Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-25T05:04:47.486Z Has data issue: false hasContentIssue false

Isotopic Determination of Growth and Longevity in Fossil and Modern Invertebrates

Published online by Cambridge University Press:  21 July 2017

Douglas S. Jones*
Affiliation:
Florida Museum of Natural History, University of Florida, Gainesville, FL 32611, USA
Get access

Extract

The isotopic composition of fossil invertebrates contains a wealth of information about the physical and chemical environments of the ancient past. The exploitation of this biogeochemical archive by paleontologists began about 50 years ago with the realization that the ratios of oxygen isotopes in the shells and skeletons of marine organisms offered the potential to accurately reconstruct paleoenvironmental conditions, particularly temperature (Urey, 1947; Urey et al., 1951). With the introduction of the oxygen isotope paleotemperature methodology (Epstein et al., 1951, 1953; Epstein and Lowenstam, 1953), the field of “isotope paleontology” was born (Wefer and Berger, 1991).

Type
Research Article
Copyright
Copyright © 1998 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

Abell, P. I. 1985. Oxygen isotope ratios in modern African gastropod shells: a data base for paleoclimatology. Chemical Geology (Isotope Geoscience Section), 58:183193.Google Scholar
Adkins, J. F., Cheng, H., Boyle, E. A., Druffel, E. R. M., and Edwards, R. L. 1998. Deep-sea coral evidence for rapid change in ventilation of the deep North Atlantic 15,400 years ago. Science, 280:725728.Google Scholar
Allmon, W. D., Jones, D. S., Aiello, R. L., Gowlett-Holmes, K., and Probert, P. K. 1994. Observations on the biology of Maoricolpus roseus (Quoy and Gaimard) (Prosobranchia: Turritellidae) from New Zealand and Tasmania. The Veliger, 37:267279.Google Scholar
Allmon, W. D., Jones, D. S., Polizzotto, K. 1997. Evolution and environment in a time of global change: Turritella cingulata-cingulatiformis (Gastropoda), Pliocene-Recent of Chile. Geological Society of America, Annual Meeting Abstracts with Programs, 29(6):A-405.Google Scholar
Allmon, W. D., Jones, D. S., and Vaughan, N. 1992. Observations on the biology of Turritella gonostoma Valenciennes (Prosobranchia: Turritellidae) from the Gulf of California. The Veliger, 35:5263.Google Scholar
Anderson, T. F., and Arthur, M. A. 1983. Stable isotopes of oxygen and carbon and their application to sedimentologic and paleoenvironmental problems, p. 1-1-1-151. In Arthur, M. A. (organizer), Stable Isotopes in Sedimentary Geology: SEPM Short Course No. 10. Society of Economic Paleontologists and Mineralogists, Tulsa, Oklahoma.Google Scholar
Andreasson, F. P., and Schmitz, B. 1996. Winter and summer temperatures of the early middle Eocene of France from Turritella δ18O profiles. Geology, 24:10671070.Google Scholar
Appledoorn, R. S. 1988. Age determination, growth, mortality and age of first reproduction in adult queen conch, Strombus gigas L., off Puerto Rico. Fisheries Research, 6:363378.CrossRefGoogle Scholar
Arthur, M. A., Williams, D. F., and Jones, D. S. 1983. Seasonal temperature-salinity changes and thermocline development in the mid-Atlantic Bight as recorded by the isotopic composition of bivalves. Geology, 11:655659.Google Scholar
Bailey, G. N., Deith, M. R., and Shackleton, N. J. 1983. Oxygen isotope analysis and seasonality determinations: Limits and potentials of a new technique. American Antiquity, 48:390398.Google Scholar
Barrera, E., Tevesz, M. J. S., Carter, J. G., and McCall, P. L. 1994. Oxygen and carbon isotopic composition and shell microstructure of the bivalve Laternula elliptica from Antarctica. Palaios, 9:275287.Google Scholar
