Measurement of oxygen isotope paleotemperatures, p. 162182 In Tongiorgi, E. (ed.), Stable Isotopes in Oceanographic Studies and Paleotemperatures. Cons. Naz. Delle Ric., Spoleto, Italy.Google Scholar
Craig, H., and Gordon, L. I. 1965. Deuterium and oxygen 18 variation in the ocean and the marine atmosphere. In Torgiorgi, R. (ed.), Second conference on stable isotopes in oceanographic studies and paleotemperatures: Consiglio Nazionale delle Richerche, Pisa, 9130.Google Scholar
Degens, E. T., and Epstein, S. 1962. Relationship between O18/O16 ratios in coexisting carbonates, cherts, and diatomites. Bulletin of the American Association of Petroleum Geologists, 46:534542.Google Scholar
Dickson, J. A. D., Wood, R. A., and Kirkland, B. L. 1996. Exceptional preservation of the sponge Fissispongia tortacloaca from the Pennsylvanian Holder Formation, New Mexico. Palaios, 11(6):559570.CrossRefGoogle Scholar
Dimarco, S. F., Strauss, J., May, N., Mullins-Perry, R. L., Grossman, E. L., and Shormann, D. 2012. Texas coastal hypoxia linked to Brazos River discharge as revealed by oxygen isotopes. Aquatic Geochemistry, 18(2):159181.CrossRefGoogle Scholar
Dutton, A., Huber, B. T., Lohmann, K. C., and Zinsmeister, W. J. 2007. High-resolution stable isotope profiles of a dimitobelid belemnite: implications for paleodepth habitat and late Maastrichtian climate seasonality. Palaios, 22:642650.Google Scholar
Epstein, S., Buchsbaum, R., Lowenstam, H., and Urey, H. C. 1951. Carbonate-water isotopic temperature scale. Geological Society of America Bulletin, 62(4):417426.Google Scholar
Epstein, S., Buchsbaum, R., Lowenstam, H. A., and Urey, H. C. 1953. Revised carbonate-water isotopic temperature scale. Geological Society of America Bulletin, 64(11):13151325.Google Scholar
Epstein, S., and Lowenstam, H. A. 1953. Temperature-shell-growth relations of recent and interglacial Pleistocene shoal-water biota from Bermuda. Journal of Geology, 61(5):424438.Google Scholar
Epstein, S. and Mayeda, T. 1953. Variation of O18 content of waters from natural sources. Geochimica et Cosmochimica Acta, 4(5):213224.Google Scholar
Erez, J., and Luz, B. 1983. Experimental paleotemperatures equation for planktonic foraminifera. Geochimica et Cosmochimica Acta, 47(6):10251031.Google Scholar
Finnegan, S., Bergmann, K., Eiler, J. M., Jones, D. S., Fike, D. A., Eisenman, I., Hughes, N. C., Tripati, A. K., and Fischer, W. W. 2011. The magnitude and duration of Late Ordovician–Early Silurian glaciation. Science, 331:903906.Google Scholar
Fielding, C. R., Frank, T. D., and Isbell, J. L. 2008a. The late Paleozoic ice age—A review of current understanding and synthesis of global climate patterns, p. 343354 In Fielding, C. R., Frank, T. D., and Isbell, J. L. (eds.), Resolving the Late Paleozoic Ice Age in Time and Space. Geological Society of America Special Paper No. 441.Google Scholar
Fielding, C. R., Frank, T. D., Birgenheier, L. P., Rygel, M. C., Jones, A. T., and Roberts, J. 2008b. Stratigraphic imprint of the Late Palaeozoic Ice Age in eastern Australia: a record of alternating glacial and non-glacial climate regime. Journal of the Geological Society, 165:129140.CrossRefGoogle Scholar
Flake, R., 2011. Circulation of North American epicontinental seas during the Carboniferous using stable isotope and trace element analyses of brachiopod shells. , Texas A&M University, College Station, 63 p.Google Scholar
Friedman, I., and O'Neil, J. R. 1977. Data of Geochemistry, Sixth Edition, Chapter KK. Compilation of stable isotope fractionation factors of geochemical interest. U. S. Geological Survey Professional Paper 440-KK, U. S. Government Printing Office, Washington.Google Scholar
Gentry, D. K., Sosdian, S., Grossman, E. L., Rosenthal, Y., Hicks, D., and Lear, C. H. 2008. Stable isotope and Sr/Ca profiles from the marine gastropod Conus ermineus: Testing a multiproxy approach for inferring paleotemperature and paleosalinity. Palaios, 23(3–4): 195209.Google Scholar
GEOSECS, 1987, GEOSECS Atlantic, Pacific, and Indian Ocean expeditions, Vol. 7, Shorebased data and graphics, National Science Foundation, Washington, D.C., 200 p.Google Scholar
