Hostname: page-component-7bb8b95d7b-dtkg6 Total loading time: 0 Render date: 2024-09-18T03:15:10.729Z Has data issue: false hasContentIssue false

Ostracodes and Their Shell Chemistry: Implications for Paleohydrologic and Paleoclimatologic Applications

Published online by Cambridge University Press:  21 July 2017

Emi Ito
Affiliation:
Department of Geology and Geophysics and Limnological Research Center, University of Minnesota, Minneapolis, MN 55455 USA
Patrick De Deckker
Affiliation:
Department of Geology and CRC LEME, The Australian National University, Canberra ACT 0200, Australia
Stephen M. Eggins
Affiliation:
Research School of Earth Sciences, The Australian National University, Canberra ACT 0200, Australia
Get access

Abstract

The shell chemistry of ostracodes is a useful indicator of past environmental conditions especially when the chemistry data are considered along with other proxy data. The complexities involved with the chemical and isotopic changes accompanying hydrologic change, solute evolution, and the autoecology of ostracodes all point to the need to exercise caution when interpreting shell chemistry. Nevertheless, the stable-isotope values and cation ratios (e.g., Mg/Ca, Sr/Ca) as well as the species assemblage of ostracodes can provide powerful tools for the reconstruction of paleoclimate and paleohydrology. In particular, the changes in Mg/Ca and Sr/Ca of well-calcified ostracodes shells record the qualitative changes in solute composition, and when the dissolved Mg/Ca remains relatively constant, the Mg/Ca in the ostracode shell is proportional to water temperature.

Type
Research Article
Copyright
Copyright © 2003 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

Adkins, J. F. 2002. Unpacking “vital effects” in biogenic carbonates using deep-sea corals. Geochimica et Cosmochimica Acta, 66:A7 (abstract).Google Scholar
Aguirre, M. L., Leng, M. J., and Spiro, B. 1998. Variation In isotopic composition (C, O and Sr) of Holocene Mactra Isabelleana (Bivalvia) From the coast of Buenos Aires Province, Argentina. Holocene, 8:613621.Google Scholar
Aizenberg, J., Albeck, S., Weiner, S., and Addadi, L. 1994. Crystal-protein interactions studied by overgrowth of calcite on biogenic skeletal elements. Journal of Crystal Growth, 142:156164.CrossRefGoogle Scholar
Alm, G. C. 1915. Monographue der Schwedischen Süsswasserostrakoden nebst Systematischen Besprechungen der Tribus Podocopa. Zoologiska Bidrag från Uppsala, 4:1246.Google Scholar
Anadón, P., Gliozzi, E., and Mazzini, I. 2002. Paleoenvironmental reconstruction of marginal marine environments from combined paleoecological and geochemiscal analysis of ostracods, p. 227247 In Holmes, J. A. and Chivas, A. R. (eds.), The Ostracoda: Applications in Quaternary Research. American Geophysical Union, Washington, D.C. Google Scholar
Bacon, S. W. 1999. Seasonal constraints on chemical composition and isotopic ratios in ostracode shells from Page Pond, Ohio. , Kent State University, Kent, 95 p.Google Scholar
Bacon, S. W., Ito, E., Smith, A. J., Palmer, D. F., and Forester, R. M. 1999. Seasonal constraints on d18O values of living freshwater ostracodes. EOS, Transactions, American Geophysical Union, 80:176 (abstract).Google Scholar
Bate, R. H., and East, B. A. 1972. The structure of the ostracode carapace. Lethaia, 5:177194.Google Scholar
Bate, R. H., and East, B. A. 1975. The ultrastructure of the ostracode (crustacea) integument. Bulletin of American Paleontolog y, 65:529547.Google Scholar
Bodergat, A. M. 1978. L'intensité lumineuse, son influence sur la teneur en phosphore des carapaces d'Ostracodes. Géobios, 11:715735.Google Scholar
Boomer, I. 1993. Palaeoenvironmental Indicators from Late Holocene and Contemporary Ostracoda of the Aral Sea. Palaeogeography, Palaeoclimatology, Palaeoecology, 103:141153.Google Scholar
Botz, R., Pokojski, H.-D., Schmidt, M., and Thomm, M. 1996. Carbon Isotope fractionation during bacterial methanogenesis by CO2 reduction. Organic Geochemistry, 25:255262.Google Scholar
Bridgwater, N. D., Heaton, T. H. E., and O'Hara, S. L. 1999. A late Holocene palaeolimnological record from central Mexico, based on faunal and stable-isotope analysis of ostracod shells. Journal of Paleolimnology, 22:383397.Google Scholar
