Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-30T01:01:59.461Z Has data issue: false hasContentIssue false

Molluscan extinction patterns across the Cenomanian-Turonian Stage boundary in the western interior of the United States

Published online by Cambridge University Press:  08 April 2016

William P. Elder*
Affiliation:
Mail Stop 915, U.S. Geological Survey, 345 Middlefield Road, Menlo Park, California 94025

Abstract

High-resolution stratigraphic analysis of 18 sections spanning the Cenomanian–Turonian Stage boundary in the western interior of the United States has allowed determination of the magnitude and pattern of molluscan extinction and disruption. Composite range data from all sections show that the faunal turnover across the stage boundary occurs in a series of narrow stratigraphic zones, defined by multiple first and last occurrences, separated by intervals displaying little or no taxonomic turnover. Two of the apparent extinction steps (bottom and top of the Neocardioceras juddii Zone) may be intercontinentally developed. The additional steps apparently reflect cyclic changes in water mass and substrate characteristics in the western interior basin produced in response to orbital forcing of climate. An interval (ca. 10-100 k.y. duration) of changing community structure and general biotic deterioration is found below each of the two potentially intercontinentally developed extinction steps. The most affected mollusks were those having intercontinental distributions (ammonites and inoceramid bivalves), suggesting that disruption of planktotrophic larval dispersal may have played a role in increasing extinction and speciation rates near the C–T boundary. The nekto-benthic ammonites were affected earlier and to a greater degree than the pelagic forms, implying progressive upward expansion of the oxygen minimum zone preceding the stage boundary.

Type
Articles
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

Literature Cited

Allen, J. A. 1958. The basic form and adaptations to habitat in the Lucinacea (Eulamellibranchia). Philosophical Transactions of the Royal Society of London, Series B 241:421484.Google Scholar
Allmon, W. D. 1988. Ecology of Recent turritelline gastropods (Prosobranchia, Turritellidae): current knowledge and paleontological implications. Palaios 3:259284.CrossRefGoogle Scholar
Alvarez, W. 1986. Toward a theory of impact crisis. Eos 67:649656.Google Scholar
Alvarez, W., Kauffman, E. G., Surlyk, F., Alvarez, L. W., Asaro, F., and Michel, H. V. 1984. Impact theory of mass extinctions and the invertebrate fossil record. Science 233:11351141.Google Scholar
Alvarez, W., and Muller, R. A. 1984. Evidence from crater ages for periodic impacts on the Earth. Science 308:718720.Google Scholar
Arthur, M. A., Dean, W. E., and Pratt, L. M. 1988. Geochemical and climatic effects of increased marine organic carbon burial at the Cenomanian/Turonian boundary. Nature 335:714717.Google Scholar
Arthur, M. A., Schlanger, S. O., and Jenkyns, H. C. 1987. The Cenomanian-Turonian Oceanic Anoxic Event, II. Paleoceanographic controls on organic matter production and preservation. Pp. 401420. In Brooks, J., and Fleet, A. (eds.), Marine Petroleum Source Rocks. Geological Society Special Publication 26.Google Scholar
Batt, R. J. 1987. Pelagic Biofacies of the Greenhorn Sea (Cretaceous)—Evidence from Ammonites and Planktonic Foraminifera. Unpublished Ph.D. dissertation, University of Colorado. Boulder, Colorado.Google Scholar
Bayer, U., and McGhee, G. R. 1984. Evolution in marginal epicontinental basins—the role of phylogenetic and ecological factors. Pp. 164220. In Bayer, U., and Seilacher, A. (eds.), Sedimentary and Evolutionary Cycles. (Friedman, G. [ed.], Lecture Notes in Earth Sciences 1.) Springer-Verlag; Berlin.Google Scholar
Birkelund, T., and Bromley, R. G. (eds.). 1979. The Maastrichtian and Danian of Denmark. Copenhagen University, Symposium Proceedings “Cretaceous–Tertiary Boundary Events” 1:1210.Google Scholar