Ghosh, P., Adkins, J., Affek, H., Balta, B., Guo, W. F., Schauble, E. A., Schrag, D., and Eller, J. M. 2006. 13C-18O bonds in carbonate minerals: A new kind of paleothermometer. Geochimica et Cosmochimica Acta, 70(6):14391456.Google Scholar
Gonfiantini, R. 1984. Advisory group meeting on stable isotope reference samples for geochemical and hydrological investigations. International Atomic Energy Agency, Vienna, 19–21 September 1983, Report to the Director General, 77 p.Google Scholar
Gonzalez, L. A., and Lohmann, K.C. 1985. Carbon and oxygen isotopic composition of Holocene reefal carbonates. Geology, 13(11):811814.Google Scholar
Gradstein, F. M., Ogg, J. G., and Smith, A. 2004. A Geologic Time Scale 2004, Cambridge University Press, 589 p.Google Scholar
Gradstein, F. M., Ogg, J. G., Schmitz, M., and Ogg, G. 2012. The Geologic Time Scale 2012, Elsevier.Google Scholar
Gregory, R. T. 1991. Oxygen isotope history of seawater revisited: Timescales for boundary event changes in the oxygen isotope composition of seawater, p. 6576 In Taylor, H. P. Jr., O'Neil, J. R., and Kaplan, I. R. (eds.), Stable isotope geochemistry: A tribute to Samuel Epstein. Special Publication No. 3. The Geochemical Society, San Antonio.Google Scholar
Gröning, M. 2004. International stable isotope reference materials. In: Handbook of Stable Isotope Analytical Techniques, Vol. 1, de Groot, P.A. (Ed.), Elsevier, Amsterdam, p. 874906 Google Scholar
Grossman, E. L. 1987. Stable isotopes in modern benthic foraminifera – a study of vital effect. Journal of Foraminiferal Research, 17(1):4861.Google Scholar
Grossman, E. L. 1994. The carbon and oxygen isotopic record during the evolution of Pangea: Carboniferous to Triassic. In Klein, G.D. (ed.), Special Paper 288, Pangea: Paleoclimate, Tectonics, and Sedimentation during Accretion, Zenith, and Breakup of a supercontinent, Geological Society of America, 207228.Google Scholar
Grossman, E. L. 2012. Ch. 10. Oxygen isotope stratigraphy. In Gradstein, F. M., Ogg, J. G., Schmitz, M., and Ogg, G. (eds.), The Geologic Time Scale 2012, Elsevier, 195220.Google Scholar
Grossman, E. L., Bruckschen, P., Mii, H-S., Chuvashov, B. I., Yancey, T. E., and Veizer, J. 2002. Carboniferous paleoclimate and global change: Isotopic evidence from the Russian Platform. In Carboniferous Stratigraphy and Paleogeography in Eurasia. Institute of Geology and Geochemistry, Russian Academy of Sciences, Urals Branch, Ekaterinburg, 6171.Google Scholar
Grossman, E. L., and Ku, T. L. 1986. Oxygen and carbon isotope fractionation in biogenic aragonite – temperature effects. Chemical Geology, 59(1):5974.Google Scholar
Grossman, E. L., Mii, H. S., and Yancey, T. E. 1993. Stable isotopes in Late Pennsylvanian brachiopods from the United States – Implications for Carboniferous paleoceanography. Geological Society of America Bulletin, 105(10):12841296.Google Scholar
Grossman, E. L., Mii, H. S., Zhang, C. L., and Yancey, T. E. 1996. Chemical variation in Pennsylvanian brachiopod shells: Diagenetic, taxonomic, microstructural, and seasonal effects. Journal of Sedimentary Research, 66:10111022.Google Scholar
Grossman, E. L., Yancey, T. E., Jones, T. E., Bruckschen, P., Chuvashov, B., Mazzullo, S. J., and Mii, H. S. 2008. Glaciation, aridification, and carbon sequestration in the Permo–Carboniferous: The isotopic record for low latitudes. Palaeogeography Palaeoclimatology Palaeoecology, 268:222233.Google Scholar
Grossman, E. L., Zhang, C. L., and Yancey, T. E. 1991. Stable isotope stratigraphy of brachiopods from Pennsylvanian shales in Texas. Geological Society of America Bulletin, 103(7):953965.Google Scholar
Hays, P. D., and Grossman, E. L. 1991. Oxygen isotopes in meteoric calcite cements as indicators of continental paleoclimate. Geology, 19(5):441444.Google Scholar
He, S., Kyser, T. K., and Caldwell, W. G. E. 2005. Paleoenvironment of the Western Interior Seaway inferred from δ18O and δ13C values of molluscs from the Cretaceous Bearpaw marine cyclothem. Palaeogeography Palaeoclimatology Palaeoecology, 217(1–2): 6785.Google Scholar
Henkes, G. A., Passey, B. H., Grossman, E. L., and Yancey, T. E. 2011. Clumped isotope geochemistry of Carboniferous brachiopods: early lessons from a novel paleothermometer. 17th International Congress on the Carboniferous and Permian, Perth, Australia.Google Scholar