Bemis, B. E., and Geary, D. H. 1996. The usefulness of bivalve stable isotope profiles as environmental indicators: Data from the eastern Pacific Ocean and the southern Caribbean Sea. Palaios, 11:328339.Google Scholar
Benavides, L. M., and Druffel, E. M. 1986. Sclerosponge growth rate as determined by 210Pb and δ14C chronologies. Coral Reefs, 4:221224.Google Scholar
Bonham, K. 1965. Growth rate of giant clam Tridacna gigas at Bikini Atoll as revealed by radioautography. Science, 149:300302.Google Scholar
Brand, U. 1989. Biogeochemistry of late Paleozoic North American brachiopods and secular variation of seawater composition. Biogeochemistry, 7:159193.Google Scholar
Buchardt, B., and Weiner, S. 1981. Diagenesis of aragonite from Upper Cretaceous ammonites: A geochemical case-study. Sedimentology, 28:423438.Google Scholar
Bucher, H., Landman, N. H., Klofak, S. M. and Guex, J. 1996. Mode and rate of growth in ammonoids, p. 407461. In Landman, N. H., Tanabe, K., and Davis, R. A. (eds.), Ammonoid Paleobiology. Plenum Press, New York.CrossRefGoogle Scholar
Buddemeier, R. W., Maragos, J. E., and Knutson, D. W. 1974. Radiographic studies of reef coral skeletons: rates and patterns of coral growth. Journal of Marine Biology and Ecology, 14:179200.CrossRefGoogle 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
Charles, C. D., Hunter, D. E., and Fairbanks, R. G. 1997. Interaction between the ENSO and the Asian Monsoon in a coral record of tropical climate. Science, 277:925928.CrossRefGoogle Scholar
Chinzei, K., Koike, H., Oba, T., Matsushima, Y., and Kitazato, H. 1987. Secular changes in the oxygen isotope ratios of mollusc shells during the Holocene of Central Japan. Palaeogeography, Palaeoclimatology, Palaeoecology, 61:155166.Google Scholar
Cochran, J. K., and Landman, N. H. 1984. Radiometric determination of the growth rate of Nautilus in nature. Nature, 308:725727.CrossRefGoogle Scholar
Cochran, J. K., Rye, D. M., and Landman, N. H. 1981. Growth rate and habitat of Nautilus pompilius inferred from radioactive and stable isotope studies. Paleobiology, 7:469480.Google Scholar
Cook, E. R. 1995. Temperature histories from tree rings and corals. Climate Dynamics, 11:211222.CrossRefGoogle Scholar
Cornu, S., Pätzold, J., Bard, E., Meco, J., and Cerda-Barcelo, J. 1993. Paleotemperature of the last interglacial period based on δ18O of Strombus bubonius from the western Mediterranean Sea. Palaeogeography, Palaeoclimatology, Palaeoecology, 103:120.CrossRefGoogle Scholar
Crocker, K. C., DeNiro, M. J., and Ward, P. D. 1985. Stable isotopic investigations of early development in extant and fossil chambered cephalopods. I. Oxygen isotopic composition of eggwater and carbon isotopic composition of siphuncle organic matter in Nautilus . Geochimica et Cosmochimica Acta, 49:25272532.Google Scholar
Deith, M. R. 1986. Subsistence strategies at a Mesolithic camp site: Evidence from stable isotope analyses of shells. Journal of Archaeological Science, 13:6178.Google Scholar
Dettman, D. L., and Lohmann, K. C. 1993. Seasonal change in Paleogene surface water δ18O: fresh-water bivalves of western North America, p. 153163. In Swart, P. K., Lohmann, K. C., McKenzie, J., and Savin, S. (eds.), Climate Change in Continental Records: AGU Geophysical Monograph Vol. 78. American Geophysical Union, Washington, D. C. Google Scholar
Dodge, R. E., and Thomson, J. 1974. The natural radiochemical and growth records in contemporary hermatypic corals from the Atlantic and Caribbean. Earth and Planetary Science Letters, 23:313322.CrossRefGoogle Scholar
