Hostname: page-component-76fb5796d-r6qrq Total loading time: 0 Render date: 2024-04-26T04:19:35.973Z Has data issue: false hasContentIssue false

Thermal potentiation and mineralogical evolution in the Bivalvia (Mollusca)

Published online by Cambridge University Press:  14 July 2015

Joseph G. Carter
Affiliation:
1Department of Geology, University of North Carolina at Chapel Hill 27599-3315
Enriqueta Barrera
Affiliation:
2Department of Geology, University of Akron, Akron, Ohio 44325
Michael J. S. Tevesz
Affiliation:
3Department of Biological, Geological, and Environmental Sciences, Cleveland State University, Cleveland, Ohio 44115

Abstract

The most important factor controlling the timing of Phanerozoic mineralogical evolution in the Bivalvia appears to be thermal potentiation of calcite deposition in colder marine and estuarine environments. Cold temperature has promoted mineralogical evolution in the Bivalvia by kinetically facilitating (potentiating) initially weak biological controls for calcite, thereby exposing their genetic basis to natural selection. Calcite has evolved in bivalve shells for a variety of selective advantages, including resistance to dissolution; resistance to chemical boring by algae and gastropods; reduced shell density in swimming and soft-bottom reclining species; enhanced flexibility in simple prismatic shell layers; and fracture localization and economy of secretion in association with certain foliated structures.

Endogenous calcite in bivalve shells varies from biologically induced to weakly and strongly biologically controlled. Biologically controlled calcite generally first appears in bivalve shells as an impersistent component of the outer shell layer, only later, in some groups, expanding to include the entire outer and then part or all of the middle and inner shell layers. The initial stages of mineralogical evolution are shown by certain modern Mytilidae, Veneridae and Petricolidae. In the latter two families, the calcite occurs as conellae in the outer part of the outer shell layer. Calcitic conellae in the inner shell layer of Pliocene Mercenaria are not barnacle plates, as previously indicated, but endogenous calcite comparable in origin to other venerid conellae. Their occurrence in Mercenaria may reflect thermal potentiation of weak biological controls for calcite, as well as local detachment of the secretory mantle epithelium near the pallial and adductor musculature.

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

Abolins-Krogis, A. 1968. Shell regeneration in Helix pomatia with special reference to the elementary calcifying particles. Symposium, Zoological Society of London, 22:7592.Google Scholar
Al-Aasm, I. S., and Veizer, J. 1986. Diagenetic stabilization of aragonite and Low-Mg calcite, I. Trace elements in rudists. Journal of Sedimentary Petrology, 56:138152.Google Scholar
Arkell, W. J., Kummel, B., and Wright, C. W. 1957. Mesozoic Ammonoidea, p. L80L490. In Moore, R. C. (ed.), Treatise on Invertebrate Paleontology, Part L, Mollusca 4, Cephalopoda, Ammonoidea. Geological Society of America and University of Kansas Press, Lawrence.Google Scholar
Bandel, K., and Hemleben, C. 1975. Anorganisches Kristallwachstum bei lebenden Mollusken. Paläontologische Zeitschrift, 49:298320.CrossRefGoogle Scholar
Begg, J. G., and Campbell, H. J. 1985. Etalia, a new Middle Triassic (Anisian) bivalve from New Zealand, and its relationship with other pteriomorphs. New Zealand Journal of Geology and Geophysics, 28:725741.CrossRefGoogle Scholar
Bernard, F. R. 1976. Living Chamidae of the eastern Pacific (Bivalvia: Heterodonta). Natural History Museum of Los Angeles County Contributions in Science, 278:143.Google Scholar