Hicks, D. W., and Mcmahon, R. F. 2002. Temperature acclimation of upper and lower thermal limits and freeze resistance in the nonindigenous brown mussel, Perna perna (L.), from the Gulf of Mexico. Marine Biology, 140(6):11671179.Google Scholar
Holmden, C. E., Creaser, R. A., Muehlenbachs, K., Leslie, S. A., and Bergstrom, S. M. 1998. Isotopic evidence for geochemical decoupling between ancient epeiric seas and bordering oceans. implications for secular curves. Geology, 26:567570.Google Scholar
Horibe, Y., and Oba, T. 1972. Temperature scales of aragonite-water and calcite-water systems. Fossils, 23–24:6978.Google Scholar
Hudson, J. D., and Anderson, T. F. 1989. Ocean temperatures and isotopic compositions through time. Transactions of the Royal Society of Edinburgh: Earth Science, 80:183192.CrossRefGoogle Scholar
Hut, G. 1987. Consultants group meeting on stable isotope reference samples for geochemical and hydrological investigations, Report to Director General, International. Atomic Energy Agency, Vienna, 42 p.Google Scholar
Isbell, J. L., Miller, M. F., Wolfe, K. L., and Lenaker, P. A. 2003. Timing of late Paleozoic glaciation in Gondwana: Was glaciation responsible for the development of northern hemisphere cyclothems? In Chan, M. A., and Archer, A. W. (eds.), Extreme depositional environments: Mega end members in geologic time. Boulder, Colorado, Geological Society of America Special Paper 370, 524.Google Scholar
Ivany, L. C. 2012. Reconstructing paleoseasonality from accretionary skeletal carbonates. In Ivany, L. C. and Huber, B. T. (eds.), Reconstructing Earth's Deep-Time Climate—The State of the Art in 2012. Paleontological Society Papers. Paleontological Society.Google Scholar
Ivany, L. C., and Runnegar, B. 2010. Early Permian seasonality from bivalve δ18O and implications for the oxygen isotopic composition of sea-water. Geology, 38(11):10271030.Google Scholar
Jaffrés, J. B. D., Shields, G. A., and Wallmann, K. 2007. The oxygen isotope evolution of seawater: A critical review of a long-standing controversy and an improved geological water cycle model for the past 3.4 billion years. Earth-Science Reviews, 83(1–2): 83122.Google Scholar
Jenkyns, H. C., Jones, C. E., Grocke, D. R., Hesselbo, S. P., and Parkinson, D. N. 2002. Chemostratigraphy of the Jurassic System: applications, limitations and implications for palaeoceanography. Journal of the Geological Society, 159:351378.Google Scholar
Jimenez-Lopez, C., Romanek, C. S., Huertas, F. J., Ohmoto, H., and Caballero, E. 2004. Oxygen isotope fractionation in synthetic magnesian calcite. Geochimica et Cosmochimica Acta, 68(16):33673377.Google Scholar
Joachimski, M.M., Breisig, S., Buggisch, W., Talent, J. A., Mawson, R., Gereke, M., Morrow, J. R., Day, J., and Weddige, K. 2009. Devonian climate and reef evolution: insights from oxygen isotopes in apatite. Earth and Planetary Science Letters, 284:599609.CrossRefGoogle Scholar
Joachimski, M. M., Lai, X., Shen, S., Jiang, H., Luo, G., Chen, B., Chen, J., and Sun, Y. 2012. Climate warming in the latest Permian and the Permian–Triassic mass extinction. Geology, 40(3):195198.Google Scholar
Joachimski, M. M., Van Geldern, R., Breisig, S., Buggisch, W., and Day, J. 2004. Oxygen isotope evolution of biogenic calcite and apatite during the Middle and Late Devonian. International Journal of Earth Sciences, 93(4):542553.Google Scholar
Joachimski, M. M., Von Bitter, P. H., and Buggisch, W. 2006. Constraints on Pennsylvanian glacioeustatic sea-level changes using oxygen isotopes of conodont apatite. Geology, 34(4):277280.Google Scholar
Johnson, K.S. 1989. Evaporite deposits in Carboniferous rocks of the U.S.A. XI Congres International de Stratigraphie et de Geologie du Carbonifere, Beiing 1987, Compte Rendu 4:5165.Google Scholar
Kim, S. T., and O'Neil, J. R. 1997. Equilibrium and nonequilibrium oxygen isotope effects in synthetic carbonates. Geochimica et Cosmochimica Acta, 61(16):34613475.Google Scholar
Kim, S. T., O'Neil, J. R., Hillaire-Marcel, C., and Mucci, A. 2007. Oxygen isotope fractionation between synthetic aragonite and water: Influence of temperature and Mg2+ concentration. Geochimica et Cosmochimica Acta, 71(19):47044715.Google Scholar
