Hostname: page-component-848d4c4894-2pzkn Total loading time: 0 Render date: 2024-06-06T11:57:25.772Z Has data issue: false hasContentIssue false

The quality of the fossil record: A sequence stratigraphic perspective

Published online by Cambridge University Press:  26 February 2019

Steven M. Holland*
Affiliation:
Department of Geology, The University of Georgia, Athens, Georgia 30602-2501. E-mail: stratum@gly.uga.edu

Abstract

As paleobiology continues to address an ever broader array of questions, it becomes increasingly important to interpret confidently the meaning of the pattern of fossil occurrences as found in outcrop. To this end, sequence stratigraphy is an important tool for paleobiologists because it predicts the distribution of unconformities, facies changes, and changes in sedimentation rate, all factors known from numerous previous studies to affect the quality of the fossil record. Computer simulations now make it possible not only to model sequence architecture within sedimentary basins, but also to model the occurrence of fossils within those basins. These models generate predictions regarding the stratigraphic distribution of first and last occurrences, changes in species abundance, changes in species morphology, and the distribution of gaps in fossil ranges. Although confirmation of some of these predictions has been found in field studies, the extent to which these predictions describe the fossil record in general is still unknown. If the predicted patterns of fossil occurrences are found to be widespread, it will suggest that a relatively simple model of fossil occurrences in outcrops could become a new tool for solving a wide array of paleobiologic and biostratigraphic problems. With such models, paleobiologists and biostratigraphers will be able to use model data to test the accuracy of newly developed methods of analysis.

Type
Research Article
Copyright
Copyright © 2000 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

Literature Cited

Abbott, S. T., and Carter, R. M. 1997. Macrofossil associations from Mid-Pleistocene cyclothems, Castlecliff Section, New Zealand: implications for sequence stratigraphy. Palaios 12:188210.Google Scholar
Armentrout, J. M. 1991. Paleontologic constraints on depositional modeling: examples of integration of biostratigraphy and seismic stratigraphy, Pliocene-Pleistocene, Gulf of Mexico. Pp. 137170 in Weimer, P. and Link, M. H., eds. Seismic facies and sedimentary processes of submarine fans and turbidite systems. Springer, New York.Google Scholar
Armentrout, J. M. 1996. High resolution sequence biostratigraphy: examples from the Gulf of Mexico Plio-Pleistocene. In Howell, J. A. and Aitken, J. F., eds. High resolution sequence stratigraphy: innovations and applications. Geological Society of London Special Publication 104:6586.Google Scholar
Armentrout, J. M., and Clement, J. F. 1991. Biostratigraphic calibration of depositional cycles: a case study in High Island–Galveston–East Breaks areas, offshore Texas. Pp. 2151 in Armentrout, J. M. and Perkins, B. F., eds. Sequence stratigraphy as an exploration tool: concepts and practices. Gulf Coast Section SEPM Foundation, eleventh annual research conference, Program and abstracts.Google Scholar
Atrops, F., and Ferry, S. 1989. Sequence stratigraphy and changes in the ammonite fauna (Upper Jurassic, S-E France). Pp. 79 in Mesozoic record of Western Tethyan margin. 2ème Congrés Français de Sédimentologie Lyon, Résumé.Google Scholar
Austin, M. P. 1987. Models for the analysis of species’ response to environmental gradients. Vegetatio 69:3545.Google Scholar
Bambach, R. K., and Gilinsky, N. L. 1988. Artifacts in the apparent timing of macroevolutionary “events.” Geological Society of America Abstracts with Programs 20:A104.Google Scholar
Baumiller, T. K. 1996. Exploring the pattern of coordinated stasis: simulations and extinction scenarios. Palaeogeography, Palaeoclimatology, Palaeoecology 127:135146.Google Scholar