Druffel, E. M. 1981. Radiocarbon in annual coral rings from the eastern tropical Pacific Ocean. Geophysical Research Letters, 8:5962.CrossRefGoogle Scholar
Druffel, E. M. 1982. Banded corals: changes in oceanic carbon-14 during the Little Ice Age. Science, 218:1319.Google Scholar
Druffel, E. M., and Benavides, L. M. 1986. Input of excess CO2 to the surface ocean based on 13C/12C ratios in a banded Jamaican sclerosponge. Nature, 321:5861.CrossRefGoogle Scholar
Druffel, E. M., King, L. L., Belastock, R. A., and Buessler, K. O. 1990. Growth rate of a deepsea coral using 210Pb and other isotopes. Geochimica et Cosmochimica Acta, 54:14931499.CrossRefGoogle Scholar
Druffel, E. M., and Linick, T. W. 1978. Radiocarbon in annual coral rings of Florida. Geophysical Research Letters, 5:913916.CrossRefGoogle Scholar
Dunbar, R. B., and Wellington, G. M. 1981. Stable isotopes in a branching coral monitor seasonal temperature variation. Nature, 293:453455.Google Scholar
Dunbar, R. B., and Wellington., G. M., Colgan, M. W., and Glynn, P. W. 1994. Eastern Pacific sea surface temperature since 1600 A.D.: The δ18O record of climate variability in Galapagos corals. Paleoceanography, 9:291315.Google Scholar
Eichler, R., and Ristedt, H. 1966. Isotopic evidence on the early life history of Nautilus pompilius (Linné). Science, 153:734736.Google Scholar
Emiliani, C., Hudson, J. H., Shinn, E. A., and George, R. Y. 1978. Oxygen and carbon isotopic growth record in a reef coral from the Florida Keys and a deep-sea coral from Blake Plateau. Science, 202:627629.Google Scholar
Epstein, S., Buchsbaum, R., Lowenstam, H. A., and Urey, H. C. 1951. Carbonate-water isotopic temperature scale. Geological Society of America Bulletin, 62:417426.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, 64:13151325.Google Scholar
Epstein, S., Buchsbaum, R., and Lowenstam, H. A. 1953. Temperature-shell-growth relations of recent and interglacial Pleistocene shoal-water biota from Bermuda. Journal of Geology, 61:424438.Google Scholar
Erez, J. and Luz, B. 1982. Temperature control of oxygen-isotope fractionation of cultured planktonic foraminifera. Nature, 297:220222.Google Scholar
Erez, J. and Luz, B. 1983. Experimental paleotemperature equation for planktonic foraminifera. Geochimica et Cosmochimica Acta, 47:10251031.Google Scholar
Erlenkeuser, H. 1976. 14C and 13C isotope concentration in modern marine mussels from sedimentary habitats. Naturwissenschaften, 63:338.Google Scholar
Erlenkeuser, H., and Wefer, G. 1981. Seasonal growth of bivalves from Bermuda recorded in their O-18 profiles. Proceedings of the Fourth International Coral Reef Symposium, Manila, 4(2):643648.Google Scholar
Fairbanks, R. G., and Dodge, R. E. 1979. Annual periodicity of the 18O/16O and 13C/12C ratios in the coral Montastrea annularis . Geochimica et Cosmochimica Acta, 43:10091020.Google Scholar
Fairbanks, R. G., Evans, M. N., Rubenstone, J. L., Mortlock, R. A., Broad, K., Moore, M. D., and Charles, C. D. 1997. Evaluating climate indices and their geochemical proxies measured in corals. Coral Reefs, 16(Suppl.):S93S100.CrossRefGoogle Scholar
Fastovsky, D. E., Arthur, M. A., Strater, N. H., and Foss, A. 1993. Freshwater bivalves (Unionidae), disequilibrium isotopic fractionation, and temperatures. Palaios, 8:602608.Google Scholar
Forester, R. M., Sandberg, P. A., and Anderson, T. F. 1973. Isotopic variability of cheilostome bryozoan skeletons, p. 7994. In Larwood, G. P. (ed.), Living and Fossil Bryozoa. Academic Press, London.Google Scholar
Forester, R. W., Caldwell, W. G. E., and Oro, F. H. 1977. Oxygen and carbon isotopic study of ammonites from the Late Cretaceous Bearpaw Formation in southwestern Saskatchewan. Canadian Journal of Earth Science, 14:20862100.Google Scholar