Knauth, L.P., and Epstein, S. 1976. Hydrogen and oxygen isotope ratios in nodular and bedded cherts. Geochimica et Cosmochimica Acta, 40(9):10951108.Google Scholar
Knauth, L. P., and Roberts, S. K. 1991. The hydrogen and oxygen isotope history of the Silurian-Permian hydrosphere as determined by direct measurement of fossil water, p. 91104 In Taylor, H. P. Jr., O'Neil, J. R., and Kaplan, I. R. (eds.), Stable isotope geochemistry: A tribute to Samuel Epstein. Special Publication No. 3. The Geochemical Society, San Antonio.Google Scholar
Kobashi, T., Grossman, E. L., Dockery, D. T., and Ivany, L. C. 2004. Water mass stability reconstructions from greenhouse (Eocene) to icehouse (Oligocene) for the northern Gulf Coast continental shelf (USA). Paleoceanography, 19(1).Google Scholar
Kolodny, Y., Luz, B., and Navon, O. 1983. Oxygen isotope variations in phosphate of biogenic apatites. 1. Fish bone apatite – rechecking the rules of the game. Earth and Planetary Science Letters, 64(3):398404.Google Scholar
Korte, C., Jasper, T., Kozur, H. W., and Veizer, J. 2005a. δ18O and δ13C of Permian brachiopods: A record of seawater evolution and continental glaciation. Palaeogeography Palaeoclimatology Palaeoecology, 224(4):333351.Google Scholar
Korte, C., Jones, P. J., Brand, Z.U., Mertmann, D., and Veizer, J. 2008. Oxygen isotope values from high-latitudes: Clues for Permian sea-surface temperature gradients and Late Palaeozoic deglaciation. Palaeogeography Palaeoclimatology Palaeoecology, 269(1–2): 116.Google Scholar
Korte, C., Kozur, H. W., and Veizer, J. 2005b. δ13C and δ18O values of Triassic brachiopods and carbonate rocks as proxies for coeval seawater and palaeotemperature. Palaeogeography Palaeoclimatology Palaeoecology, 226(3–4): 287306.Google Scholar
Land, L. S. 1995. Oxygen and carbon isotopic composition of Ordovician brachiopods—implication for coeval seawater—Comment. Geochimica et Cosmochimica Acta, 59(13):28432844.Google Scholar
Lécuyer, C., and Allemand, P. 1999. Modeling of the oxygen isotope evolution of seawater: Implications for the climate interpretation of the δ18O of marine sediments. Geochimica et Cosmochimica Acta, 63(3–4): 351361.Google Scholar
Lécuyer, C., Grandjean, P., and Emig, C. C., 1996. Determination of oxygen isotope fractionation between water and phosphate from living lingulids: potential application to palaeoenvironmental studies. Palaeogeography Palaeoclimatology Palaeoecology, 126:101108.Google Scholar
Leder, J. J., Swart, P. K., Szmant, A. M., and Dodge, R. E. 1996. The origin of variations in the isotopic record of scleractinian corals: 1. Oxygen. Geochimica et Cosmochimica Acta, 60(15):28572870.Google Scholar
Lee, R. W. 2003. Thermal tolerances of deep-sea hydrothermal vent animals from the Northeast Pacific. Biological Bulletin, 205(2):98101.Google Scholar
Lee, X-Q., and Wan, G-J. 2000. No vital effect on δ18O and δ13C values of fossil brachiopod shells, Middle Devonian of China. Geochimica et Cosmochimica Acta, 64:26492664.Google Scholar
Legrande, A. N., and Schmidt, G. A. 2006. Global gridded data set of the oxygen isotopic composition in seawater. Geophysical Research Letters, 33(12), L12604, doi:10.1029/2006GL026011.Google Scholar
Lepzelter, C. G., Anderson, T. F., and Sandberg, P. A. 1983. Stable isotope variation in modern articulate brachiopods (abs.): American Association of Petroleum Geologists Bulletin, 67:500501.Google Scholar
Lhomme, N., Clarke, G. K. C., and Ritz, C. 2005. Global budget of water isotopes inferred from polar ice sheets. Geophysical Research Letters, 32(20).Google Scholar
Lloyd, R. M. 1964. Variations in the oxygen and carbon isotope ratios of Florida Bay mollusks and their environmental significance. Journal of Geology, 72(1):84111.Google Scholar
Longinelli, A., and Nuti, S. 1973. Revised phosphate-water isotopic temperature scale. Earth and Planetary Science Letters, 19(3):373376.Google Scholar
Lowenstam, H. A., 1961. Mineralogy, O18/O16 ratios, and strontium and magnesium contents of recent and fossil brachiopods and their bearing on the history of the oceans. Journal of Geology 69:241260.Google Scholar