Bralower, T. J. 1988. Calcareous nannofossil biostratigraphy and assemblages of the Cenomanian–Turonian boundary interval—implications for the origin and timing of oceanic anoxia. Paleoceanography 3:275316.Google Scholar
Brumsack, H. J., and Thurow, J. 1986. The geochemical facies of black shales from the Cenomanian–Turonian boundary event (CTBE). Mitteilungen aus dem Geologisch-Paläontologischen Institut der Universität Hamburg, SCOPE/UNEP Sonderbond Heft 60:247265.Google Scholar
Cappetta, H. 1987. Extinctions et renouvellements fauniques chez les Selaciens post-jurassiques. Mémoires de la Société Géologique de France 150:113131.Google Scholar
Caron, M. 1985. Cretaceous planktic foraminifera. Pp. 1786. In Bolli, H. M., Saunders, J. B., and Perch-Nielson, K. (eds.), Plankton Stratigraphy. Cambridge University Press; Cambridge.Google Scholar
Cobban, W. A. 1984. Mid-Cretaceous ammonite zones, Western Interior, United States. Bulletin of the Geological Society of Denmark 33:7189.Google Scholar
Cobban, W. A. 1985. Ammonite record from the Bridge Creek Member of the Greenhorn Limestone at Pueblo Reservoir State Recreation Area, Colorado. Pp. 135138. In Pratt, L., Kauffman, E. G., and Zelt, F. (eds.), Fine-Grained Deposits and Biofacies of the Cretaceous Western Interior Seaway—Evidence of Cyclic Sedimentary Processes. Society of Economic Paleontologists and Mineralogists Field Trip Guidebook 4; 1985 Midyear Meeting; Golden, Colorado.Google Scholar
Cobban, W. A. 1988. The Upper Cretaceous Ammonite Watinoceras Warren in the Western Interior of the United States. United States Geological Survey Bulletin 1788.CrossRefGoogle Scholar
Cobban, W. A., and Scott, G. R. 1975. Stratigraphy and Ammonite Fauna of the Graneros Shale and Greenhorn Limestone Near Pueblo, Colorado. United States Geological Survey Professional Paper 645.Google Scholar
Diner, R. 1986. Foraminifera/water-mass relationships across the Cenomanian–Turonian Stage boundary. Fourth North American Paleontological Convention, Abstracts with Program:A12.Google Scholar
Eicher, D. L., and Diner, R. 1985. Foraminifera as indications of water mass in the Cretaceous Greenhorn Sea, Western Interior. Pp. 6071. In Pratt, L., Kauffman, E. G., and Zelt, F. (eds.), Fine-Grained Deposits and Biofacies of the Cretaceous Western Interior Seaway—Evidence of Cyclic Sedimentary Processes. Society of Economic Paleontologists and Mineralogists Field Trip Guidebook 4; 1985 Midyear Meeting; Golden, Colorado.Google Scholar
Eicher, D. L., and Worstell, P. 1970. Cenomanian and Turonian foraminifera from the Great Plains, United States. Micropaleontology 16:269324.CrossRefGoogle Scholar
Elder, W. P. 1985. Biotic patterns across the Cenomanian-Turonian extinction boundary near Pueblo, Colorado. Pp. 157169. In Pratt, L., Kauffman, E. G., and Zelt, F. (eds.), Fine-Grained Deposits and Biofacies of the Cretaceous Western Interior Seaway—Evidence of Cyclic Sedimentary Processes. Society of Economic Paleontologists and Mineralogists Field Trip Guidebook 4; 1985 Midyear Meeting; Golden, Colorado.Google Scholar
Elder, W. P. 1987. The paleoecology of the Cenomanian-Turonian (Cretaceous) Stage boundary at Black Mesa, Arizona. Palaios 2:2440.Google Scholar
Elder, W. P. 1988. Geometry of Upper Cretaceous bentonite beds—implications about volcanic source areas and paleowind patterns western interior, United States. Geology 16:835838.2.3.CO;2>CrossRefGoogle Scholar
Elder, W. P., and Kirkland, J. I. 1985. Stratigraphy and depositional environments of the Bridge Creek Limestone Member of the Greenhorn Limestone at Rock Canyon anticline near Pueblo, Colorado. Pp. 122134. In Pratt, L., Kauffman, E. G., and Zelt, F. (eds.), Fine-Grained Deposits and Biofacies of the Cretaceous Western Interior Seaway—Evidence of Cyclic Sedimentary Processes. Society of Economic Paleontologists and Mineralogists Field Trip Guidebook 4; 1985 Midyear Meeting; Golden, Colorado.Google Scholar
Fischer, A. G. 1980. Gilbert-bedding rhythms and geochronology. Geological Society of America Special Paper 183:93104.Google Scholar