Fritz, P., and Poplawski, S. 1974. 18O and 13C in the shells of freshwater mollusks and their environment. Earth and Planetary Science Letters, 24:9198.Google Scholar
Geary, D. H., Brieske, T. A., and Bemis, B. E. 1992. The influence and interaction of temperature, salinity, and upwelling on the stable isotopic profiles of strombid gastropod shells. Palaios, 7:7785.Google Scholar
Grossman, E. L., and Ku, T.-L. 1986. Oxygen and carbon isotopic fractionation in biogenic aragonite: temperature effects. Chemical Geology, 59:5974.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, 105:12841296.Google Scholar
Grossman, E. L., Zhang, C., and Yancey, T. E. 1991. Stable-isotope stratigraphy of brachiopods from Pennsylvanian shales in Texas. Geological Society of America Bulletin, 103:953965.Google Scholar
Gruszczynski, M., Halas, S., Hoffman, A., and Malkowski, K. 1989. A brachiopod calcite record of the oceanic carbon and oxygen shifts at the Permian/Triassic transition. Nature, 337:6468.CrossRefGoogle Scholar
Hart, S. R., and Blusztajn, J. 1998. Clams as recorders of ocean ridge volcanism and hydrothermal vent field activity. Science, 280:883886.CrossRefGoogle ScholarPubMed
Heller, J. 1990. Longevity in molluscs. Malacologia, 31:259295.Google Scholar
Higsmith, R. D. 1979. Coral growth rates and environmental control of density banding. Journal of Experimental Marine Biology and Ecology, 37:105125.Google Scholar
Horibe, Y., and Oba, T. 1972. Temperature scales of aragonite-water and calcite-water systems. Fossils, 23/24:6979.Google Scholar
Hudson, J. H., Shinn, E., Halley, R., and Lidz, B. 1976. Sclerochronology: A new tool for interpreting past environments. Geology, 4:361364.Google Scholar
James, N. P., Bone, Y., and Kyser, T. K. 1997. Brachiopod δ18O values do reflect ambient oceanography: Lacepede Shelf, southern Australia. Geology, 25:551554.Google Scholar
Jones, D. S. 1983. Sclerochronology: reading the record of the molluscan shell. American Scientist, 71:384391.Google Scholar
Jones, D. S., and Allmon, W. D. 1995. Records of upwelling, seasonality and growth in stable-isotope profiles of Pliocene mollusks from Florida. Lethaia, 28:6174.Google Scholar
Jones, D. S., Arthur, M. A., and Allard, D. J. 1989. Sclerochronological records of temperature and growth from shells of Mercenaria mercenaria from Narragansett Bay, Rhode Island. Marine Biology, 102:225234.Google Scholar
Jones, D. S., and Quitmyer, I. R. 1996. Marking time with bivalve shells: oxygen isotopes and season of annual increment formation. Palaios, 11:340346.Google Scholar
Jones, D. S., Thompson, I., and Ambrose, W. G. 1978. Age and growth rate determinations for the Atlantic surf clam Spisula solidissima (Bivalvia: Mactracea), based on internal growth lines in shell cross-sections. Marine Biology, 47:6370.CrossRefGoogle Scholar
Jones, D. S., Williams, D. F., and Arthur, M. A. 1983. Growth history and ecology of the Atlantic surf clam, Spisula solidissima (Dillwyn), as revealed by stable isotopes and annual shell increments. Journal of Experimental Marine Biology and Ecology, 73:225242.Google Scholar
Jones, D. S., Williams, D. F., and Romanek, C. S. 1986. Life history of symbiont-bearing giant clams from stable isotope profiles. Science, 231:4648.Google Scholar
Jordan, R., and Stahl, W. 1970. Isotopische Paläotemperatur-Bestimmungen an jurassischen Ammoniten und grundsätzliche Voraussetzungen für diese Methode. Geologische Jahrbüch, 89:3362.Google Scholar