Berner, R. A. 1971. Principles of Chemical Sedimentology. McGraw-Hill, New York, 240 p.Google Scholar
Berner, R. A. 1975. The role of magnesium in the crystal growth of calcite and aragonite from sea water. Geochimica et Cosmochimica Acta, 39:489504.CrossRefGoogle Scholar
Bischoff, W. D., and Burke, C. D. 1991. Phanerozoic carbonate skeletal mineralogy and atmospheric carbon dioxide, p. 261274. In Schneider, S. H. and Boston, P. J. (eds.), Scientists on Gaia. The MIT Press, Cambridge, Massachusetts.Google Scholar
Burton, E. A., and Walter, L. M. 1987. Relative precipitation rates of aragonite and Mg calcite from seawater: temperature or carbonate ion control? Geology, 15:111114.2.0.CO;2>CrossRefGoogle Scholar
Cairns, S. D., and MacIntyre, I. G. 1992. Phylogenetic implications of calcium carbonate mineralogy in the Stylasteridae (Cnidaria: Hydrozoa). Palaios, 7:96107.CrossRefGoogle Scholar
Campbell, D. C., Hoekstra, K. T., and Carter, J. G. In press. 18s ribosomal DNA and evolutionary relationships within the Bivalvia. In Johnston, P. A. and Haggart, J. (eds.), The Bivalvia: Half a Billion Years of Evolution—Essays in Honor of Norman D. Newell. First International Symposium on Bivalve Evolution, Proceedings, University of Calgary Press, Canada.Google Scholar
Carter, J. G. 1980a. Environmental and biological controls of bivalve shell mineralogy and microstructure, p. 69113. In Rhoads, D. C., and Lutz, R. A. (eds.), Skeletal Growth of Aquatic Organisms. Plenum Press, New York.CrossRefGoogle Scholar
Carter, J. G. 1980b. Guide to bivalve shell microstructures, p. 645673. In Rhoads, D. C., and Lutz, R. A. (eds.), Skeletal Growth of Aquatic Organisms. Plenum Press, New York.Google Scholar
Carter, J. G. 1990. Evolutionary significance of shell microstructure in the Palaeotaxodonta, Pteriomorphia and Isofilibranchia (Bivalvia: Mollusca), p. 135296. In Carter, J. G. (ed.), Skeletal Biomineralization: Patterns, Processes and Evolutionary Trends, Volume I. Van Nostrand Reinhold, New York.Google Scholar
Carter, J. G., and Ambrose, W. W. 1989. Techniques for studying molluscan shell microstructure, p. 101119. In Feldman, R. M., Chapman, R. E., and Hannibal, J. T. (eds.), Paleotechniques. The Paleontological Society, Special Publication, 4, 358 p.Google Scholar
Carter, J. G. and Lutz, R. A. 1990. Part 2, Bivalvia (Mollusca), Plates 1-121, p. 528. In Carter, J. G. (ed.), Skeletal Biomineralization: Patterns, Processes and Evolutionary Trends, Volume II, Atlas. Van Nostrand Reinhold, New York.Google Scholar
Carter, J. G. and Seed, R. In press. Thermal potentiation and mineralogical evolution in Mytilus. In Johnston, P. A. and Haggart, J. (eds.), The Bivalvia: Half a Billion Years of Evolution—Essays in Honor of Norman D. Newell. First International Symposium on Bivalve Evolution, Proceedings, University of Calgary Press, Canada.Google Scholar
Carter, J. G., Bandel, K., de Buffrénil, V., Carlson, S. J., Castanet, J., Crenshaw, M. A., Dalingwater, J. E., Francillon-Vieillot, H., Géraudie, J., Meunier, F. J., Mutvei, H., de Ricqles, A., Sire, J. Y., Smith, A. B., Wendt, J., Williams, A., and Zylberberg, L. 1990. Glossary of Skeletal Biomineralization, p. 609671. In Carter, J. G. (ed.), Skeletal Biomineralization, Patterns, Processes and Evolutionary Trends, v. 1. Van Nostrand Reinhold, New York.Google Scholar
Carter, J. G., Rossbach, T. J., Robertson, K. J., and Ward, L. W. 1994. Morphological and microstructural evidence for the origin and early evolution of Ecphora (Mollusca, Gastropoda). Journal of Paleontology, 68:905907.CrossRefGoogle Scholar
Chinzei, K. 1995. Adaptive significance of the lightweight shell structure in soft bottom oysters. Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen, 195:217227.CrossRefGoogle Scholar