Brouwers, E. M. 1988. Sediment transport detected from the analysis of ostracod population structure: an example from the Alaskan continental shelf, p. 231244 In De Deckker, P., Colin, J.-P., and Peypouquet, J.-P. (eds.), Ostracoda in the Earth Sciences. Elsevier, Amsterdam.Google Scholar
Brown, S. J., and Elderfield, H. 1996. Variations in Mg/Ca and Sr/Ca ratios of planktonic foraminifera caused by postdepositional dissolution: Evidence of shallow Mg-dependent dissolution. Paleoceanography, 11:543551.Google Scholar
Cadot, H. M., and Kaesler, R. L. 1977. Magnesium content of calcite in carapaces of benthic marine Ostracoda. The University of kansas Paleontological Contributions, 87:123.Google Scholar
Carroll, A. R. 2003. Sr-isotopic records of paleoweathering and paleodrainage in ancient lake basins (abstract). 3rd International Limnogeology Congress, Tucson, Arizona, March 29-April 2, 2003:46 (abstract).Google Scholar
Chappell, J. 1974. Geology of coral terraces, Huon Peninsula, New Guinea: A study of Quaternary tectonic movements and sea-level changes. Geological Society of America Bulletin, 85:553570.Google Scholar
Chappell, J., and Shackleton, N. J. 1986. Oxygen isotopes and sea level. Nature, 324:137140.Google Scholar
Chave, K. E. 1954. Aspects of the biogeochemistry of magnesium. 1. Calcareous marine organisms. Journal of Geology, 62:266283.Google Scholar
Chester, R. 2003. Marine Geochemistry. Blackwell Publishing, Oxford, 506 p.Google Scholar
Chivas, A. R., De Deckker, P., Cali, J. A., Chapman, A., Kiss, E., and Shelley, J. M. G. 1993. Coupled stable-isotope and trace-element measurements of lacustrine carbonates as paleoclimatic indicators, p. 113121 In Swart, P. K., Lohmann, K. C., Mckenzie, J., and Savin, S. (eds.), Climate Change in Continental Isotopic Records. American Geophysical Union.Google Scholar
Chivas, A. R., De Deckker, P., and Shelley, J. M. G. 1983. Magnesium, strontium, and barium partitioning in nonmarine ostracode shells and their use in paleoenvironmental reconstructions - a preliminary study, p. 238249 In Maddocks, R. F. (ed.), Applications of Ostracoda. University of Houston, Geosciences Department, Houston.Google Scholar
Chivas, A. R., De Deckker, P., and Shelley, J. M. G. 1985. Strontium content of ostracods indicates lacustrine palaeosalinity. Nature, 316:251253.CrossRefGoogle Scholar
Chivas, A. R., De Deckker, P., and Shelley, J. M. G. 1986a. Magnesium and strontium in nonmarine ostracod shells as indicators of palaeosalinity and palaeotemperature. Hydrobiologia, 143:135142.Google Scholar
Chivas, A. R., De Deckker, P., and Shelley, J. M. G. 1986b. Magnesium content of nonmarine ostracod shells: a new palaeosalinometer and palaeothermometer. Palaeogeography Palaeoclimatology Palaeoecology, 54:4361.Google Scholar
Chivas, A. R., De Deckker, P., Wang, S. X., and Cali, J. A. 2002. Oxygen-isotope systematics of the nektic ostracod Austracypris robusta., p. 301313 In Holmes, J. A. and Chivas, A. R. (eds.), The Ostracoda: Applications in Quaternary Research. American Geophysical Union, Washington, D.C. Google Scholar
Cole, J. J., Caraco, N. F., Kling, G. W., and Kratz, T. K. 1994. Carbon dioxide supersaturation in the surface waters of lakes. Science, 265:15681570.Google Scholar
Corrège, T. 1993. Preliminary results of palaeotemperature reconstruction using the magnesium to calcium ratio of deep-sea ostracode shells from the Late Quaternary of Site 822, Leg 133 (western Coral Sea). Proceedings of the ODP, Scientific Results, 133:175180.Google Scholar
Corrège, T., and De Deckker, P. 1997. Faunal and geochemical evidence for changes in intermediate water temperature and salinity in the western Coral Sea (Northeast Australia) during the late Quaternary: The late Quaternary palaeoceanography of the Australasian region. Palaeogeography, Palaeoclimatology, Palaeoecology, 131:183205.CrossRefGoogle Scholar
Cox, R. A. 1965. The physical properties of seawater., p. 73120 In Riley, J. P. and Skirrow, G. (eds.), Chemical Oceanography. Volume 1. Academic Press, New York.Google Scholar
Craig, H. 1961. Isotopic variations in meteoric waters. Science, 133:17021703.Google Scholar
Craig, H., and Gordon, L. 1965. Deuterium and oxygen-18 isotope variations in the ocean and marine atmosphere, p. 9130 In Tongiorgi, E. (ed.), Stable Isotopes in Oceanographic Studies and Paleotemperatures, Spoleto 1965. Conoglio Nazionale delle Ricerche, Pisa, Italy.Google Scholar