Luz, B., Kolodny, Y., and Kovach, J. 1984. Oxygen isotope variations in phosphate of biogenic apatites. 3. Conodonts. Earth and Planetary Science Letters, 69(2):255262.Google Scholar
Lynch-Stieglitz, J., Curry, W. B., and Slowey, N. 1999. A geostrophic transport estimate for the Florida Current from the oxygen isotope composition of benthic foraminifera. Paleoceanography, 14(3):360373.CrossRefGoogle Scholar
Machel, H. G. 1985. Cathodoluminescence in calcite and dolomite and its chemical interpretation. Geoscience Canada. 12(4):139147.Google Scholar
Macleod, K. G. 2012. The δ18O paleothermometer applied to phosphate oxygen measurements of bioapatite. In Ivany, L. C. and Huber, B. T. (eds.), Reconstructing Earth's Deep-Time Climate—The State of the Art in 2012. Paleontological Society Papers. Paleontological Society.Google Scholar
Malchus, N. and Steuber, T. 2002. Stable isotope records (O, C) of Jurassic aragonitic shells from England and NW Poland: palaeoecologic and environmental implications. Geobios, 35(1):2939.Google Scholar
Marshall, J. D. 1992, Climatic and oceanographic isotopic signals from the carbonate rock record and their preservation: Geological Magazine, 129:143160.Google Scholar
Marshall, J. D., Brenchley, P. J., Mason, P., Wolff, G. A., Astini, R. A., Hints, L., and Meidla, T. 1997. Global carbon isotopic events associated with mass extinction and glaciation in the late Ordovician. Palaeogeography Palaeoclimatology Palaeoecology, 132(1–4): 195210.Google Scholar
Marshall, J. D., and Middleton, P. D. 1990. Changes in marine isotopic composition and late Ordovician glaciation. Journal of the Geological Society, London, 147:14.Google Scholar
Mason, R. A. 1987. Ion microprobe analysis of trace-elements in calcite with an application to the cathodoluminescence zonation of limestone cements from the Lower Carboniferous of South-Wales, UK. Chemical Geology, 64(3–4): 209224.Google Scholar
Mazzullo, S. J., Boardman, D. R., Grossman, E. L., and Dimmick-Wells, K. 2007. Oxygen-carbon isotope stratigraphy of Upper Pennsylvanian to Lower Permian marine deposits in Kansas and northeastern Oklahoma: implications for seawater isotopic composition, glaciation, and depositional cyclicity. Carbonates and Evaporites, 22:5572.Google Scholar
Mccrea, J. M. 1950. On the isotopic chemistry of carbonates and a paleotemperature scale. Journal of Chemical Physics, 18:849857.Google Scholar
Mii, H. S., and Grossman, E. L. 1994. Late Pennsylvanian seasonality reflected in the 18O and elemental composition of a brachiopod shell. Geology, 22:661664.Google Scholar
Mii, H. S., Grossman, E. L., and Yancey, T. E. 1999. Carboniferous isotope stratigraphies of North America: Implications for Carboniferous paleoceanography and Mississippian glaciation. Geological Society of America Bulletin, 111(7):960973.Google Scholar
Mii, H. S., Grossman, E. L., Yancey, T. E., Chuvashov, B., and Egorov, A. 2001. Isotopic records of brachiopod shells from the Russian Platform—evidence for the onset of mid-Carboniferous glaciation. Chemical Geology, 175(1–2): 133147.Google Scholar
Mii, H.-S., Shi, G. R., Cheng, C.-J., and Chen, Y.-Y. 2012. Permian Gondwanaland paleoenvironment inferred from carbon and oxygen isotope records of brachiopod fossils from Sydney Basin, southeast Australia. Chemical Geology, 291:87103.Google Scholar
Moriya, K., Nishi, H., Kawahata, H., Tanabe, K., and Takayanagi, Y. 2003. Demersal habitat of Late Cretaceous ammonoids: Evidence from oxygen isotopes for the Campanian (Late Cretaceous) northwestern Pacific thermal structure. Geology, 31:167170.Google Scholar
Muehlenbachs, K. 1986. Ch. 12. Alteration of the oceanic crust and the 18O history of seawater, p. 425444 In Valley, J. W., Taylor, H. P. Jr., and O'Neil, J. R. (eds.), Stable isotopes in high temperature geological processes. Reviews in Mineralogy, Volume 16.Google Scholar
Muehlenbachs, K. 1998. The oxygen isotopic composition of the oceans, sediments and the seafloor. Chemical Geology, 145(3–4): 263273.Google Scholar
Muehlenbachs, K., and Clayton, R. N. 1976. Oxygen isotope composition of oceanic-crust and its bearing on seawater. Journal of Geophysical Research, 81(23):43654369.Google Scholar