Hallam, A. 1968. Morphology, paleoecology and evolution of the genus Gryphaea in the British Lias. Philosophical Transactions of the Royal Society of London, Series B 254:91128.Google Scholar
Haq, B. U., Hardenbol, J., and Vail, P. R. 1987. Chronology of fluctuating sealevels since the Triassic. Science 235:11561167.CrossRefGoogle Scholar
Hattin, D. E. 1971. Widespread, synchronously deposited beds in the Greenhorn Limestone (Upper Cretaceous) of Kansas and southeastern Colorado. American Association of Petroleum Geologists Bulletin 55:110119.Google Scholar
Hattin, D. E. 1975. Stratigraphy and Depositional Environment of the Greenhorn Limestone (Upper Cretaceous) of Kansas. Kansas Geological Survey Bulletin 209.Google Scholar
Hattin, D. E. 1985. Distribution and significance of widespread, time-parallel pelagic limestone beds in the Greenhorn Limestone (Upper Cretaceous) of the central Great Plains and southern Rocky Mountains. Pp. 2837. In Pratt, L., Kauffman, E. G., and Zelt, F. (eds.), Fine-Grained Deposits and Biofacies of the Cretaceous Western Interior Seaway—Evidence of Cyclic Sedimentary Processes. Society of Economic Paleontologists and Mineralogists Field Trip Guidebook 4; 1985 Midyear Meeting; Golden, Colorado.CrossRefGoogle Scholar
Hattin, D. E. 1986a. Interregional model for deposition of Upper Cretaceous pelagic rhythmites, U.S. western interior. Paleoceanography 1:483494.Google Scholar
Hattin, D. E. 1986b. Carbonate substrates of the Late Cretaceous sea, central Great Plains and southern Rocky Mountains. Palaios 1:347367.CrossRefGoogle Scholar
Hilbrecht, H. 1986. On the correlation of the Upper Cenomanian and Lower Turonian of England and Germany (Boreal and Tethys). Newsletters on Stratigraphy 15:115138.CrossRefGoogle Scholar
Hilbrecht, H., and Hoefs, J. 1986. Geochemical and palaeontological studies of the 13C anomaly in boreal and north Tethyan Cenomanian-Turonian sediments in Germany and adjacent areas. Palaeogeography, Palaeoclimatology, Palaeoecology 53:169189.Google Scholar
Hut, P., Alvarez, W., Elder, W. P., Hansen, T., Kauffman, E. G., Keller, G., Shoemaker, E. M., and Weissman, P. R. 1987. Comet showers as a cause of mass extinctions. Nature 329:118126.Google Scholar
Jablonski, D. 1985. Causes and consequences of mass extinctions—a comparative approach. Pp. 183229. In Elliott, D. K. (ed.), Dynamics of Extinctions. John Wiley and Sons; New York.Google Scholar
Jablonski, D. 1986. Background and mass extinctions—the alternation of macroevolutionary regimes. Science 231:129133.CrossRefGoogle ScholarPubMed
Jablonski, D., and Flessa, K. W. 1986. The taxonomic structure of shallow-water marine faunas—implications for Phanerozoic extinctions. Malacologia 27:4366.Google Scholar
Jarvis, I., Carson, G. A., Cooper, M. K. E., Hart, M. B., Leary, P. N., Tocher, B. A., Horne, D., and Rosenfeld, A. 1988. Microfossil assemblages and the Cenomanian-Turonian (late Cretaceous) Oceanic Anoxic Event. Cretaceous Research 9:3103.Google Scholar
Jefferies, R. P. S. 1961. The paleoecology of the Actinocamax plenus subzone (lowest Turonian) in the Anglo-Paris Basin. Palaeontology 4:609647.Google Scholar
Jones, D. S., and Nicol, D. 1986. Origination, survivorship, and extinction of rudist taxa. Journal of Paleontology 60:107115.Google Scholar
Kauffman, E. G 1977a. Cyclothems, biotas and environments, Rock Canyon anticline, Colorado. Pp. 129152. In Kauffman, E. G. (ed.), Cretaceous Facies, Faunas, and Paleoenvironments across the Western Interior Basin. Field Guide, North American Paleontological Convention II. Mountain Geologist 13.Google Scholar
Kauffman, E. G. 1977b. Evolutionary rates and biostratigraphy. Pp. 109141. In Kauffman, E., and Hazel, J. (eds.), Concepts and Methods of Biostratigraphy. Dowden, Hutchinson and Ross, Incorporated; Stroudsburg, Pennsylvania.Google Scholar
Kauffman, E. G. 1978. Evolutionary rates and patterns among Cretaceous Bivalvia. Philosophical Transactions of the Royal Society of London, Series B 284:277304.Google Scholar