Cope, J. C. W. 1996. Early Ordovician (Arenig) bivalves from the Llangynog inlier, South Wales. Palaeontology, 39:9791025.Google Scholar
Cox, L. R. 1969. Family Bakevelliidae, p. N306N310. In Cox, L. R. (ed.), Treatise on Invertebrate Paleontology, Part N, Volume 1, Mollusca 6. Geological Society of America and University of Kansas Press, Lawrence.Google Scholar
Cox, L. R., and Newell, N. D. 1969. Family Oxytomidae, p. N344346. In Cox, L. R. (ed.), Treatise on Invertebrate Paleontology, Part N, Volume 1, Mollusca 6. Geological Society of America and University of Kansas Press, Lawrence.Google Scholar
Crenshaw, M. A. 1972. The inorganic composition of molluscan extrapallial fluid. Biological Bulletin, Marine Biological Laboratory, Woods Hole, Massachusetts, 143:506512.CrossRefGoogle ScholarPubMed
Crowley, T. J. 1994. Pangean climates, p. 2539. In Klein, G. D. (ed.), Pangea: Paleoclimate, Tectonics, and Sedimentation during Accretion, Zenith, and Breakup of a Supercontinent, The Geological Society of America Special Paper 288.Google Scholar
Dakin, W. J. 1928. The anatomy and phylogeny of Spondylus, with a particular reference to the lamellibranch nervous system. Proceedings of the Royal Society of London, series B, 103:337365.Google Scholar
Dauphin, Y., Cuif, J. P., and Denis, A. 1989. Mineralogy, chemistry and ultrastructure of the external shell-layer in ten species of Haliotis with reference to Haliotis tuberculata (Mollusca: Archaeogastropoda). Bulletin of the Geological Institutions of the University of Uppsala, N.S., 15:738.Google Scholar
Dechaseaux, C. 1939. Megalodon, Pachyerisma, Protodiceras, Diceras, Pterocardium et l'origine des Diceras. Bulletin de la Société Géologique de France, ser. 5, 9:207218.Google Scholar
de Renzi, M., and Márquez-Aliaga, A. 1980. Primary and diagenetic features in the microstructure of some Triassic bivalves. Revista del Instituto de Investigaciones Geológicas, Diputación Provincial, Universidad de Barcelona, 34:101116.Google Scholar
Dodd, J. R. 1963. Palaeoecological implications of shell mineralogy in two pelecypod species. Journal of Geology, 71:111.CrossRefGoogle Scholar
Dodd, J. R. 1964. Environmentally controlled variation in the shell structure of a pelecypod species. Journal of Paleontology, 38:10651071.Google Scholar
Duckworth, D. L. 1976. A model for the physiological control of mineralogy and trace element composition of biogenic carbonate. Geological Society of America Abstracts, 1976:844.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, 64:13151326.CrossRefGoogle Scholar
Falini, G., Albeck, S., Weiner, S., and Addadi, L. 1996. Control of aragonite or calcite polymorphism by mollusk shell macromolecules. Science, 271:6769.CrossRefGoogle Scholar
Feigl, F. 1937. Qualitative Analysis by Spot Test. Nordemann Publishing Company, New York, 400 p.Google Scholar
Fischer-Piette, É., and Métivier, B. 1971. Révision des Tapetinae (Mollusques Bivalves). Mémoires du Muséum National d'Histoire Naturelle, série A, Zoologie, 71:1106, +16 pls.Google Scholar
François, L. M., and Gerard, J.-C. 1986. A numerical model of the evolution of ocean sulfate and sedimentary sulfur during the last 800 million years. Geochimica et Cosmochimica Acta, 50:22892302.CrossRefGoogle Scholar
Freeman, T. 1980. Comment on “Biomineralization, paleooceanography and the evolution of calcareous marine organisms. Geology, 8:265266.2.0.CO;2>CrossRefGoogle Scholar
Fyfe, W. S. and Bischoff, J. L. 1965. The calcite-aragonite problem. Society of Economic Paleontologists and Mineralogists Special Publication, 13:313.Google Scholar
