Hostname: page-component-848d4c4894-5nwft Total loading time: 0 Render date: 2024-06-06T08:22:15.185Z Has data issue: false hasContentIssue false

Paleoenvironmental analysis of Middle Jurassic (Callovian) ammonoids from Poland: trace elements and stable isotopes

Published online by Cambridge University Press:  19 May 2016

Uwe Brand*
Affiliation:
Department of Geological Sciences, Brock University, St. Catharines, Ontario L2S 3A1, Canada

Abstract

Mineralogical and microstructural analyses of Quenstedtoceras shells (Callovian, Poland) show that the outer shell material is preserved as nacreous aragonite, while the septal material shows different degrees of alteration. The septa in part are pyritized or show signs of cementation by equant/acicular calcite cement; paleoenvironmental interpretation of the least-altered outer shell material using the Sr/Na ratio suggests that the shell material was deposited in seawater which had a salinity of about 33 ppt and a range of about 32–35 ppt. Other data, such as Mn and Fe, imply that the oxygen level of the seawater was normal. The paleotemperature, using δ18O values with a salinity correction, was calculated to be about 11.0–14.5°C with an average water temperature of about 12.5°C. These paleoenvironmental parameters, in part, account for the faunal character of the Boreal Realm of the Jurassic.

Type
Research Article
Copyright
Copyright © 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