Keith, M. L., Anderson, G. M., and Eichler, E. 1964. Carbon and oxygen isotopic composition of mollusk shells from marine and fresh-water environments. Geochimica et Cosmochimica Acta, 28:17571786.Google Scholar
Killingley, J. S. 1980. Migrations of California gray whales tracked by oxygen-18 variations in their epizoic barnacles. Science, 207:759760.Google Scholar
Killingley, J. S., and Berger, W. H. 1979. Stable isotopes in a mollusk shell: detection of upwelling events. Science, 205:186188.Google Scholar
Killingley, J. S., and Newman, W. A. 1982. 18O fractionation in barnacle calcite: a barnacle paleotemperature equation. Journal of Marine Research, 40:893901.Google Scholar
Klein, R. T., Lohmann, K. C., and Kennedy, G. L. 1997. Elemental and isotopic proxies of paleotemperature and paleosalinity: Climate reconstruction of the marginal northeast Pacific ca. 80 ka. Geology, 25:363366.Google Scholar
Klein, R. T., Lohmann, K. C., and Thayer, C. W. 1996. Bivalve skeletons record sea-surface temperature and δ18O via Mg/Ca and 18O/16O ratios. Geology, 24:415418.2.3.CO;2>CrossRefGoogle Scholar
Knutson, D. W., Buddemeier, R. W., and Smith, S. V. 1972. Coral chronometers: Seasonal growth bands in reef corals. Science, 177:270272.Google Scholar
Krantz, D. E. 1990. Mollusk-isotope records of Plio-Pleistocene marine paleoclimate, U. S. Middle Atlantic Coastal Plain. Palaios, 5, 317335.Google Scholar
Krantz, D. E., Jones, D. S., and Williams, D. F. 1984. Growth rates of the sea scallop, Placopecten magellanicus, determined from the 18O/16O record in shell calcite. Biological Bulletin, 167:186199.Google Scholar
Krantz, D. E., Williams, D. F., and Jones, D. S. 1987. Ecological and paleoenvironmental information using stable isotope profiles from living and fossil molluscs. Palaeogeography, Palaeoclimatology, Palaeoecology, 58:249266.CrossRefGoogle Scholar
Landman, N. H., and Cochran, J. K. 1987. Growth and longevity of Nautilus , p. 401420. In Saunders, W. B. and Landman, N. H. (eds.), Nautilus. The Biology and Paleobiology of a Living Fossil. Plenum Press, New York.Google Scholar
Landman, N. H., and Cochran, J. K., Rye, D. M., Tanabe, K., and Arnold, J. M. 1994. Early life history of Nautilus: evidence from isotopic analyses of aquarium-reared specimens. Paleobiology, 20:4051.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
Lutz, R. A., Fritz, L. W., and Cerrato, R. M. 1988. A comparison of bivalve (Calyptogena magnifica) growth at two deep-sea hydrothermal vents in the eastern Pacific. Deep-Sea Research, 35:17931810.Google Scholar
Lutz, R. A., Fritz, L. W., and Rhoads, D. C. 1985. Molluscan growth at deep-sea hydrothermal vents. Bulletin of the Biological Society of Washington, 6:199210.Google Scholar
Marshall, J. D., Duncan, P., Clarke, A., Nolan, C. P., and Sharman, J. 1997. Stable-isotopic composition of skeletal carbonates from living Antarctic marine invertebrates. Lethaia, 29:203212.Google Scholar
McMichael, D. F. 1974. Growth rate, population size and mantle coloration in the small giant clam Tridacna maxima (Röding), at One Tree Island, Capricorn Group, Queensland. Proceedings of the Second International Coral Reef Symposium, Brisbane, 2:241254.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., and Grossman., E. L. and Yancey, T. E. 1997. Stable carbon and oxygen isotope shifts in Permian seas of West Spitsbergen—Global change or diagenetic artifact? Geology, 25:227230.2.3.CO;2>CrossRefGoogle Scholar
Mook, W. G. 1971. Paleotemperatures and chlorinities from stable carbon and oxygen isotopes in shell carbonates. Palaeogeography, Palaeoclimatology, Palaeoecology, 9:245263.Google Scholar
Mook, W. G., and Vogel, J. C. 1968. Isotopic equilibrium between shells and their environment. Science, 159:874875.Google Scholar