Grant-Mackie, J. A. 1980. Mode of life and adaptive evolution in the cosmopolitan Triassic bivalve Monotis. Journal of the Malacological Society of Australia, 4:223.Google Scholar
Grossman, E. L., and Ku, T. 1986. Oxygen and carbon isotope fractionation in biogenic aragonite: temperature effects. Chemical Geology (Isotope Geoscience Section), 59:5974.CrossRefGoogle Scholar
Harper, E. M. 1994. Molluscivory by the forcipulate asteroid Coscinasterias acutispina (Stimpson), p. 405425. In Morton, B. (ed.), The Malacofauna of Hong Kong and Southern China III, Hong Kong University Press, Hong Kong.Google Scholar
Harper, E. M., and Skelton, P. W. 1993a. The Mesozoic marine revolution and epifaunal bivalves. Scripta Geologica Special Issue, 2:127153.Google Scholar
Harper, E. M., and Skelton, P. W. 1993b. A defensive value of the thickened periostracum in the Mytiloidea. The Veliger, 36:3642.Google Scholar
Harper, E. M., Palmer, T. J., and Alphey, J. R. 1997. Evolutionary response by bivalves to changing Phanerozoic sea-water chemistry. Geological Magazine, 134:403407.CrossRefGoogle Scholar
Hölder, H. 1952. Über Gehäusebau, insbesondere hohlkie Jurassischer Ammoniten. Palaeontographica, Abt. A, 102:1848.Google Scholar
Holland, H. D. 1965. The history of ocean water and its effect on the chemistry of the atmosphere. Proceedings of the National Academy of Sciences, 53:11731182.CrossRefGoogle Scholar
Hopkins, S. H., Anderson, J. W., and Horvath, K. 1974. Biology of the clam Rangia cuneata: what we know and what it means. Proceedings of the National Shellfisheries Association, 64:4.Google Scholar
Ichikawa, K. 1958. Zur Taxonomie und Phylogenie der triadischen “Pteriidae” (Lamellibranch.) mit besonderer Berücksichtigung der Gattung Claraia, Eumorphotis, Oxytoma, und Monotis. Palaeontographica, Abteilung A, 111:131212.Google Scholar
Isozaki, Y. 1997. Permo-Triassic boundary superanoxia and stratified superocean: records from lost deep sea. Science, 276:235238.CrossRefGoogle ScholarPubMed
Jaccarini, V., Bannister, W. H., and Micallef, H. 1968. The pallial glands and rock boring in Lithophaga lithophaga (Lamellibranchia: Mytilidae). Journal of Zoology, 154:397401.CrossRefGoogle Scholar
Johnson, C. C., and Kauffman, E. G. 1987. Adaptive radiation of rudistids in Cretaceous carbonate platform environments. Program with Abstracts, L. J. Chubb Centennial Meeting, Kingston, Jamaica, p. 11.Google Scholar
Johnston, P. A. 1991. Systematics and ontogeny of a new bivalve, Umburra cinefacta, from the Silurian of Australia: implications for Pteriomorphian evolution. Alcheringa, 15:293319.CrossRefGoogle Scholar
Johnston, P. A. 1995. Evolutionary pathways in the bivalve subclass Pteriomorphia. Canadian Paleontology Conference Programs and Abstracts, 5:1718.Google Scholar
Jux, U., and Omara, S. 1983. Pachypteria sinaitica n. sp.,—eine aufgewachsene, austernähnliche Muschel aus dem Unterkarbon ägyptens. Paläontologische Zeitschrift, 57:7991.Google Scholar
Karhu, J., and Epstein, S. 1986. The implication of the oxygen isotope records in coexisting cherts and phosphates. Geochimica et Cosmochimica Acta, 50:17451756.CrossRefGoogle Scholar
Kauffman, E. G., and Johnson, C. C. 1988. The morphological and ecological evolution of Middle and Upper Cretaceous reef-building rudistids. Palaios, 3:194216.CrossRefGoogle Scholar
Kennedy, W. J., and Taylor, J. D. 1968. Aragonite in rudists. Proceedings of the Geological Society of London, 1645:325331.Google Scholar
Kennedy, W. J., Taylor, J. D., and Hall, A. 1969. Environmental and biological controls on bivalve shell mineralogy. Biological Reviews of the Cambridge Philosophical Society, 44:499530.CrossRefGoogle ScholarPubMed
Kennedy, W. J., Morris, N. J., and Taylor, J. D. 1970. The shell structure, mineralogy and relationships of the Chamacea (Bivalvia). Palaeontology, 13:379413.Google Scholar