Kauffman, E. G. 1981. Ecological reappraisal of the German Posidonien-schiefer. Pp. 311382. In Gray, J., Boucot, A. J., and Berry, W. B. N. (eds.), Communities of the Past. Hutchinson Ross Publishing Company; Stroudsburg, Pennsylvania.Google Scholar
Kauffman, E. G. 1984a. The fabric of Cretaceous marine extinctions. Pp. 151246. In Breggen, W., and Van Couvering, J. (eds.), Catastrophies and Earth History—The New Uniformi-tarianism. Princeton University Press; Princeton, New Jersey.Google Scholar
Kauffman, E. G. 1984b. Dynamic paleobiogeography and evolutionary response in the Cretaceous Western Interior of North America. Pp. 273306. In Westermann, G. E. G. (ed.), Jurassic-Cretaceous Biochronology and Paleogeography of North America. Geological Association of Canada Special Paper 27.Google Scholar
Kirkland, J. I. 1987a. Upper Jurassic and Cretaceous lungfish tooth plates from the western interior, the last dipnoan faunas of North America. Hunteria 2:116.Google Scholar
Kirkland, J. I. 1987b. Integrated high-resolution event and biotic data using graphic correlation—examples from the mid-Cretaceous of Arizona and Colorado. Geological Society of America Abstracts with Program 19:287.Google Scholar
Koch, C. F. 1977. Evolutionary and ecological patterns of Upper Cenomanian (Cretaceous) mollusc distribution in the Western Interior of North America. Unpublished Ph.D. dissertation, George Washington University. Washington, D.C.Google Scholar
Koch, C. F. 1980. Bivalve species duration, aerial extent and population size in a Cretaceous sea. Paleobiology 6:184192.Google Scholar
Kuhnt, W., Thurow, J., Wiedmann, J., and Herbin, J. P. 1986. Oceanic anoxic conditions around the Cenomanian-Turonian boundary and the response of the biota. Mitteilungen aus dem Geologisch-Paläontologischen Institut der Universität Hamburg, SCOPE/UNEP Sonderbond Heft 60:205246.Google Scholar
LaBarbera, M. 1981. The ecology of Mesozoic Gryphaea, Exogyra, and Ilymatogyra (Bivalvia-Mollusca) in a modern ocean. Paleobiology 7:510526.Google Scholar
Leckie, R. M. 1985. Foraminifera of the Cenomanian-Turonian boundary interval, Greenhorn Formation, Rock Canyon Anticline, Pueblo, Colorado. Pp. 139150. In Pratt, L., Kauffman, E. G., and Zelt, F. (eds.), Fine-Grained Deposits and Biofacies of the Cretaceous Western Interior Seaway—Evidence of Cyclic Sedimentary Processes. Society of Economic Paleontologists and Mineralogists Field Trip Guidebook 4; 1985 Midyear Meeting; Golden, Colorado.Google Scholar
Lerbekmo, J. F., Sweet, A. R., and St. Louis, R. M. 1987. The relationship between the iridium anomaly and palynological floral events at three Cretaceous-Tertiary boundary localities in western Canada. Geological Society of America Bulletin 99:325340.2.0.CO;2>CrossRefGoogle Scholar
Levinton, J. 1974. Trophic group and evolution in bivalve molluscs. Paleontology 7:579585.Google Scholar
Morse, D. E. 1985. Neurotransmitter-mimetic inducers of larval settlement and metamorphosis. Bulletin of Marine Science 37:697706.Google Scholar
North American Commission on Stratigraphic Nomenclature. 1983. North American stratigraphic code. American Association of Petroleum Geologists Bulletin 67:841875.Google Scholar
Orth, C. J., Attrep, M. Jr., Mao, X. Y., Kauffman, E. G., Diner, R., and Elder, W. P. 1988a. Iridium abundance maxima in the upper Cenomanian extinction interval. Geophysical Research Letters 15:346349.Google Scholar
Orth, C. J., Attrep, M. Jr., Quintana, L. R., Diner, R., and Elder, W. P. 1988b. Siderophile abundance maxima at upper Cenomanian marine invertebrate extinction horizon, western interior of North America. Nineteenth Lunar and Planetary Science Conference Abstract, Part 2:893894.Google Scholar
Perch-Nielson, K., McKenzie, J., and He, Q. 1982. Biostratigraphy and isotope stratigraphy and the “catastrophic” extinction of calcareous nannoplankton at the Cretaceous/Tertiary boundary. Geological Society of America Special Paper 190:353371.Google Scholar
Pratt, L. M. 1984. Influence of paleoenvironmental factors on preservation of organic matter in middle Cretaceous Greenhorn Formation, Pueblo, Colorado. American Association of Petroleum Geologists Bulletin 68:11461159.Google Scholar