Moore, W. S., and Krishnaswami, S. 1972. Coral growth rates using 228Ra and 210Pb. Earth and Planetary Science Letters, 15:187192.Google Scholar
Moore, W. S., and Krishnaswami, S. 1974. Correlation of X-radiography revealed banding in corals with radiometric growth rates. Proceedings of the Second International Coral Reef Symposium, Brisbane, 2:269276.Google Scholar
Nozaki, Y., Rye, D. M., Turekian, K. K., and Dodge, R. E. 1978. A 200 year record of carbon-13 and carbon-14 variations in a Bermuda coral. Geophysical Research Letters, 5:825828.Google Scholar
Oba, T., Kai, M., and Tanabe, K. 1992. Early life history and habitat of Nautilus pompilius inferred from oxygen isotope examinations. Marine Biology, 113:211217.Google Scholar
Oba, T., and Tanabe, K. 1983. Oxygen isotope analysis of the shells of Nautilus pompilius . Kagoshima University Research Center for the South Pacific, Occasional Papers, 1:2629.Google Scholar
Pannella, G., and MacClintock, C. 1968. Biological and environmental rhythms reflected in molluscan shell growth. Journal of Paleontology, Memoir 2, 42(5, suppl.):6480.Google Scholar
Pätzold, J. 1984. Growth rhythms recorded in stable isotopes and density bands in the reef coral Porites lobata (Cebu, Philippines). Coral Reefs, 3:7890.Google Scholar
Pätzold, J., Ristedt, H., and Wefer, G. 1987. Rate of growth and longevity of a large colony of Pentapora foliacea (Bryozoa) recorded in their oxygen isotope profiles. Marine Biology, 96:535538.Google Scholar
Peterson, C. H. 1986. Quantitative allometry of gamete production by Mercenaria mercenaria into old age. Marine Ecology Progress Series, 29:9397.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, 97:12621269.Google Scholar
Purton, L., and Brasier, M. 1997. Gastropod carbonate δ18O and δ13C values record strong seasonal productivity and stratification shifts during the late Eocene in England. Geology, 25:871874.Google Scholar
Quitmyer, I. R., Jones, D. S., and Arnold, W. S. 1997. The sclerochronology of hard clams, Mercenaria spp., from the south-eastern U.S.A.: A method of elucidating the zooarchaeological records of seasonal resource procurement and seasonality in prehistoric shell middens. Journal of Archaeological Science, 24:825840.Google Scholar
Rhoads, D. C., and Lutz, R. A., eds. 1980. Skeletal Growth of Aquatic Organisms: Biological Records of Environmental Change. Plenum Press, New York, 750 p.Google Scholar
Rhoads, D. C., and Lutz, R. A., Revelas, E. C. and Cerrato, R. M. 1981. Growth of bivalves at deep-sea hydrothermal vents along the Galapagos Rift. Science, 214:911913.CrossRefGoogle ScholarPubMed
Rio, M., Roux, M., Renard, M., and Schein, E. 1992. Chemical and isotopic features of present day bivalve shells from hydrothermal vents or cold seeps. Palaios, 7:351360.Google Scholar
Rollins, H. B., Sandweiss, D. H., Brand, U., and Rollins, J. C. 1987. Growth increment and stable isotope analysis of marine bivalves: Implications for the geoarchaeological record of El Niño. Geoarchaeology, 2:181197.Google Scholar
Romanek, C. S., and Grossman, E. L. 1989. Stable isotope profiles of Tridacna maxima as environmental indicators. Palaios, 4:402413.CrossRefGoogle Scholar
Romanek, C. S., Jones, D. S., Williams, D. F., Krantz, D. E., and Radtke, R. 1987. Stable isotopic investigation of physiological and environmental changes recorded in shell carbonate from the giant clam Tridacna maxima . Marine Biology, 94:385393.CrossRefGoogle Scholar
Ropes, J. W., Jones, D. S., Murawski, S. A., Serchuk, F. M., and Jearld, A. Jr. 1984. Documentation of annual growth lines in Arctica islandica (Linné). Fishery Bulletin, 82:119.Google Scholar