Kitano, Y., Kanamori, N., and Yoshioka, S. 1976. Influence of chemical species on the crystal type of calcium carbonate, p. 191202. In Watabe, N., and Wilbur, K. M. (eds.), The Mechanisms of Mineralization in the Invertebrates and Plants. University of South Carolina Press, Columbia.Google Scholar
Korringa, P. 1951. On the nature and function of “chalky” deposits in the shell of Ostrea edulis. Proceedings of the California Academy of Sciences, Series 4, 27:133158.Google Scholar
Kutzbach, J. E. 1994. Idealized Pangean climates: sensitivity to orbital change, p. 4155. In Klein, G. D. (ed.), Pangea: Paleoclimate, Tectonics, and Sedimentation during Accretion, Zenith, and Breakup of a Supercontinent, The Geological Society of America Special Paper 288.Google Scholar
Lamprell, K., and Whitehead, T. 1992. Bivalves of Australia. Volume 1. Crawford House Press, Bathurst, 182 p.Google Scholar
Lorens, R. B., and Bender, M. L. 1977. Physiological exclusion of magnesium from Mytilus edulis calcite. Nature, 269:793794.CrossRefGoogle Scholar
Lorens, R. B., and Bender, M. L. 1980. The impact of solution chemistry on Mytilus edulis calcite and aragonite. Geochimica et Cosmochimica Acta, 44:12651278.CrossRefGoogle Scholar
Lowenstam, H. A. 1954a. Environmental relations of modification compositions of certain carbonate secreting marine invertebrates. National Academy of Sciences, U.S.A., Proceedings, 40:3948.Google ScholarPubMed
Lowenstam, H. A. 1954b. Factors affecting the aragonite-calcite ratios in carbonatesecreting marine organisms. Journal of Geology, 62:284322.CrossRefGoogle Scholar
Lowenstam, H. A. 1954c. Status of invertebrate paleontology, 1953, XI. Systematic, paleoecologic and evolutionary aspects of skeletal building materials. Bulletin of the Museum of Comparative Zoology, Harvard College, 112:287317.Google Scholar
Lowenstam, H. A. 1981. Minerals formed by organisms. Science, 211:11261131.CrossRefGoogle ScholarPubMed
Lowenstam, H. A., and Weiner, S. 1989. On Biomineralization. Oxford University Press, New York, 324 p.CrossRefGoogle Scholar
Macgillavry, H. J. 1937. Geology of the province of Camaguey, Cuba, with revisional studies in rudist paleontology. Geographische en Geologische Mededelingen, Physiographisch Geologische Reeks, Geographisch Institut (Utrecht), Thése 14:1168.Google Scholar
Mackenzie, F. T., and Pigott, J. D. 1981. Tectonic controls of Phanerozoic sedimentary rock cycling. Geological Society of London, Journal, 138:183196.CrossRefGoogle Scholar
Mackenzie, F. T., and Agegian, C. R. 1986. Biomineralization and tentative links to plate tectonics, p. 1127. In Crick, R. E. (ed.), Origin, Evolution and Modern Aspects of Biomineralization in Plants and Animals. Plenum Press, New York.Google Scholar
McRoberts, C. A., and Carter, J. G. 1994. Nacre in an early gryphaeid bivalve (Mollusca). Journal of Paleontology, 68:14051408.CrossRefGoogle Scholar
Malchus, N. 1990. Revision der Kreide-Austern (Bivalvia: Pteriomorphia) Ägyptens (Biostratigraphie, Systematik). Berliner Geowissenschaftliche Abhandlungen, Reihe A, 125:1231.Google Scholar
Mann, S. 1983. Mineralization in biological systems. Structure and Bonding, 54:125174.CrossRefGoogle Scholar
Morse, J. W. 1983. The kinetics of calcium carbonate dissolution and precipitation. Reviews in Mineralogy, 11:227264.Google Scholar
Mucci, A. 1983. The solubility of calcite and aragonite in seawater at various salinities, temperatures and one atmosphere total pressure. American Journal of Science, 283:780799.CrossRefGoogle Scholar
Mutvei, H. 1986. Structure of molluscan prismatic shell layers, p. 137151. In Crick, R. E. (ed.), Origin, Evolution and Modern Aspects of Biomineralization in Plants and Animals, Plenum Press, New York.Google Scholar