Bayer, U., and McGhee, G. R. 1985. Evolution in marginal epicontinental basins: the role of phylogenetic and ecologic factors (Ammonite replacements in the German Lower and Middle Jurassic). Pp. 164220 in Bayer, U. and Seilacher, A., eds. Sedimentary and evolutionary cycles. Springer, New York.Google Scholar
Bosscher, H., and Southam, J. 1992. CARBPLAT—a computer model to simulate the development of carbonate platforms. Geology 20:235238.Google Scholar
Boucot, A. J. 1981. Principles of benthic marine paleoecology. Academic Press, New York.Google Scholar
Bretsky, P. W. 1970. Late Ordovician ecology of the central Appalachians. Peabody Museum of Natural History Bulletin 34:1150.Google Scholar
Brett, C. E. 1995. Sequence stratigraphy, biostratigraphy, and taphonomy in shallow marine environments. Palaios 10:597616.Google Scholar
Brett, C. E. 1998. Sequence stratigraphy, paleoecology, and evolution: biotic clues and responses to sea-level fluctuations. Palaios 13:241262.Google Scholar
Brett, C. E., and Baird, G. C. 1995. Coordinated stasis and evolutionary ecology of Silurian to Middle Devonian faunas in the Appalachian Basin. Pp. 285315 in Erwin, D. H. and Anstey, R. L., eds. New approaches to speciation in the fossil record. Columbia University Press, New York.Google Scholar
Brett, C. E., Miller, K. B., and Baird, G. C. 1990. A temporal hierarchy of paleoecologic processes within a middle Devonian epeiric sea. In Miller, W. III, ed. Paleocommunity temporal dynamics: the long term development of multispecies assemblies. Paleontological Society Special Publication 5:178209.Google Scholar
Brett, C. E., Ivany, L. C., and Schopf, K. M. 1996. Coordinated stasis: an overview. Palaeogeography, Palaeoclimatology, Palaeoecology 127:120.Google Scholar
Brown, J. H. 1984. On the relationship between abundance and distribution of species. American Naturalist 124:255279.Google Scholar
Brown, J. H. 1995. Macroecology. University of Chicago Press, Chicago.Google Scholar
Buhl-Mortensen, L., and H⊘isæter, T. 1993. Mollusc fauna along an offshore-fjord gradient. Marine Ecology Progress Series 97:209224.Google Scholar
Cisne, J. L., Chandlee, G. O., Rabe, B. D., and Cohen, J. A. 1982. Clinal variation, episodic evolution, and possible parapatric speciation: the trilobite Flexicalymene senaria along an Ordovician depth gradient. Lethaia 15:325341.Google Scholar
Connell, S. D., and Lincoln-Smith, M. P. 1999. Depth and the structure of assemblages of demersal fish: experimental trawling along a temperate coast. Estuarine, Coastal and Shelf Science 48:483495.Google Scholar
Culver, S. J. 1988. New foraminiferal depth zonation of the northwestern Gulf of Mexico. Palaios 3:6985.Google Scholar
Dockery, D. T. III. 1986. Punctuated succession of Paleogene mollusks in the northern Gulf Coastal Plain. Palaios 1:582589.Google Scholar
Donovan, S. K., and Paul, C. R. C. 1998. The adequacy of the fossil record. Wiley, Chichester, England.Google Scholar
Eldredge, N., and Gould, S. J. 1972. Punctuated equilibria: an alternative to phyletic gradualism. Pp. 82115 in Schopf, T. J. M., ed. Models in paleobiology. Freeman, Cooper, San Francisco.Google Scholar
Elrick, M., and Read, J. F. 1991. Cyclic ramp-to-basin carbonate deposits, Lower Mississippian, Wyoming and Montana: a combined field and computer modeling study. Journal of Sedimentary Petrology 61:11941224.Google Scholar
Emery, D., and Myers, K. J. 1996. Sequence stratigraphy. Blackwell Scientific, Oxford.Google Scholar
Fernández-López, S., and Meléndez, G. 1996. Phylloceratina ammonoids in the Iberian Basin during the Middle Jurassic: a model of biogeographical and taphonomic dispersal related to relative sea-level changes. Palaeogeography, Palaeoclimatology, Palaeoecology 120:291302.Google Scholar
Finney, S. C., Berry, W. B. N., Cooper, J. D., Ripperdan, R. L., Sweet, W. C., Jacobson, S. R., Soufiane, A., Achab, A., and Noble, P. J. 1999. Late Ordovician mass extinction: a new perspective from stratigraphic sections in central Nevada. Geology 27:215218.Google Scholar