Pratt, L. M. 1985. Isotopic studies of organic matter and carbonate in rocks of the Greenhorn marine cycle. Pp. 3848. In Pratt, L., Kauffman, E. G., and Zelt, F. (eds.), Fine-Grained Deposits and Biofacies of the Cretaceous Western Interior Seaway—Evidence of Cyclic Sedimentary Processes. Society of Economic Paleontologists and Mineralogists Field Trip Guidebook 4; 1985 Midyear Meeting; Golden, Colorado.Google Scholar
Pratt, L. M., and Threlkeld, C. N. 1984. Stratigraphic significance of 13C/12C ratios in mid-Cretaceous rocks of the Western Interior, U.S.A. Pp. 305312. In Scott, D. F., and Glass, D. J. (eds.), Canadian Society of Petroleum Geologists Memoir 9.Google Scholar
Raup, D. M., and Sepkoski, J. J. Jr. 1982. Mass extinctions in the marine fossil record. Science 215:15011503.Google Scholar
Raup, D. M., and Sepkoski, J. J. Jr. 1986. Periodic extinction of families and genera. Science 231:833836.Google Scholar
Scheltema, R. S. 1971. The dispersal of the larvae of shoal-water benthic invertebrate species over long distances by ocean currents. Pp. 728. In Crisp, D. J. (ed.), Fourth European Marine Biology Symposium. Cambridge University Press; Cambridge, England.Google Scholar
Scheltema, R. S. 1977. Dispersal of marine invertebrate organisms—paleobiogeographic and biostratigraphic implications. Pp. 73108. In Kauffman, E. G., and Hazel, J. E. (eds.), Concepts and Methods of Biostratigraphy. Dowden, Hutchinson and Ross, Incorporated; Stroudsburg, Pennsylvania.Google Scholar
Schlanger, S. O., Arthur, M. A., Jenkyns, H. C., and Scholle, P. A. 1987. The Cenomanian-Turonian Oceanic Anoxic Event, I. Stratigraphy and distribution of organic-rich beds and the marine δ13C excursion. Pp. 371399. In Brooks, J., and Fleet, A. (eds.), Marine Petroleum Source Rocks. Geological Society Special Publication 26.Google Scholar
Schwartz, R. D., and James, P. B. 1984. Periodic mass extinctions and the sun's oscillation about the galactic plane. Nature 308:712717.Google Scholar
Sohl, N. F. 1987. Cretaceous gastropods—contrasts between Tethys and Temperate provinces. Journal of Paleontology 61:10851111.Google Scholar
Southam, J. R., Peterson, W. H., and Brass, G. W. 1982. Dynamics of anoxia. Palaeogeography, Palaeoclimatology, Palaeoecology 40:183198.Google Scholar
Stanley, S. M. 1984. Marine mass extinctions—a dominant role for temperatures. Pp. 69118. In Nitecki, M. (ed.), Extinctions. University of Chicago Press; Chicago.Google Scholar
Summerhayes, C. P. 1987. Organic-rich Cretaceous sediments from the North Atlantic. Pp. 301316. In Brooks, J. and Fleet, A. (eds.), Marine Petroleum Source Rocks. Geological Society Special Publication 26.Google Scholar
Thorson, G. 1950. Reproduction and larval ecology of marine bottom invertebrates. Biology Review 25:145.Google Scholar
Valentine, J. W., and Moores, E. M. 1983. Provinciality and diversity across the Permian-Triassic boundary. Pp. 759766. In Logan, A., and Hills, L. V. (eds.), The Permian and Triassic Systems and Their Mutual Boundary. Canadian Society of Petroleum Geologists Memoir 2.Google Scholar
Watkins, D. K. 1986a. Calcareous nannofossil paleoceanography of the Cretaceous Greenhorn Sea. Geological Society of America Bulletin 97:12391249.Google Scholar
Watkins, D. K. 1986b. Nannoplankton productivity fluctuations and rhythmically bedded pelagic carbonates of the Greenhorn Limestone (Upper Cretaceous). North American Paleontological Convention IV Abstracts:A48.Google Scholar
Whitmire, D. P., and Jackson, A. A. IV. 1984. Are periodic mass extinctions driven by a distant solar companion? Nature 308:713715.Google Scholar
Wright, C. W., and Kennedy, W. J. 1981. The Ammonoidea of the Plenus Marls and the Middle Chalk. Monograph of the Palaeontographical Society of London 134.Google Scholar
Zachos, J. C., and Arthur, M. A. 1986. Paleoceanography of the Cretaceous/Tertiary boundary event—inferences from stable isotopic and other data. Paleoceanography 1:526.Google Scholar