Criss, R. E. 1999. Principles of Stable Isotope Distribution. Oxford University Press, New York, 254 p.Google Scholar
Cronin, T., Dwyer, G. S., Baker, P. A., Rodriguez-Lazaro, J., and Briggs, W. M. 1996. Deep-sea ostracode shell chemistry (Mg/Ca ratios) and late Quaternary Arctic Ocean history., p. 188–134 In Andrews, J. E. (ed.), Late Glacial Paleoceanography of the North Atlantic Margins. Royal Society of Edinburgh, Edinburgh, Scotland.Google Scholar
Curry, B. B., Anderson, T. F., and Lohmann, K. C. 1997. Unusual carbon and oxygen isotopic ratios of ostracodal calcite from last interglacial (Sangamon episode) lacustrine sediment in Raymond basin, Illinois, USA. Journal of Paleolimnology, 17:421435.Google Scholar
Cwynar, L. C., and Levesque, A. J. 1995. Chironomid evidence for late-glacial climatic reversals in Maine. Quaternary Research, 43:405413.Google Scholar
De Deckker, P. 2002. Ostracod paleoecology, p. 121134 In Holmes, J. A. and Chivas, A. R. (eds.), The Ostracoda: Applications in Quaternary Research. American Geophysical Union, Washington, D.C. Google Scholar
De Deckker, P., Chivas, A. R., and Shelley, J. M. G. 1988a. Paleoenvironment of the Messinian Mediterranean “Lago Mare” from strontium and magnesium in ostracode shells. Palaios, 3:352358.Google Scholar
De Deckker, P., Chivas, A. R., and Shelley, J. M. G. 1999. Uptake of Mg and Sr in the euryhaline ostracod Cyprideis determined from in vitro experiments. Palaeogeography Palaeoclimatology Palaeoecology, 148:105116.Google Scholar
De Deckker, P., Chivas, A. R., Shelley, J. M. G., and Torgersen, T. 1988b. Ostracod shell chemistry: a new palaeoenvironmental indicator applied to a regressive/transgressive record from the Gulf of Carpentaria, Australia. Palaeogeogr., Palaeoclimatol., Palaeoecol., 66:231241.Google Scholar
De Deckker, P., and Forester, R. M. 1988. The use of ostracods to reconstruct continental palaeoenvironmental records, p. 175198 In De Deckker, P., Colin, J.-P., and Peypouquet, J.-P. (eds.), Ostracoda in the Earth Sciences. Elsevier, Amsterdam.Google Scholar
De Villiers, S., and Flecker, R. 2002. Foraminiferal shell heterogeneity and selective diagenesis revealed: LA-ICP-MS as a poweful new tool. Geochimica et Cosmochimica Acta, 66:A173 (abstract).Google Scholar
Dean, W. E. 1999. The carbon cycle and biogeochemical dynamics in lake sediments. Journal of Paleolimnology, 21:375393.Google Scholar
Delorme, L. D. 1989. Methods in Quaternary ecology. No. 7: Freshwater Ostracoda. Geosciences Canada, 16:8590.Google Scholar
Dettman, D. L., Smith, A. J., Rea, D. K., Moore, T. C., and Lohmann, K. C. 1995. Glacial meltwater in Lake Huron during early postglacial time as inferred from single-valve analysis of oxygen isotopes in ostracodes. Quaternary Research, 43:297310.Google Scholar
Didié, C. 2001. Late Quaternary climate variations recorded in North Atlantic deep-sea benthic ostracodes. Polarforschung Meeresforschung, 390.Google Scholar
Didié, C., and Bauch, H. A. 2002. Implications of Upper Quaternary stable isotope records of marine ostrcaodes and benthic foraminifers for paleoecological and paleoceanogrphical investigations, p. 279299 In Holmes, J. A. and Chivas, A. R. (eds.), The Ostracoda: Applications in Quaternary Research. American Geophysical Union, Washington, D.C. Google Scholar
Didié, C., Bauch, H. A., and Helmke, J. P. 2002. Late Quaternary deep-sea ostracodes in the polar and subpolar North Atlantic: paleoecological and paleoenvironmental implications. Palaeogeography Palaeoclimatology Palaeoecology, 184:195212.Google Scholar
Dodge, R. E., Fairbanks, R. G., Benninger, L. K., and Maurrasse, F. 1983. Pleistocene sea-levels from raised coral reefs of Haiti. Science 219:14231424.Google Scholar
Donovan, J. J., Smith, A. J., Panek, V. A., Engstrom, D. R., and Ito, E. 2002. Climate-driven hydrologic transients in lake sediment records: Calibration of groundwater conditions using 20th century drought. Quaternary Science Reviews, 21:605624.Google Scholar
Dwyer, G. S., Cronin, T. M., and Baker, P. A. 2002. Trace elements in marine ostracodes, p. 205225 In Holmes, J. A. and Chivas, A. R. (eds.), The Ostracoda: Applications in Quaternary Research. American Geophysical Union, Washington, D.C. CrossRefGoogle Scholar