Flemings, P., and Grotzinger, J. P. 1996. STRATA: freeware for analyzing classic stratigraphic problems. GSA Today 6:17.Google Scholar
Foote, M., and Raup, D. M. 1996. Fossil preservation and the stratigraphic ranges of taxa. Paleobiology 22:121140.Google Scholar
Franseen, E. K., Watney, W. L., Kendall, C. G. S. C., and Ross, W., eds. 1991. Sedimentary modeling: computer simulations and methods for improved parameter definition. Kansas Geological Survey Bulletin 233.Google Scholar
Freeman, S. M., Richardson, C. A., and Seed, R. 1998. The distribution and occurrence of Acholoë squamosa (Polychaeta: Polynoidae), a commensal with the burrowing starfich Astropecten irregularis (Echinodermata: Asteroidea). Estuarine, Coastal and Shelf Science 47:107118.Google Scholar
Fürsich, F. T., Oschmann, W., Jaitly, A. K., and Singh, I. B. 1991. Faunal response to transgressive-regressive cycles: example from the Jurassic of western India. Palaeogeography, Palaeoclimatology, Palaeoecology 85:149159.Google Scholar
Gaskell, B. A. 1991. Extinction patterns in Paleogene benthic foraminiferal faunas: relationship to climate and sea level. Palaios 6:216.Google Scholar
Gerrodette, T. 1979. Equatorial submergence in a solitary coral Balanophyllia elegans, and the critical life stage excluding the species from shallow water in the south. Marine Ecology Progress Series 1:227235.Google Scholar
Goldhammer, R. K., Lehmann, P. J., and Dunn, P. A. 1993. The origin of high-frequency platform carbonate cycles and third-order sequences (Lower Ordovician El Paso Group, west Texas): constraints from outcrop data and stratigraphic modeling. Journal of Sedimentary Petrology 63:318359.Google Scholar
Gräfe, K.-U. 1999. Foraminiferal evidence for Cenomanian sequence stratigraphy and palaeoceanography of the Boulonnais (Paris basin, northern France). Palaeogeography, Palaeoclimatology, Palaeoecology 153:4170.Google Scholar
Grassle, J. F., Sanders, H. L., Hessler, R. R., Rowe, G. T., and McLellan, T. 1975. Pattern and zonation: a study of the bathyal megafauna using the research submersible Alvin. Deep-Sea Research 22:457481.Google Scholar
Harris, M. T., and Sheehan, P. M. 1997. Carbonate sequences and fossil communities from the Upper Ordovician-Lower Silurian of the Eastern Great Basin. Brigham Young University Geology Studies 42:105128.Google Scholar
Holland, S. M. 1995a. Sequence stratigraphy, facies control, and their effects on the stratigraphic distribution of fossils. Pp. 123 in Haq, B. U., ed. Sequence stratigraphy and depositional response to eustatic, tectonic and climatic forcing. Kluwer Academic, Dordrecht, Netherlands.Google Scholar
Holland, S. M. 1995b. The stratigraphic distribution of fossils. Paleobiology 21:92109.Google Scholar
Holland, S. M. 1996a. Guidelines for interpreting the stratigraphic record of extinctions: distinguishing pattern from artifact. Pp. 174 in Repetski, J. E., ed. Sixth North American paleontological convention, Abstracts of papers. Paleontological Society Special Publication 8:174.Google Scholar
Holland, S. M. 1996b. Recognizing artifactually generated coordinated stasis: implications of numerical models and strategies for field tests. Palaeogeography, Palaeoclimatology, Palaeoecology 127:147156.Google Scholar
Holland, S. M. 1997. Using time/environment analysis to recognize faunal events in the Upper Ordovician of the Cincinnati Arch. Pp. 309334 in Brett, C. E. and Baird, G. C., eds. Paleontological events: stratigraphic, ecological and evolutionary implications. Columbia University Press, New York.Google Scholar
Holland, S. M., and Patzkowsky, M. E. 1998. Depositional sequences and the stratigraphic distribution of fossils: isolating the effects of condensation. Pp. A307(1–4) in 1998 AAPG Annual Convention Extended Abstracts, Vol. 1. American Association of Petroleum Geologists, Salt Lake City.Google Scholar
Holland, S. M., and Patzkowsky, M. E. 1999. Models for simulating the fossil record. Geology 27:491494.Google Scholar