Müller-Lupp, T., and Bauch, H. 2005. Linkage of Arctic atmospheric circulation and Siberian shelf hydrography: A proxy validation using δ18O records of bivalve shells. Global and Planetary Change, 48(1–3): 175186.CrossRefGoogle Scholar
Mutterlose, J., Pauly, S., and Steuber, T. 2009. Temperature controlled deposition of early Cretaceous (Barremian–early Aptian) black shales in an epicontinental sea. Palaeogeography Palaeoclimatology Palaeoecology, 273:330345.Google Scholar
Nützel, A., Joachimski, M., and López Correa, M. 2010. Seasonal climatic fluctuations in the Late Triassic tropics—High-resolution oxygen isotope records from aragonitic bivalve shells (Cassian Formation, Northern Italy). Palaeogeography Palaeoclimatology Palaeoecology, 285:194204.Google Scholar
O'Neil, J. R., Clayton, R. N., and Mayeda, T. K. 1969. Oxygen isotope fractionation in divalent metal carbonates. Journal of Chemical Physics, 51(12):55475558.Google Scholar
Noret, J. R., Grossman, E. L., Yancey, T. E., and Chuvashov, B. I. 2009. Global climatic and ecological correlations during the Early Permian (Cisuralian). South-Central GSA Abstracts with Programs.Google Scholar
Panchuk, K.M., Holmden, C. E., and Leslie, S. A. 2006. Local controls on carbon cycling in the Ordovician mid-continent region of North America with implications for carbon isotope secular curves. Journal of Sedimentary Research 76 DOI:10.2110/jsr.2006.017.Google Scholar
Passey, B., Henkes, G., Grossman, E., and Yancey, T. 2011. Deep time paleoclimate reconstruction using carbonate clumped isotope thermometry: a status report. 17th International Congress on the Carboniferous and Permian (Perth, Australia).Google Scholar
Pearson, P. N. 2012. The Marine Stable Isotope Record. In Ivany, L. C. and Huber, B. T. (eds.), Reconstructing Earth's Deep-Time Climate—The State of the Art in 2012. Paleontological Society Papers. Paleontological Society.Google Scholar
Pearson, P. N., Ditchfield, P. W., Singano, J., Harcourt-Brown, K. G., Nicholas, C. J., Olsson, R. K., Shackleton, N. J., and Hall, M. A. 2001. Warm tropical sea surface temperatures in the Late Cretaceous and Eocene epochs. Nature, 413(6855):481487.Google Scholar
Pérez-Huerta, A., Cusack, M., and England, J. 2007. Crystallography and diagenesis in fossil craniid brachiopods. Palaeontology, 50:757763.Google Scholar
Perry, E. C. 1967. Oxygen isotope chemistry of ancient cherts. Earth and Planetary Science Letters, 3:6266.Google Scholar
Podlaha, O. G., Mutterlose, J., and Veizer, J. 1998. Preservation of δ18O and δ13C in belemnite rostra from the Jurassic Early Cretaceous successions. American Journal of Science, 298(4):324347.Google Scholar
Popp, B. N. 1986. The record of carbon, oxygen, sulfur, and strontium isotopes and trace elements in late Paleozoic brachiopods []. Urbana, Illinois, University of Illinois, 199 p.Google Scholar
Popp, B. N., Anderson, T. F., and Sandberg, P. A. 1986. Brachiopods as indicators of original isotopic compositions in some Paleozoic limestones. Geological Society of America Bulletin, 97(10):12621269.Google Scholar
Pörtner, H. O. 2002. Climate variations and the physiological basis of temperature dependent bio-geography: systemic to molecular hierarchy of thermal tolerance in animals. Comparative Biochemistry and Physiology a—Molecular and Integrative Physiology, 132(4):739761.Google Scholar
Prokoph, A., Shields, G. A., and Veizer, J. 2008. Compilation and time-series analysis of a marine carbonate δ18O, δ13C, 87Sr/86Sr and δ34S database through Earth history. Earth-Science Reviews, 87(3–4): 113133.Google Scholar
Pucéat, E., Joachimski, M. M., Bouilloux, A., Monna, F., Bonin, A., Motreuil, S., Moriniere, P., Henard, S., Mourin, J., Dera, G., and Quesne, D. 2010. Revised phosphate-water fractionation equation reassessing paleotemperatures derived from biogenic apatite. Earth and Planetary Science Letters, 298:135142.Google Scholar
Qing, H.R. and Veizer, J., 1994. Oxygen and carbon isotopic composition of Ordovician brachiopods – Implications for coeval seawater. Geochimica et Cosmochimica Acta, 58(20):44294442.Google Scholar
Railsback, L. B., Anderson, T. F., Ackerly, S. C., and Cisne, J. L. 1989. Paleoceanographic modeling of temperature-salinity profiles from stable isotope data: Paleoceanography, 4:585591.Google Scholar