Dwyer, G. S., Cronin, T. M., Baker, P. A., Raymo, M. E., Buzas, J. S., and Correge, T. 1995. North Atlantic deepwater temperature change during late Pliocene and late Quaternary climatic cycles. Science, 270:13471351.Google Scholar
Eggins, S. M., De Deckker, P., and Marshall, J. in press. Mg/Ca variation in planktonic foraminifera tests: Implications for reconstructing paleoseawater temperature and habitat migration. Earth and Planetary Science Letters.Google Scholar
Eggins, S. M., Kinsley, L. K., and Shelley, J. M. G. 1998. Deposition and element fractionation processes occurring during atmospheric pressure laser sampling for analysis by ICPMS. Applied Surface Science, 127-129:278286.Google Scholar
Emiliani, C. 1992. Planet Earth: Cosmology, Geology, and the Evolution of Life and Environment. Cambridge University Press, Cambridge, 719 p.Google Scholar
Engstrom, D. R., Xia, J., and Ito, E. 1993. A study of the variability of Mg/Ca and Sr/Ca on ostracode calcite by laboratory culture and field collection. EOS, 74:365 (abstract).Google Scholar
Erez, J., Bentov, S., Brownwlee, C., Raz, M., and Rinkevich, B. 2002. Biomineralization mechanisms in foraminifera and corals and their paleoceanographic implications. Geochimica et Cosmochimica Acta, 66:A216 (abstract).Google Scholar
Eugster, H. P., and Hardie, L. A. 1978. Saline lakes, p. 237293 In Lerman, A. (ed.), Lakes–Chemistry, Geology and Physics. Springer-Verlag, New York.CrossRefGoogle Scholar
Eugster, H. P., and Jones, B. F. 1979. Behavior of major solutes during closed-basin brine evolution. American Journal of Science, 279:609631.Google Scholar
Fairbanks, R. G., and Matthews, R. K. 1978. The marine oxygen isotope record in Pleistocene coral, Barbados, West Indies. Quaternary Research, 10:181196.Google Scholar
Falini, G., Albeck, S., Weiner, S., and Addadi, L. 1996. Control of aragonite or calcite polymorphism by mollusk shell macromolecules. Science, 271:6769.Google Scholar
Forester, R. M. 1983. Relationship of two lacustrine ostracode species to solute composition and salinity: Implications for paleohydrochemistry. Geology, 11:435438.Google Scholar
Forester, R. M. 1986. Determination of the dissolved anion composition of ancient lakes from fossil ostracodes. Geology, 14:796798.Google Scholar
Fritz, P., and Fontes, J. C. 1980. Handbook of Environmental Isotope Geochemistry, 1, The Terrestrial Environment, A. Elsevier Scientific Publishing Company, Amsterdam, 545 p.Google Scholar
Fritz, P., and Fontes, J. C. 1989. Handbook of Environmental Isotope Geochemistry, 3, The Marine Environment, A. Elsevier Scientific Publishing Company, Amsterdam, 428 p.Google Scholar
Fritz, P., and Fontes, J. C. 1986. Handbook of Environmental Isotope Geochemistry, 2, The Terrestrial Environment, B. Elsevier Scientific Publishing Company, Amsterdam, 557 p.Google Scholar
Fritz, S. C. 1990. 20th-Century Salinity and Water-Level Fluctuations in Devils Lake, North-Dakota - Test of a Diatom-Based Transfer-Function. Limnology and Oceanography, 35:17711781.Google Scholar
Fritz, S. C., Juggins, S., Battarbee, R. W., and Engstrom, D. R. 1991. Reconstruction of past changes in salinity and climate using a diatom-based transfer function. Nature, 352:706708.Google Scholar
Gat, J. R. 1995. Stable isotopes of fresh and saline lakes., p. 139165 In Lerman, A., Imboden, D., and Gat, J. R. (eds.), Physics and Chemistry of Lakes. Springer-Verlag, Berlin.Google Scholar
Goldsmith, J. R., and Newton, R. C. 1969. P-T-X relations in the system CaCO3-MgCO3 at high temperatures and pressures. American Journal of Science, 267-A:160190.Google Scholar
Gonfiantini, R. 1986. Environmental isotopes in lake studies, p. 113168 In Fritz, P. and Fontes, J. C. (eds.), Handbook of Environmental Isotope Geochemistry, 2, Terrestrial Environment, B. Elsevier Scientific Publishing Company.Google Scholar
Haskell, B. J., Engstrom, D. R., and Fritz, S. C. 1996. Late Quaternary paleohydrology in the North American Great Plains inferred from the geochemistry of endogenic carbonate and fossil ostracodes from Devils Lake, North Dakota, USA. Palaeogeography, Palaeoclimatology, Palaeoecology, 124:179193.Google Scholar
Hendry, J. P., and Kalin, R. M. 1997. Are oxygen and carbon isotopes of mollusc shells reliable palaeosalinity indicators in marginal marine environments? A case study from the Middle Jurassic of England. Journal of the Geological Society, 154:321333.Google Scholar