Horton, B. P., Edwards, R. J., and Lloyd, J. M. 1999. A foraminiferal-based transfer function: implications for sea-level studies. Journal of Foraminiferal Research 29:117129.Google Scholar
Ivany, L. C., Newton, C. R., and Mullins, H. T. 1994. Benthic invertebrates of a modern carbonate ramp: a preliminary survey. Journal of Paleontology 68:417433.Google Scholar
Jablonski, D. 1980. Apparent versus real biotic effects of transgressions and regressions. Paleobiology 6:397407.Google Scholar
Jablonski, D. 1986. Background and mass extinctions: the alternation of macroevolutionary regimes. Science 231:129133.Google Scholar
Jervey, M. T. 1988. Quantitative geological modelling of siliciclastic rock sequences and their seismic expression. Pp. 4769 in Wilgus, C. K., Hastings, B. S., Kendall, C. G. S. C., Posamentier, H. W., Ross, C. A., and Van Wagoner, J. C., eds. Sea-level changes: an integrated approach. Society of Economic Paleontologists and Mineralogists, Tulsa, Okla.Google Scholar
Kendall, C. G. S. C., Strobel, J., Cannon, R., Bezdak, J., and Biswas, G. 1991. The simulation of the sedimentary fill of basins. Journal of Geophysical Research 96:69116929.Google Scholar
Kidwell, S. M. 1991a. Condensed deposits in siliciclastic sequences: expected and observed features. Pp. 682695 in Einsele, G., Ricken, W. and Seilacher, A., eds. Cycles and events in stratigraphy. Springer, Berlin.Google Scholar
Kidwell, S. M. 1991b. The stratigraphy of shell concentrations. Pp. 211290 in Allison, P. A. and Briggs, D. E. G., eds. Taphonomy: releasing the data locked in the fossil record. Plenum, New York.Google Scholar
Kidwell, S. M., and Bosence, D. W. J. 1991. Taphonomy and time-averaging of marine shelly faunas. Pp. 115209 in Allison, P. A. and Briggs, D. E. G., eds. Taphonomy: releasing the data locked in the fossil record. Plenum, New York.Google Scholar
Ludvigsen, R., Westrop, S. R., Pratt, B. R., Tuffnell, P. A., and Young, G. A. 1986. Dual biostratigraphy: zones and biofacies. Geoscience Canada 13:139154.Google Scholar
MacLeod, N. 1991. Punctuated anagenesis and the importance of stratigraphy to paleobiology. Paleobiology 17:167188.Google Scholar
Mancini, E. A., and Tew, B. H. 1995. Geochronology, biostratigraphy and sequence stratigraphy of a marginal marine to marine shelf stratigraphic succession: Upper Paleocene and Lower Eocene, Wilcox Group, eastern Gulf Coastal Plain, U.S.A. In Berggren, W. A., Kent, D. V., Aubry, M.-P. and Hardenbol, J., eds. Geochronology, Time Scales and Global Stratigraphic Correlation. SEPM Special Publication 54:281293. Tulsa, Okla.Google Scholar
Marshall, C. R. 1990. Confidence intervals on stratigraphic ranges. Paleobiology 16:110.Google Scholar
Marshall, C. R. 1994. Confidence intervals on stratigraphic ranges: partial relaxation of the assumption of randomly distributed fossil horizons. Paleobiology 20:459469.Google Scholar
Marshall, C. R. 1997. Confidence intervals on stratigraphic ranges with nonrandom distributions of fossil horizons. Paleobiology 23:165173.Google Scholar
Marshall, C. R., and Ward, P. D. 1996. Sudden and gradual molluscan extinctions in the latest Cretaceous of Western-European Tethys. Science 274:13601363.Google Scholar
McGhee, G. R. Jr., Bayer, U., and Seilacher, A. 1991. Biological and evolutionary responses to transgressive-regressive cycles. Pp. 696708 in Ricken, W. and Seilacher, A., eds. Cycles and events in stratigraphy. Springer, Berlin.Google Scholar
McKinney, M. L. 1986. Biostratigraphic gap analysis. Geology 14:3638.Google Scholar
McKinney, M. L. 1991. Completeness of the fossil record: an overview. Pp. 6683 in Donovan, S. K., ed. The processes of fossilization. Columbia University Press, New York.Google Scholar
Minchin, P. R. 1987. Simulation of multidimensional community patterns: towards a comprehensive model. Vegetatio 71:145156.Google Scholar
Murray, J. W. 1991. Ecology and palaeoecology of benthic foraminifera. Wiley, New York.Google Scholar
Newell, R. C. 1970. Biology of intertidal animals. Elsevier, New York.Google Scholar