Rozanski, K., Araguás-Araguás, L., and Gonfiantini, R. 1993. Isotopic patterns in modern global precipitation, p. 136 In Swart, P. K., Lohmann, K. C., McKenzie, J., and Savin, S. (eds.), Climate Change in Continental Isotopic Records. Geophysical Monograph 78. American Geophysical Union, Washington.Google Scholar
Rush, P. F., and Chafetz, H. S. 1990. Fabric-retentive, non-luminescent brachiopods as indicators of original δ13C and δ18O composition: a test. Journal of Sedimentary Petrology, 60:968981.Google Scholar
Samtleben, C., Munnecke, A., Bickert, T. and Pätzold, J. 1996. The Silurian of Gotland (Sweden): Facies interpretation based on stable isotopes in brachiopod shells. Geologische Rundschau, 85(2):278292.Google Scholar
Savard, M. M., Veizer, J., and Hinton, R. 1995. Cathodoluminescence at low Fe and Mn concentrations—a SIMS study of zones of natural calcites. Journal of Sedimentary Research Section a—Sedimentary Petrology and Processes 65(1):208213.Google Scholar
Savin, S. M. 1977. The history of the Earth's surface temperature during the past 100 million years. Annual Review of Earth and Planetary Sciences, 5:319355.Google Scholar
Schmidt, G. A. 1999. Forward modeling of carbonate proxy data from planktonic foraminifera using oxygen isotope tracers in a global ocean model. Paleoceanography, 14(4):482497.Google Scholar
Schmidt, G.A., Bigg, G.R., and Rohling, E.J. 1999. Global Seawater Oxygen-18 Database. http://data.giss.nasa.gov/o18data/.Google Scholar
Schrag, D.P. 1999. Effects of diagenesis on the isotopic record of late Paleogene tropical sea surface temperatures. Chemical Geology, 161(1–3): 215224.Google Scholar
Seuss, B., Titschack, J., Seifert, S., Neubauer, J., and Nützel, A. 2012. Oxygen and stable carbon isotopes from a nautiloid from the middle Pennsylvanian (Late Carboniferous) impregnation Lagerstätte ‘Buckhorn Asphalt Quarry’ — Primary paleo-environmental signals versus diagenesis. Palaeogeography, Palaeoclimatology, Palaeoecology, 319–320:115.Google Scholar
Shackleton, N. J. 1974. Attainment of isotopic equilibrium between ocean water and the benthonic foraminifera genus Uvigerina: isotopic changes in the ocean during the last glacial. Centre National de la Recherche Scientifique Colloq. Internationau, 219:203209.Google Scholar
Sharp, Z. 2007. Principles of Stable Isotope Geochemistry. Pearson, Upper Saddle River, NJ, 344 p.Google Scholar
Spero, H. J., Bijma, J., Lea, D. W., and Bemis, B. E. 1997. Effect of seawater carbonate concentration on foraminiferal carbon and oxygen isotopes. Nature, 390(6659):497500.Google Scholar
Stahl, W., and Jordan, R. 1969. General considerations on isotopic paleotemperature determinations and analyses on Jurassic ammonites. Earth and Planetary Science Letters, 6:173178.Google Scholar
Swart, P.K., Moore, M., Charles, C., and Bohm, F. 1998. Sclerosponges may hold new keys to marine paleoclimate. Eos, 633, 636.Google Scholar
Swart, P. K., and Price, R. 2002. Origin of salinity variations in Florida Bay. Limnology and Oceanography, 47:12341241.CrossRefGoogle Scholar
Tarutani, T., Clayton, R. N., and Mayeda, T. K. 1969. The effect of polymorphism and magnesium substitution on oxygen isotope fractionation between calcium carbonate and water. Geochimica et Cosmochimica Acta, 33:987996.Google Scholar
Trotter, J. A., Williams, I. S., Barnes, C. R., Lecuyer, C., and Nicoll, R. S. 2008. Did cooling oceans trigger Ordovician biodiversification? Evidence from conodont thermometry. Science, 321(5888):550554.CrossRefGoogle ScholarPubMed
Urey, H. C. 1947. The thermodynamic properties of isotopic substances. Journal of the Chemical Society of London, 1947:562581.Google Scholar
Urey, H. C., Lowenstam, H. A., Epstein, S., and Mckinney, C. R. 1951. Measurement of paleotemperatures and temperatures of the upper Cretaceous of England, Denmark, and the southeastern United States. Geological Society of America Bulletin, 62:399416.Google Scholar
Van De Schootbrugge, B., Follmi, K. B., Bulot, L. G., and Burns, S. J. 2000. Paleoceanographic changes during the early Cretaceous (Valanginian–Hauterivian): evidence from oxygen and carbon stable isotopes. Earth and Planetary Science Letters, 181(1–2): 1531.Google Scholar