Ross, C. A., Moore, G. T., and Hayashida, N. 1992. Late Jurassic paleoclimate simulation – paleoecological implications for ammonoid provinciality. Palaios, 7:487507.Google Scholar
Roulier, L. M., and Quinn, T. M. 1995. Seasonal- to decadal-scale climatic variability in southwest Florida during the middle Pliocene: Inferences from a coralline stable isotope record. Paleoceanography, 10:429443.Google Scholar
Rye, D. M., and Sommer, M. A. II. 1980. Reconstructing paleotemperature and paleosalinity regimes with oxygen isotopes, p. 169202. In Rhoads, D. C. and Lutz, R. A. (eds.), Skeletal Growth of Aquatic Organisms. Plenum Press, New York.Google Scholar
Shackleton, N. J. 1973. Oxygen isotope analysis as a means of determining season of occupation of prehistoric midden sites. Archaeometry, 15:133141.Google Scholar
Smith, J. E., Risk, M. J., Schwarcz, H. P., and McConnaughey, T. A. 1997. Rapid climate change in the North Atlantic during the Younger Dryas recorded by deepsea corals. Nature, 386:818820.Google Scholar
Spero, H. J. 1988. Ultrastructural examination of chamber morphogenesis and biomineralization in the planktonic foraminifer Orbulina universa . Marine Biology, 99:920.Google Scholar
Spero, H. J., and Williams, D. F. 1989. Opening the carbon isotope “vital effect” black box. 1. Seasonal temperatures in the euphotic zone. Paleoceanography, 4:593601.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
Stanley, G. D. Jr., and Swart, P. K. 1995. Evolution of the coral-zooxanthellae symbiosis during the Triassic: a geochemical approach. Paleobiology, 21:179199.Google Scholar
Stecher, H. A. III, Krantz, D. E., Lord, C. J. III, Luther, G. W. III, and Bock, K. W. 1996. Profiles of strontium and barium in Mercenaria mercenaria and Spisula solidissima shells. Geochimica et Cosmochimica Acta, 60:34453456.Google Scholar
Steuber, T. 1996. Stable isotope sclerochronology of rudist bivalves: Growth rates and Late Cretaceous seasonality. Geology, 24:315318.Google Scholar
Stevens, G. R., and Clayton, R. N. 1971. Oxygen isotope studies on Jurassic and Cretaceous belemnites from New Zealand and their biogeographic significance. New Zealand Journal of Geology and Geophysics, 14:829887.Google Scholar
Swart, P. K., McNeill, D., Grammer, M., Jull, T., and Beck, W. 1994. Intra-annual, inter-annual, and decadal climate variations as recorded in the skeletons of sclerosponges from the Bahamas. Geological Society of America Abstracts with Programs, 26(7):A-228.Google Scholar
Swart, P. K., Dodge, R. E., and Hudson, H. J. 1996. A 240-year stable oxygen and carbon isotopic record in a coral from South Florida: implications for the prediction of precipitation in southern Florida. Palaios, 11:362375.Google Scholar
Tasch, P. 1980. Paleobiology of the Invertebrates: Data Retrieval from the Fossil Record. John Wiley & Sons, New York, 975 p.Google Scholar
Taylor, B. E., and Ward, P. D. 1983. Stable isotopic studies of Nautilus macromphalus Sowerby (New Caledonia) and Nautilus pompilius L. (Fiji). Palaeogeography, Palaeoclimatology, Palaeoecology, 41:116.Google Scholar
Teranes, J. L., Geary, D. H., and Bemis, B. E. 1996. The oxygen isotopic record of seasonality in Neogene bivalves from the Central American Isthmus, p. 105129. In Jackson, J. B. C., Budd, A. F., and Coates, A. G. (eds.), Evolution and Environment in Tropical America. University of Chicago Press, Chicago.Google Scholar
Thompson, I., Jones, D. S., and Dreibelbis, D. 1980. Annual internal growth banding and life history of the ocean quahog, Arctica islandica (Mollusca: Bivalvia). Marine Biology, 57:2534.Google Scholar