Henrichs, M. L., Walker, I. R., and Mathewes, R. W. 2001. Chironomid-based paleosalinity records in southern British Columbia, Canada: a comparison of transfer functions. Journal of Paleolimnology, 26:147159.Google Scholar
Holmes, J. A., and Chivas, A. R. 2002. Ostracod shell chemistry: overview. In Holmes, J. A. and Chivas, A. R. (eds.), The Ostracoda: Applications in Quaternary Research. American Geophysical Union, Washington, D.C. Google Scholar
Hostetler, S. W., and Benson, L. V. 1994. Stable isotopes of oxygen and hydrogen in the Truckee River-Pyramid Lake surface-water system. 2. A predictive model of δ18O and δ2H in Pyramid Lake. Limnology and Oceanography, 39:356364.Google Scholar
Hutchinson, G. E. 1958. Concluding remarks. Cold Spring Harbor Symposia on Quantitative Biology (proceedings), 22:415427.Google Scholar
Hutchinson, G. E. 1978. An Introduction to Population Ecology. Yale University Press, New Haven, 260 p.Google Scholar
Imbrie, J., Hays, J. D., Martinson, D. G., Mcintyre, A., Mix, A. C., Morley, J. J., Pisias, N. G., Prell, W. L., and Shackleton, N. J. 1984. The orbital theory of Pleistocene climate: Support from a revised chronology of the marine 18O record, p. 269305 In A. L. B. E. AL. (ed.), Milankovitch and Climate. Volume 1.Google Scholar
Ingram, B. L., Conrad, M. E., and Ingle, J. C. 1996a. Stable isotope and salinity systematics in estuarine waters and carbonates: San Francisco Bay. Geochimica et Cosmochimica Acta, 60:455467.Google Scholar
Ingram, B. L., Ingle, J. C., and Conrad, M. E. 1996b. Stable isotope record of late Holocene salinity and river discharge in San Francisco Bay, California. Earth and Planetary Science Letters, 141:237247.Google Scholar
Ito, E. 2002. Mg/Ca, Sr/Ca, δ18O and δ13C chemistry of Quaternary lacustrine ostracode shells from the North American continental interior, p. 267278 In Holmes, J. A. and Chivas, A. R. (eds.), The Ostracoda: Applications in Quaternary Research. American Geophysical Union, Washington, D.C. Google Scholar
Ito, E., Bacon, S. W., Smith, A. J., and Palmer, D. F. 2002. Geochemistry of ostracode calcite: Empirical calibration of 3 species from page Pond, Ohio, U.S.A. Geochimica et Cosmochimica Acta, 66:A357 (abstract).Google Scholar
Ito, E., and Curry, B. B. 1998. Last interglacial ostracode geochemistry and assemblage from Pittsburg basin, Illinois. GSA Abstracts with Programs, 30:A-260 (abstract).Google Scholar
Keatings, K. W., Heaton, T. H. E., and Holmes, J. A. 2002. Carbon and oxygen isotope fractionation in nonmarine ostracods: Results from a ‘natural culture’ environment. Geochimica et Cosmochimica Acta, 66:17011711.Google Scholar
Keatings, K. W., Ito, E., Engstrom, D. R., Yu, Z. C., Heaton, T. H. E., and Haskell, B. J. 1999. An investigation into the effect on ostracod shell chemistry of some chemical and physical cleaning techniques. EOS, Supplement, 80:S176 (abstract).Google Scholar
Kim, S.-T., and O'Neil, J. R. 1997. Equilibrium and nonequilibrium oxygen isotope effects in synthetic carbonates. Geochimica et Cosmochimica Acta, 61:34613475.Google Scholar
Labaugh, J. W., and Swanson, G. A. 1992. Changes in the chemical characteristics of water in selected wetlands in the Cottonwood Lake area, North Dakota, U.S.A., 1967-1989, p. 149162 In Robarts, R. D. and Bothwell, M. L. (eds.), Aquatic Ecosystems in Semi-Arid Regions: Implications for Resource Management. Environment Canada, Saskatoon, Saskatchewan.Google Scholar
Last, W. M. 2001. Mineralogical analysis of lake sediments, p. 143187 In Last, W. M. and Smol, J. P. (eds.), Tracking Environmental Change using Lake Sediments. Volume 2: Physical and Geochemical Methods. Kluwer Academic Publishers, Dordrecht, The Netherlands.Google Scholar
Lea, D. W., Pak, D. K., and Spero, H. J. 2000. Climate impact of Late Quaternary equatorial Pacific sea syrface temperature variations. Science, 289:17191724.Google Scholar
Martinez, J. I., De Deckker, P., and Chivas, A. R. 1997. New estimates for salinity changes in the Western Pacific warm pool during the last glacial maximum; oxygen-isotope evidence. Marine Micropaleontology, 32:311340.Google Scholar
Mccrea, J. M. 1950. On the isotopic chemistry of carbonates and a paleotemperature scale. Journal of Chemical Physics, 18:849857.Google Scholar