Mutvei, H., Dauphin, Y., and Cuif, J.-P. 1985. Observations sur l'organisation de la couche externe du test des Haliotis (Gastropoda): un cas exceptionnel de variabilité minéralogique et microstructurale. Bulleti. du Muséum Nationale d'Histoire Naturelle, Paris, série 4, 7(A):7391.CrossRefGoogle Scholar
Newell, N. D. 1942. Late Paleozoic pelecypods: Mytilacea. State Geological Survey of Kansas [Report], 10(2):180.Google Scholar
Newell, N. D., and Boyd, D. W. 1970. Oyster-like Permian Bivalvia. American Museum of Natural History, Bulletin, 143:217282.Google Scholar
Newell, N. D., and Boyd, D. W. 1989. Phylogenetic implications of shell microstructure in the Pseudomonotidae, extinct Bivalvia. American Museum Novitates, Number 2933, 12 p.Google Scholar
Palmer, A. R. 1992. Calcification in marine molluscs: how costly is it? Proceedings of the National Academy of Sciences, U.S.A., 89:13791382.CrossRefGoogle Scholar
Palmer, T. J., Hudson, J. D., and Wilson, M. A. 1988. Palaeoecological evidence for early aragonite dissolution in ancient calcite seas. Nature, 335:809810.CrossRefGoogle Scholar
Pareyn, C., Termier, G., and Termier, H. 1971. Les bivalves ostréiformes du Sahara: Société Geologique du Nord, Annales, 91:229238.Google Scholar
Railsback, L. B., and Anderson, T. F. 1987. Control of Triassic seawater chemistry and temperature on the evolution of post-Paleozoic aragonite-secreting faunas. Geology, 15:10021005.2.0.CO;2>CrossRefGoogle Scholar
Reitner, J. 1992. “Coralline Spongien”: Der Versuch einer phylogenetisch-taxonomischen Analyse. Berliner Geowissenschaftliche Abhandlungen, Reihe E, 1:1352.Google Scholar
Rubison, M., and Clayton, R. N. 1969. Carbon-13 fractionation between aragonite and calcite. Geochimica et Cosmochimica Acta, 33:9971002.CrossRefGoogle Scholar
Runnegar, B. 1985. Shell microstructures of Cambrian molluscs replicated by phosphate. Alcheringa, 9:245257.CrossRefGoogle Scholar
Sandberg, P. A. 1983. An oscillating trend in Phanerozoic non-skeletal carbonate mineralogy. Nature, London, 305:1922.CrossRefGoogle Scholar
Sandberg, P. A. 1985. Nonskeletal aragonite and pCO2 in the Phanerozoic and Proterozoic, p. 585594. In Sundquist, E. T. and Broecker, W. S. (eds.), The Carbon Cycle and Atmospheric CO2: Natural Variations Archean to Present. American Geophysical Union, Washington, D.C.Google Scholar
Scarlato, O. A., and Starobogatov, Y. I. 1984. The systematics of the suborder Mytileina (Bivalvia). Malacological Review, 17:115116.Google Scholar
Seilacher, A. 1972. Divaricate patterns in pelecypod shells. Lethaia, 5:325343.CrossRefGoogle Scholar
Senobari-Daryan, B. 1991. “Sphinctozoa”: an overview, p. 224241. In Reitner, J. and Keupp, H. (eds.), Fossil and Recent Sponges, Springer Verlag, Berlin.CrossRefGoogle Scholar
Siewert, W. 1972. Schalenbau und Stammesgeschichte von Austern. Stuttgarter Beiträge zur Naturkunde, series B (Geologie und Paläontologie), 1:157.Google Scholar
Skelton, P. W. 1974. Aragonitic shell structures in the rudist Biradiolites and some paleobiological inferences. Géologie Méditerranéenne, 1:6374.CrossRefGoogle Scholar
Skelton, P. W. 1976. Functional morphology of the Hippuritidae. Lethaia, 9:83100.CrossRefGoogle Scholar
Skelton, P. W. 1978. The evolution of functional design in rudists (Hippuritacea) and its taxonomic implications. Royal Society of London, Philosophical Transactions, series B, 284:305318.Google Scholar
Skelton, P. W. 1979. Preserved ligament in a radiolitid rudist bivalve and its implication of mantle marginal feeding in the group. Paleobiology, 5:90106.CrossRefGoogle Scholar