Al-Aasm, I. and Veizer, J. 1982. Chemical stabilization of low-Mg calcite: an example of brachiopods. Journal of Sedimentary Petrology, 52:11011109.Google Scholar
Brand, U. 1981. Mineralogy and chemistry of the lower Pennsylvanian Kendrick fauna, eastern Kentucky — 1: trace elements. Chemical Geology, 32:116.CrossRefGoogle Scholar
Brand, U. 1982. The oxygen and carbon isotope composition of Carboniferous fossil components: seawater effects. Sedimentology, 29:139147.CrossRefGoogle Scholar
Brand, U. 1983a. Mineralogy and chemistry of the lower Pennsylvanian Kendrick fauna, eastern Kentucky—3: diagenetic and paleoenvironmental analysis. Chemical Geology, 40:167181.Google Scholar
Brand, U. 1983b. Geochemical analysis of Nautilus pompilius from Fiji, South Pacific. Marine Geology, 53:M1–M5.Google Scholar
Brand, U. 1984a. A salinity equation: chemical evaluation of molluscan aragonite. Society of Economic Paleontologists and Mineralogists Midyear Meeting, Book of Abstracts, 1:16.Google Scholar
Brand, U. 1984b. Biochemical evolution of molluscan aragonite. Influence of seawater chemistry. Geological Association of Canada, Annual Meeting, London, May 14–16, 1984, Book of Abstracts, 9:48.Google Scholar
Brand, U. and Veizer, J. 1980. Chemical diagenesis of a multicomponent carbonate system — 1: trace elements. Journal of Sedimentary Petrology, 50:12191236.Google Scholar
Brand, U. and Veizer, J. 1981. Chemical diagenesis of a multicomponent carbonate system—2: stable isotopes. Journal of Sedimentary Petrology, 51:987997.Google Scholar
Craig, H. 1957. Isotopic standards for carbon and oxygen and correction factors for mass spectrometer analysis of carbon dioxide. Geochimica et Cosmochimica Acta, 12:133149.Google Scholar
Emiliani, C. 1955. Pleistocene temperatures. Journal of Geology, 63:538578.Google Scholar
Epstein, S. and Mayeda, T. 1953. Variations in O18 content of waters from natural sources. Geochimica et Cosmochimica Acta, 27:213224.Google Scholar
Fritz, P. 1965. O18/O16 Isotopenanalysen und Paleotemperatur Bestimmungen an Belemniten aus dem Schwäbischen Jura. Geologische Rundschau, 54:261269.Google Scholar
Hallam, A. 1975. Jurassic Environments. Cambridge University Press, Cambridge, 269 p.Google Scholar
Hudson, J. D. 1977. Stable isotopes and limestone lithification. Journal of the Geological Society, London, 133:637660.CrossRefGoogle Scholar
Jordan, R. and Stahl, W. 1970. Isotopische Paläotemperatur Bestimmungen an jurassischen Ammoniten and grundsätzliche Voraussetzungen für diese Methode. Geologisches Jahrbuch, 89:3362.Google Scholar
Kitano, Y., Ikumara, M. and Idogaki, M. 1975. Incorporation of sodium chloride and sulfate with calcium carbonate. Geochemical Journal (Japan), 9:7584.CrossRefGoogle Scholar
Kolesar, P. T. 1978. Magnesium in calcite from a coralline alga. Journal of Sedimentary Petrology, 48:815819.Google Scholar
Lowenstam, H. A. 1963. Biologic problems relating to the composition and diagenesis of sediments, p. 137195. In Donnelly, T. W. (ed.), The Earth Sciences, Problems and Progress in Current Research. University of Chicago Press, Chicago.Google Scholar
Makowski, H. 1952. La faune Callovienne de Lukow en Pologne. Palaeontologica Polonicia, 4, 64 p.Google Scholar
McCrea, J. M. 1950. On the isotope chemistry of carbonates and a paleotemperature scale. Journal of Chemical Physics, 18:849857.CrossRefGoogle Scholar
Milliman, J. D. 1974. Marine Carbonates. Springer-Verlag, Berlin, 375 p.Google Scholar
Morrison, J. O. and Brand, U. 1983. Chemical diagenesis of Pennsylvanian Brush Creek (Pennsylvania) carbonate components: trace elements (abstract). American Association of Petroleum Geologists, Bulletin, 67:520.Google Scholar
Morrison, J. O. and Brand, U. 1984. Secular and environmental variation of seawater: an example of brachiopod chemistry. Geological Association of Canada, Annual Meeting, London, May 14–16, 1984, Book of Abstracts, 9:91.Google Scholar
Pingitore, N. E. Jr. 1976. Vadose and phreatic diagenesis: processes, products and their recognition in corals. Journal of Sedimentary Petrology, 46:9851006.Google Scholar
Pingitore, N. E. Jr. 1982. The role of diffusion during carbonate diagenesis. Journal of Sedimentary Petrology, 52:2739.Google Scholar
Ragland, P. C., Pilkey, O. H. and Blackwelder, B. W. 1979. Diagenetic changes in the elemental composition of unrecrystallized mollusk shells. Chemical Geology, 25:123134.Google Scholar
Rubinson, M. and Clayton, R. N. 1969. Carbon-13 fractionation between aragonite and calcite. Geochimica et Cosmochimica Acta, 33:991002.Google Scholar
Shackleton, N. J. and Kennett, J. P. 1975. Paleotemperature history of the Cenozoic and the initiation of Antarctic glaciation: oxygen and carbon isotopic analysis in DSDP sites 277, 279, and 281, p. 743755. In Kennett, J. P. and Houtz, R. E. (eds.), Initial Reports of the Deep Sea Drilling Project. U.S. Government Printing Office, Washington.Google Scholar
Smith, A. G., Briden, J. C. and Drewry, G. E. 1973. Phanerozoic world maps. Special Papers in Palaeontology, 12:139.Google Scholar
Squires, R. L. 1973. Burial environment, diagenesis, mineralogy and Mg & Sr contents of skeletal carbonates in the Buckhorn asphalt of Middle Pennsylvanian age, Arbuckle Mountains, Oklahoma. Unpubl. Ph.D. Dissertation, California Institute of Technology, Pasadena, 184 p.Google Scholar
Stevens, G. R. 1971. Relationship of isotopic temperatures and faunal realms to Jurassic-Cretaceous paleogeography, particularly of the S.W. Pacific. Journal of the Royal Society, New Zealand, 1:145158.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, 14:829897.Google Scholar
Tarutani, T., Clayton, R. N. and Mayeda, T. K. 1969. The effect of polymorphism and magnesium substitution on oxygen isotope formation between calcium carbonate and water. Geochimica et Cosmochimica Acta, 33:982996.Google Scholar
Turner, J. B. 1982. Kinetic fractionation of Carbon-13 during calcium carbonate precipitation. Geochimica et Cosmochimica Acta, 46:11831191.Google Scholar
Veizer, J. and Fritz, P. 1976. Possible control of postdepositional alteration in oxygen paleotemperature determinations. Earth and Planetary Sciences Letters, 33:266280.Google Scholar
White, A. F. 1977. Sodium and potassium co-precipitation in aragonite. Geochimica et Cosmochimica Acta, 41:613625.CrossRefGoogle Scholar