Mcculloch, M. T., and De Deckker, P. 1989. Sr isotope constraints on the Mediterranean environment at the end of the Messinian salinity crisis. Nature, 342:6265.Google Scholar
Mcculloch, M. T., De Deckker, P., and Chivas, A. R. 1989. Strontium isotope variations in single ostracod valves from the Gulf of Carpentaria, Australia: A palaeoenvironmental indicator. Geochimica et Cosmochimica Acta, 53:17031710.Google Scholar
Mezquita, F., Roca, J. R., and Wansard, G. 1999. Moulting, survival and calcification: the effects of temperature and water chemistry on an ostracod crustacean (Herpetocypris intermedia) under experimental conditions. Archiv für Hydrobiologie, 146:219238.CrossRefGoogle Scholar
Mikalsen, G., and Sejrup, H. P. 2000. Oxygen isotope composition of fjord and river water in the Sognefjorden drainage area, western Norway. Implications for paleoclimate studies. Estuarine Coastal & Shelf Science, 50:441448.Google Scholar
Mix, A. C., and Ruddiman, W. F. 1984. Oxygen isotope analyses and Pleistocene ice volumes. Quaternary Research, 21:120.Google Scholar
Müller, A., and De Deckker, P. 2002. Magnesium, calcium and strontium in waters of the southern Tasman Sea at the confluence of the Indian, Pacific and Southern Oceans. Marine and Freshwater Research, 53:11151128.Google Scholar
Müller, G., Irion, G., and Förstner, U. 1972. Formation and diagenesis of inorganic CaMg carbonates in the lacustrine environment. Naturwissenschaften, 59:158164.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, 51:55475558.Google Scholar
Palacios-Fest, M. R. 1996. Geoquímica de la concha de ostrácodos (Limnocythere staplini): Un método de regresión múltiple como indicador paleoclimático. GEOS, 16:130136.Google Scholar
Palacios-Fest, M. R., Carreno, A. L., Ortega-Ramirez, J. R., and Alvarado-Valdez, G. 2002. A paleoenvironmental reconstruction of Laguna Babicora, Chihuahua, Mexico based on ostracode paleoecology and trace element shell chemistry. Journal of Paleolimnology, 27:185206.Google Scholar
Palacios-Fest, M. R., and Dettman, D. L. 2001. Temperature controls monthly variation in Ostracode valve Mg/Ca: Cypridopsis vidua from a small lake in Sonora, Mexico. Geochimica et Cosmochimica Acta, 65:24992507.Google Scholar
Park, L. E., Cohen, A. S., Martens, K., and Bralek, R. 2003. The impact of taphonomic processes on interpreting paleoecological changes in large lake ecosystems: Ostracodes in lakes Tanganyika and Malawi. Journal of Paleolimnology, in press.Google Scholar
Pedone, V. A., Olsen, J. A., and Hemming, G. 2003. Paleohydrologic history of the transgressive of lake Bonneville defined by secular variation in Sr isotopes. 3rd International Limnogeology Congress, Tucson, Arizona, March 29-April 2, 2003:212 (abstract).Google Scholar
Pietras, J. T., Carroll, A. R., Rhodes, M. K., Johnson, C. M., and Beard, B. L. 2003. Stratigraphic and strontium geochemical evidence of lacustrine basin evolution: example from the Green River Formation, Wyoming. 3rd International Limnogeology Congress, Tucson, Arizona, March 29-April 2, 2003:214 (abstract).Google Scholar
Rathburn, A. E., and De Deckker, P. 1997. Magnesium and strontium compositions of Recent benthic Foraminifera from the Coral Sea, Australia and Prydz Bay, Antarctica. Marine Micropaleontology, 32:231248.Google Scholar
Ricketts, R. D., Johnson, T. C., Brown, E. T., Rasmussen, K. A., and Romanovsky, V. V. 2001. The Holocene paleolimnology of Lake Issyk-Kul, Kyrgyzstan: trace element and stable isotope composition of ostracodes. Palaeogeography Palaeoclimatology Palaeoecology, 176:207227.Google Scholar
Rosenfeld, A. 1982. The secretion process of the ostracode carapace, p. 1224 In Bate, R. H., Robinson, E., and Sheppard, L. M. (eds.), Fossil and Recent Ostracods. Ellis Horwood Ltd., Chichester.Google Scholar
Rozanski, K., Araguas-Araguas, 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. American Geophysical Union.Google Scholar
Schornikov, E. I. 1980. Ostracodes in terrestrial habitats (biotopes). Zoologicheski Journal, 59:13061319 (in Russian).Google Scholar
Schrag, D., and Depaolo, D. 1993. Determination of δ18O of seawater in deep ocean during the last glacial maximum. Paleoceanography, 8:16.Google Scholar