Stenzel, H. B. 1971. Volume 3, Oysters, p. N953N1224. In Cox, L. R., et al., Treatise on Invertebrate Paleontology, Part N, Mollusca 6. Geological Society of America and University of Kansas Press, Lawrence.Google Scholar
Sverdrup, H. U., Johnson, M. W., and Fleming, R. H. 1942. The Oceans. Prentice Hall, New Jersey, 1087 p.Google 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.CrossRefGoogle Scholar
Tate, R. 1887. Descriptions of some new species of South Australian marine and freshwater Mollusca. Transactions, Proceedings and Reports of the Royal Society of South Australia, 9:6275, pls. 4-5.Google Scholar
Taylor, J. D. 1973. The structural evolution of the bivalve shell. Palaeontology, 16:519534.Google Scholar
Taylor, J. D., and Kennedy, W. J. 1969. The shell structure and mineralogy of Chama pellucida Broderip. The Veliger, 11:391398.Google Scholar
Taylor, J. D., and Layman, M. 1972. The mechanical properties of bivalve (Mollusca) shell structures. Palaeontology, 15:7387.Google Scholar
Taylor, J. D., and Reid, D. G. 1990. Shell microstructure and mineralogy of the Littorinidae: ecological and evolutionary significance. Hydrobiologia, 193:199215.CrossRefGoogle Scholar
Taylor, J. D., Cleevely, R. J., and Morris, N. 1983. Predatory gastropods and their activities in the Blackdown Greensand (Albian) of England. Palaeontology, 26:521533.Google Scholar
Taylor, J. D., Kennedy, W. J., and Hall, A. 1969. The shell structure and mineralogy of the Bivalvia. Introduction. Nuculacea-Trigonacea. Bulletin of the British Museum (Natural History), Zoology, Supplement, 3:1125.Google Scholar
Taylor, J. D., Kennedy, W. J., and Hall, A. 1973. The shell structure and mineralogy of the Bivalvia. II. Lucinacea-Clavagellacea. Conclusions. Bulletin of the British Museum (Natural History), Zoology, 22:253294.CrossRefGoogle Scholar
Teichert, C. 1964. Morphology of hard parts, p. K13K59. In Moore, R. C. (ed.), Treatise on Invertebrate Paleontology, Part K, Mollusca 3. Geological Society of America and University of Kansas Press, Lawrence.Google Scholar
Termier, H., and Termier, G. 1972. Identification de fragments de coquilles de mollusques dans les sédiments Paléozoiques et Triasiques. Haliotis, 2:209214.Google Scholar
Van de Poel, H. M., and Schlager, W. 1994. Variations in Mesozoic-Cenozoic skeletal carbonate mineralogy. Geologie en Mijnbouw, 73:3151.Google Scholar
Vermeij, G. J. 1983. Traces and trends of predation, with special reference to bivalved animals. Palaeontology, 26:455465.Google Scholar
Vermeij, G. J. 1987. Evolution and Escalation: an Evolutionary History of Life. Princeton University Press, Princeton, New Jersey, 527 p.CrossRefGoogle Scholar
Wada, K., and Fujinuki, T. 1976. Biomineralization in bivalve molluscs with emphasis on the chemical composition of the extrapallial fluid, p. 175190. In Watabe, N. and Wilbur, K. M. (eds.), The Mechanisms of Mineralization in the Invertebrates and Plants. University of South Carolina Press, Columbia.Google Scholar
Waller, T. R. 1972. The functional significance of some shell microstructures in the Pectinacea (Mollusca: Bivalvia). Proceedings of the International Geological Congress, 24th Session, Montreal, Canada, Section 7, Paleontology, p. 4856.Google Scholar
Waller, T. R. 1978. Morphology, morphoclines and a new classification of the Pteriomorphia. Royal Society of London, Philosophical Transactions, Series B, 284:34365.Google Scholar
Walter, L. M., and Morse, J. W. 1984. Reactive surface area of skeletal carbonates during dissolution: effect of grain size. Journal of Sedimentary Petrology, 54:10811090.Google Scholar
Watabe, N. 1981. Crystal growth of calcium carbonate in the invertebrates. Progress in Crystal Growth and Characterization, 4:99147.CrossRefGoogle Scholar