Tourtelot, H. A., and Rye, R. O. 1969. Distribution of oxygen and carbon isotopes in fossils of Late Cretaceous age, western interior region of North America. Geological Society of America Bulletin, 80:19031922.Google Scholar
Tudhope, A. W., Lea, D. W., Shimmield, G. B., Chilcott, C. P., and Head, S. 1996. Monsoon climate and Arabian Sea coastal upwelling recorded in massive corals from southern Oman. Palaios, 11:347361.Google Scholar
Turekian, K. K., and Cochran, J. K. 1981. Growth rate of a vesicomyid clam from the Galapagos spreading center. Science, 214:909911.Google Scholar
Turekian, K. K., and Cochran., J. K., Kharkar, D. P., Cerrato, R. M., Vaisnys, J. R., Sanders, H. L., Grassle, J. F., and Allen, J. A. 1975. Slow growth rate of a deep-sea clam determined by 228Ra chronology. Proceedings of the National Academy of Sciences, 72:28292832.Google Scholar
Turekian, K. K., and Cochran., J. K., and Nozaki, Y. 1979. Growth rate of a clam from the Galapagos Rise hot spring field using natural radionuclides. Nature, 280:385387.Google Scholar
Turekian, K. K., and Cochran., J. K., and Nozaki., Y., Thompson, I., and Jones, D. S. 1982. Determination of shell deposition rates of Arctica islandica from the New York Bight using natural 228Ra and 228Th and bomb-produced 14C. Limnology and Oceanography, 27:737741.CrossRefGoogle Scholar
Urey, H. C. 1947. The thermodynamic properties of isotopic substances. Journal of the Chemical Society, 1947:562581.Google Scholar
Urey, H. C., Lowenstam, H. A., Epstein, S., and McKinney, C. R. 1951. Measurements 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
Veinott, G. I., and Cornett, R. J. 1996. Identification of annually produced opaque bands in the shell of the freshwater mussel Elliptio complanata using the seasonal cycle of δ18O. Canadian Journal of Fisheries and Aquatic Science, 53:372379.Google Scholar
Veizer, J. 1983. Chemical diagenesis of carbonates: theory and application of trace element technique, p. 3-1-3-100. In Arthur, M. A. (organizer), Stable Isotopes in Sedimentary Geology: SEPM Short Course No. 10. Society of Economic Paleontologists and Mineralogists, Tulsa, Oklahoma.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, 50:16791696.Google Scholar
Wefer, G., and Berger, W. H. 1980. Stable isotopes in benthic foraminifera: Seasonal variation in large tropical species. Science, 209:803805.Google Scholar
Wefer, G., and Berger, W. H. 1991. Isotope paleontology: growth and composition of extant calcareous species. Marine Geology, 100:207248.Google Scholar
Wefer, G., and Killingley, J. S. 1980. Growth histories of strombid snails from Bermuda recorded in their O-18 and C-13 profiles. Marine Biology, 60:129135.Google Scholar
Weidman, C. R., Jones, G. A., and Lohmann, K. C. 1994. The long-lived mollusc Arctica islandica. A new paleoceanographic tool for the reconstruction of bottom temperatures for the continental shelves of the northern North Atlantic Ocean. Journal of Geophysical Research, 99:18,305–18,314.Google Scholar
Whittaker, S. G., Kyser, T. K., and Caldwell, W. G. E. 1987. Paleoenvironmental geochemistry of the Clagett marine cyclothem in south-central Saskatchewan. Canadian Journal of Earth Science, 24:967984.Google Scholar
Williams, D. F., Arthur, M. A., Jones, D. S., and Healy-Williams, N. 1982. Seasonality and mean annual sea surface temperatures from isotopic and sclerochronological records. Nature, 296:432434.Google Scholar
Williams, D. F., Sommer, M. A. II, and Bender, M. L. 1977. Carbon isotopic compositions of recent planktonic foraminifera of the Indian Ocean. Earth and Planetary Science Letters, 36:391403.Google Scholar