Sebacher, D. I., Harriss, R. C., and Bartlett, K. B. 1985. Methane emissions to the atmosphere through aquatic plants. Journal of Environmental Quality, 14:4046.Google Scholar
Shackleton, N. J., and Opdyke, N. D. 1973. Oxygen isotope and paleomagmetic stratigraphy of equatorial Pacific core V28–238; oxygen isotope temperatures and ice volumes on a 105 and 106 year scale. Quaternary Research, 3:3955.Google Scholar
Shanley, J. B., Kendall, C., Pendall, E., Stevens, L. R., Michel, R. L., Phillips, P. J., Forester, R. M., Naftz, D. L., Liu, B., Stern, L., Wolfe, B. B., Chamberlain, P., Leavitt, S. W., Heaton, T., Mayer, B., Cecil, L. D., Lyons, W. B., Katz, B. G., Betancourt, J., Mcknight, D. M., Blum, J. D., Edwards, T. W. D., House, H. R., Ito, E., Aravena, R., and Whelan, J. F. 1998. Isotopes as Indicators of Environmental Change, p. 762816 In Kendall, C. and Mcdonnell, J. J. (eds.), Isotope Tracers in Catchment Hydrology. Elsevier Science B.V., Amsterdam.Google Scholar
Siegenthaler, U., and Eicher, U. 1986. Stable oxygen and carbon isotope analyses, p. 407422 In Berglund, B. E. (ed.), Handbook of Holocene Palaeoecology and Palaeohydrology. J. Wiley, New York.Google Scholar
Smith, A. J. 1993. Ostracodes as hydrochemical indicators in lakes of the northcentral United States. Journal of Paleolimnology, 8:121134.Google Scholar
Smith, A. J., Donovan, J., Ito, E., and Engstrom, D. R. 1997. Groundwater processes controlling prairie lake response to mid-Holocene drought. Geology, 25:391394.Google Scholar
Smith, A. J., Donovan, J. J., Ito, E., Engstrom, D. R., and Panek, V. A. 2002. Climate-driven hydrologic transients in lake sediment records: multiproxy record of mid-Holocene drought. Quaternary Science Reviews, 21:625646.Google Scholar
Smith, T. M., and Bate, R. H. 1983. The shell of the ostracod Halocypris inflata (Dana, 1849) examined by the ion bean etch technique. Journal of Micropalaeontology, 2:105110.Google Scholar
Sohn, I. G., and Kornicker, L. S. 1988. Ultrastructure of Myodocopid shells, p. 243258 In Hanai, T., Ikeya, N., and Ishizaki, K. (eds.), Evolutionary Biology of Ostracoda: Its Fundamentals and Applications. Proceedings of the Ninth International Symposium on Ostracoda, held in Shizuoka, Japan, 29 July-2 August 1985. Volume 11. Co-published by Kodansha and Elsevier, Amsterdam.Google Scholar
Stein, R., Schubert, C., Vogt, C., and Futterer, D. 1994. Stable isotope stratigraphy, sedimentation rates, and salinity changes in the Latest Pleistocene to Holocene eastern central Arctic Ocean. Marine Geology, 119:333355.Google Scholar
Stoll, H. M., and Schrag, D. P. 1998. Effects of Quaternary sea level changes on strontium in sea water. Geochimica et Cosmochimica Acta, 62:11071118.Google Scholar
Surge, D. M., and Lohmann, K. C. 2002. Temporal and spatial differences in salinity and water chemistry in SW Florida estuaries: Effects of human-impacted watersheds. Estuaries, 25:393408.Google Scholar
Swart, P. K., Burns, S. J., and Leder, J. J. 1991. Fractionation of the stable isotopes of oxygen and carbon in carbon dioxide during the reaction of calcite with phosphoric acid as a function of temperature and technique. Chemical Geology (Isotope Geoscience Section), 86:8996.Google Scholar
Swart, P. K., and Price, K. 2002. Origin of salinity variations in Florida Bay. Limnology & Oceanography, 47:12341241.Google Scholar
Talbot, M. R., Brendeland, K.-I., Russell, J. M., and Laerdal, T. 2003. Using Sr-isotopes to trace changes in basin configuration:the Holocene Sr-isotope stratigraphy of Lake Edwards, Uganda/Congo. 3rd International Limnogeology Congress, Tucson, Arizona, March 29-April 2, 2003:291 (abstract).Google Scholar
Talbot, M. R., and Kelts, K. 1986. Primary and diagenetic carbonates in the anoxic sediments of Lake Bosumtwi, Ghana. Geology, 14:912916.Google Scholar
Talbot, M. R., and Kelts, K. 1990. Palaeolimnological signatures from carbon and oxygen isotopic ratios in carbonates from organic carbon-rich lacustríne sediments, p. 99112 In Katz, B. J. and Rosendahl, B. R. (eds.), Lacustrine exploration: Case studies and modern analogues. American Association of Petroleum Geologists (AAPG).Google Scholar
Talbot, M. R., Williams, M. A. J., and Adamson, D. A. 2000. Strontium isotope evidence for late Pleistocene reestablishment of an integrated Nile drainage network. Geology, 28:343346.2.0.CO;2>CrossRefGoogle Scholar