Waterhouse, J. B. 1982. Permian Pectinacea and Limacea (Bivalvia) from New Zealand. New Zealand Geological Survey Paleontological Bulletin 49:175, 25 pls.Google Scholar
Weiner, S. 1979. Aspartic acid-rich proteins: major components of the soluble organic matrix of mollusk shells. Calcified Tissue International, 29:163167.CrossRefGoogle ScholarPubMed
Weiner, S. 1986. Organization of extracellularly mineralized tissues: a comparative study of biological crystal growth. Critical Reviews in Biochemistry, 20:365408.CrossRefGoogle ScholarPubMed
Wignall, P. B., and Twitchett, R. J. 1996. Oceanic anoxia and the end Permian mass extinction. Science, 272:11551158.CrossRefGoogle ScholarPubMed
Wilbur, K. M. 1964. Shell formation and regeneration, p. 243282. In Wilbur, K. M. and Yonge, C. M. (eds.), Physiology of Mollusca, Volume 1. Academic Press, New York.CrossRefGoogle Scholar
Wilbur, K. M. 1976. Recent studies on invertebrate mineralization, p. 79108. In Watabe, N. and Wilbur, K. M., (eds.), The Mechanisms of Mineralization in the Invertebrates and Plants. University of South Carolina Press, Columbia, South Carolina.Google Scholar
Wilbur, K. M., and Watabe, N. 1963. Experimental studies on calcification in molluscs and the alga Coccolithus huxleyi. Annals of the New York Academy of Sciences, 109:82112.CrossRefGoogle Scholar
Wilkinson, B. H. 1979. Biomineralization, paleooceanography and the evolution of calcareous marine organisms. Geology, 7:524527.2.0.CO;2>CrossRefGoogle Scholar
Wilkinson, B. H. 1980. Reply to Tom Freeman's comment on “Biomineralization, paleooceanography and the evolution of calcareous marine organisms.” Geology, 8:266267.Google Scholar
Wilkinson, B. H., and Given, R. K. 1986. Secular variation in abiotic marine carbonates: constraints on Phanerozoic atmospheric carbon dioxide contents and ocean Mg/Ca ratios. Journal of Geology, 94:321333.CrossRefGoogle Scholar
Wilkinson, B. H., Owen, R. M., and Carroll, A. R. 1985. Submarine hydrothermal weathering, global eustacy, and carbonate polymorphism in Phanerozoic marine oolites. Journal of Sedimentary Petrology, 55:171183.Google Scholar
Wilson, D. 1983. The Lee Creek enigma, McLellania aenigma, a new taxon in fossil Cirrhipedia, p. 483498. In Ray, C. E. (ed.), Geology and Paleontology of the Lee Creek Mine, North Carolina, I. Smithsonian Contributions to Paleobiology, 53:1529.CrossRefGoogle Scholar
Wilson, K. M., Pollard, D., Hay, W. W., Thompson, S. L., and Wold, C. N. 1994. General circulation model simulations of Triassic climates: preliminary results, p. 91116. In Klein, G. D. (ed.), Pangea: Paleoclimate, Tectonics, and Sedimentation during Accretion, Zenith, and Breakup of a Supercontinent, The Geological Society of America Special Paper 288.Google Scholar
Worsley, T. R., Moore, T. L., Fraticelli, C. M., and Scotese, C. R. 1994. Phanerozoic CO2 levels and global temperatures inferred from changing paleogeography, p. 5773. In Klein, G. D. (ed.), Pangea: Paleoclimate, Tectonics, and Sedimentation during Accretion, Zenith, and Breakup of a Supercontinent, The Geological Society of America Special Paper 288.CrossRefGoogle Scholar
Wray, J. L., and Daniels, F. 1957. Precipitation of calcite and aragonite. Journal of the American Chemical Society, 79(9):20312034.CrossRefGoogle Scholar
Yonge, C. M. 1973. Functional morphology with particular reference to hinge and ligament in Spondylus and Plicatula and a discussion on relations within the superfamily Pectinacea (Mollusca: Bivalvia). Philosophical Transactions of the Royal Society of London, Series B, 267:173208.Google Scholar
Yonge, C. M. 1981. On adaptive radiation in the Pectinacea with a description of Hemipecten forbesianus Adams and Reeve, 1849. Malacologia, 21:2324.Google Scholar