Van Geldern, R., Joachimski, M. M., Day, J., Jansen, U., Alvarez, F., Yolkin, E. A., and Ma, X. P. 2006. Carbon, oxygen and strontium isotope records of Devonian brachiopod shell calcite. Palaeogeography Palaeoclimatology Palaeoecology, 240(1–2): 4767.Google Scholar
Veizer, J. 1983. Ch. 3. Chemical diagenesis of carbonates: Theory and application of trace element technique. In Stable Isotopes in Sedimentary Geology, SEPM Short Course No. 10:3–1 to 3–100.Google Scholar
Veizer, J. 1992. Depositional and diagenetic histroy of limestones: Stable and radiogenic isotopes, p. 1348 In Clauer, N. and Chaudhuri, S. (ed.), Isotopic Signatures and Sedimentary Rocks. Springer-Verlag, Berlin.Google Scholar
Veizer, J. 1995. Oxygen and carbon isotopic composition of Ordovician brachiopods—implication for coeval seawater—Reply. Geochimica et Cosmochimica Acta, 59(13):28452846.Google Scholar
Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Buhl, D., Bruhn, F., Carden, G. A. F., Diener, A., Ebneth, S., Godderis, Y., Jasper, T., Korte, C., Pawellek, F., Podlaha, O. G., and Strauss, H. 1999. 87Sr/86Sr, δ13C and δ18O evolution of Phanerozoic seawater. Chemical Geology, 161(1–3): 5988.Google Scholar
Veizer, J., Bruckschen, P., Pawellek, F., Diener, A., Podlaha, O. G., Carden, G. A. F., Jasper, T., Korte, C., Strauss, H., Azmy, K., and Ala, D. 1997. Oxygen isotope evolution of Phanerozoic seawater. Palaeogeography Palaeoclimatology Palaeoecology, 132(1–4): 159172.Google Scholar
Veizer, J., Fritz, P., and Jones, B. 1986. Geochemistry of brachiopods – oxygen and carbon isotopic records of Paleozoic oceans. Geochimica et Cosmochimica Acta, 50(8):16791696.CrossRefGoogle Scholar
Veizer, J., and Hoefs, J. 1976. The nature of O18/O16 and C13/C12 secular trends in sedimentary carbonate rocks. Geochimica et Cosmochimica Acta, 40:13871395.Google Scholar
Vennemann, T.W., Fricke, H. C., Blake, R.E., O'Neil, J. R., and Colman, A. 2002. Oxygen isotope analysis of phosphates: a comparison of techniques for analysis of Ag3PO4 . Chemical Geology, 185(3–4): 321336.Google Scholar
Wallmann, K. 2001. The geological water cycle and the evolution of marine δ18O values. Geochimica et Cosmochimica Acta, 65(15):24692485.Google Scholar
Ward, W. B., and Reeder, R. J. 1993. The use of growth microfabrics and transmission electron microscopy in understanding replacement processes in carbonates, p. 253264 In Rezak, R. and Lavoie, D. L. (eds.), Carbonate Microfacies. Springer-Verlag, New York.Google Scholar
Wefer, G., and Berger, W. H. 1991. Isotope paleontology – growth and composition of extant calcareous species. Marine Geology, 100(1–4): 207248.Google Scholar
Wenzel, B., and Joachimski, M. M. 1996. Carbon and oxygen isotopic composition of Silurian brachiopods (Gotland/Sweden): Palaeoceanographic implications. Palaeogeography Palaeoclimatology Palaeoecology, 122(1–4): 143166.Google Scholar
Wenzel, B., Lecuyer, C., and Joachimski, M. M. 2000. Comparing oxygen isotope records of Silurian calcite and phosphate - δ18O compositions of brachiopods and conodonts. Geochimica et Cosmochimica Acta, 64(11):18591872.Google Scholar
Wierzbowski, H., and Joachimski, M. 2007. Reconstruction of late Bajocian–Bathonian marine palaeoenvironments using carbon and oxygen isotope ratios of calcareous fossils from the Polish Jura Chain (central Poland). Palaeogeography Palaeoclimatology Palaeoecology, 254(3–4): 523540.Google Scholar
Wierzbowski, H., and Joachimski, M. M. 2009. Stable isotopes, elemental distribution, and growth rings of belemnopsid belemnite rostra: Proxies for belemnite life habitat. Palaios, 25(5–6): 377386.Google Scholar
Wright, E.K. 1987. Stratification and paleocirculation of the Late Cretaceous Western Interior Seaway of North America. Geological Society of America Bulletin, 99(4):480490.Google Scholar
Zachos, J. C., Stott, L. D., and Lohmann, K. C. 1994. Evolution of early Cenozoic marine temperatures. Paleoceanography, 9(2):353387.Google Scholar
Zeebe, R. E. 1999. An explanation of the effect of seawater carbonate concentration on foraminiferal oxygen isotopes. Geochimica et Cosmochimica Acta, 63(13–14): 20012007.Google Scholar