Taylor, L. C. 1992. The response of spring-dwelling ostracides to intra-regional differences in groundwater chemistry with road salting practices in southern Ontario: a test using urban-rural transect. , University of Toronto, Toronto, 222 p.Google Scholar
Ullman, W. J., and Collerson, K. D. 1994. The Sr-isotope record of Late Quaternary hydrologic changes around Lake Frome, South-Australia. Australian Journal of Earth Sciences, 41:3745.Google Scholar
Van Doninck, K., Schön, I., Martens, K., and Goddeeris, B. in press. The life-cycle of ancient asexual ostracod Darwinula stevensoni (Brady & Robertson, 1870) (Crustacea, Ostracoda). Hydrobiologia.Google Scholar
Van Morkhoven, F. P. C. M. 1962. Post-Palaeozoic Ostracoda, Their Morphology, Taxonomy, and Economic Use. Elsevier Publishing Company, Amsterdam, Volume I, 204 p.Google Scholar
Veizer, J. 1983. Trace elements and isotopes in sedimentary carbonates., p. 265300 In Reeder, R. J. (ed.), Carbonates: Mineralogy and Chemistry. Reviews in Mineralogy, Volume 11. Mineralogical Society of America, Washington, D. C. Google Scholar
Von Grafenstein, U., Erlenkeuser, H., Müller, J., and Kleinmann-Eisenmann, A. 1992. Oxygen isotope records of benthic ostracods in Bavarian lake sediments. Naturwissenschaften, 79:145152.Google Scholar
Von Grafenstein, U., Erlernkeuser, H., and Trimborn, P. 1999. Oxygen and carbon isotopes in modern freshwater ostracod valves: assessing vital offsets and autecological effects of interest for palaeoclimate studies. Palaeogeography Palaeoclimatology Palaeoecology, 148:133152.Google Scholar
Wachter, E. A., and Hayes, J. M. 1985. Exchange of oxygen isotopes in carbon-dioxide-phosphoric acid systems. Chemical Geology (Isotope Geoscience Section), 52:365374.Google Scholar
Walker, I. R., Levesque, A. J., Cwynar, L. C., and Lotter, A. F. 1997. An expanded surface-water palaeotemperature inference model for use with fossil midges from eastern Canada. Journal of Paleolimnology, 18:165178.Google Scholar
Wansard, G., De Deckker, P., and Julià, R. 1998. Variability in ostracod partition coefficients D(Sr) and D(Mg): Implications for lacustrine palaeoenvironmental reconstructions. Chemical Geology, 146:3954.Google Scholar
Weiner, S., and Addadi, L. 2002. Calcium carbonateformation in biology: the involvement of an amorphous calcium carbonate precursor phase. Geochimica et Cosmochimica Acta, 66:A827 (abstract).Google Scholar
Whatley, R. C. 1988. Population structure of ostracods: some general principles for the recognition of palaeoenvironments., p. 245256 In De Deckker, P., Colin, J.-P., and Peypouquet, J.-P. (eds.), Ostracoda in the Earth Sciences. Elsevier, Amsterdam.Google Scholar
Whiticar, M. J., Faber, E., and Schoell, M. 1986. Biogenic methane formation in marine and freshwater environments: CO2 reduction vs acetate fermetation-isotope evidence. Geochimica et Cosmochimica Acta, 50:693709.Google Scholar
Whittaker, R. H., Levin, S. A., and Root, R. B. 1973. Niche, Habitat, and Ecotope. American Naturalist, 107:321338.Google Scholar
Wilson, J. O., Crill, P. M., Bartlett, K. B., Sebacher, D. I., Harriss, R. C., and Sas, R. L. 1989. Seasonal variation of methane emissions from a temperate swamp. Biogeochemistry, 8:5571.Google Scholar
Xia, J., Engstrom, D. R., and Ito, E. 1997a. Geochemistry of ostracode calcite: 2. the effects of water chemistry and seasonal temperature variation on Candona rawsoni . Geochimica et Cosmochimica Acta, 61:383391.Google Scholar
Xia, J., Haskell, B. J., Engstrom, D. R., and Ito, E. 1997b. Holocene climate reconstructed from tandem trace-element and stable-isotope composition of ostracodes from Coldwater Lake, North Dakota, U.S.A. Journal of Paleolimnology, 17:85100.Google Scholar
Xia, J., Ito, E., and Engstrom, D. R. 1997c. Geochemistry of ostracode calcite: 1. An experimental determination of oxygen isotope fractionation. Geochimica et Cosmochimica Acta, 61:377382.Google Scholar
Yu, Z., and Ito, E. 1999. Possible solar forcing of century-scale drought frequency in the northern Great Plains. Geology, 27:263266.Google Scholar
Yu, Z., Ito, E., Engstrom, D. R., and Fritz, S. C. 2002. A 2100-year trace-element and stable-isotope record at decadal resolution from Rice Lake in the northern Great Plains, USA. Holocene, 12:605617.Google Scholar