Osleger, D., and Read, J. F. 1993. Comparative analysis of methods used to define eustatic variations in outcrop: late Cambrian interbasinal sequence development. American Journal of Science 293:157216.Google Scholar
Pachut, J. F., Cuffey, R. J., and Kobluk, D. R. 1995. Depth-related associations of cryptic-habitat bryozoans from the leeward fringing reef of Bonaire, Netherlands Antilles. Palaios 10:254267.Google Scholar
Palmer, A. R. 1965. Biomere—a new kind of biostratigraphic unit. Journal of Paleontology 39:149153.Google Scholar
Patzkowsky, M. E., and Holland, S. M. 1993. Biotic response to a Middle Ordovician paleoceanographic event in eastern North America. Geology 21:619622.Google Scholar
Patzkowsky, M. E., and Holland, S. M. 1996. Extinction, invasion, and sequence stratigraphy: patterns of faunal change in the Middle and Upper Ordovician of the eastern United States. In Witzke, B. J., Ludvigsen, G. A. and Day, J. E., eds. Paleozoic sequence stratigraphy: views from the North American craton. Geological Society of America Special Paper 306:131142.Google Scholar
Patzkowsky, M. E., and Holland, S. M. 1999. Biofacies replacement in a sequence stratigraphic framework: Middle and Upper Ordovician of the Nashville Dome, Tennessee, USA. Palaios 14:301323.Google Scholar
Paul, C. R. C. 1982. The adequacy of the fossil record. Pp. 75117 in Joysey, K. A. and Friday, A. E., eds. Problems of phylogenetic reconstruction. Academic Press, New York.Google Scholar
Piepenburg, D., and Schmid, M. K. 1996. Distribution, abundance, biomass and mineralization potential of the epibenthic megafauna of the Northeast Greenland Shelf. Marine Biology 125:321332.Google Scholar
Piepenburg, D., Chernova, N. V., von Dorrien, C. F., Gutt, J., Neyelov, A. V., Rachor, E., Saldanha, L., and Schmid, M. K. 1996. Megabenthic communities in the waters around Svalbard. Polar Biology 16:431446.Google Scholar
Piepenburg, D., Voß, J., and Gutt, J. 1997. Assemblages of sea stars (Echinodermata: Asteroidea) and brittle stars (Echinodermata: Ophiuroidea) in the Weddell Sea (Antarctica) and off Northeast Greenland (Arctic): a comparison of diversity and abundance. Polar Biology 17:305322.Google Scholar
Pineda, J. 1993. Boundary effects on the vertical ranges of deep-sea benthic species. Deep-Sea Research I 70:21792192.Google Scholar
Pineda, J., and Caswell, H. 1998. Bathymetric species-diversity patterns and boundary constraints on vertical range distributions. Deep-Sea Research II 45:83101.Google Scholar
Posamentier, H. W., and Allen, G. P. 1993. Variability of the sequence stratigraphic model: effects of local basin factors. Sedimentary Geology 86:91109.Google Scholar
Posamentier, H. W., and James, D. P. 1993. An overview of sequence-stratigraphic concepts: uses and abuses. Pp. 318 in Posamentier, H. W., Summerhayes, C. P., Haq, B. U., and Allen, G. P., eds. Sequence stratigraphy and facies associations. Blackwell Scientific, Oxford.Google Scholar
Posamentier, H. W., Allen, G. P., James, D. P., and Tesson, M. 1992. Forced regressions in a sequence stratigraphic framework: concepts, examples, and exploration significance. American Association of Petroleum Geologists Bulletin 76:16871709.Google Scholar
Raup, D. M. 1985. Mathematical models of cladogenesis. Paleobiology 11:4252.Google Scholar
Read, J. F., Grotzinger, J. P., Bova, J. A., and Koerschner, W. F. 1986. Models for generation of carbonate cycles. Geology 14:107110.Google Scholar
Ross, C. A., and Ross, J. R. P. 1995. Foraminiferal zonation of late Paleozoic depositional sequences. Marine Micropaleontology 26:469478.Google Scholar
Saltzman, M. R. 1999. Upper Cambrian carbonate platform evolution, Elvinia and Taenicephalus Zones (Pterocephaliid-Ptychaspid biomere boundary), northwestern Wyoming. Journal of Sedimentary Research 69:926938.Google Scholar
Siebenaller, J. F., and Somero, G. N. 1989. Biochemical adaptation to the deep-sea. Reviews in Aquatic Sciences 1:125.Google Scholar
Smith, R. I., and Carlton, J. T. 1975. Light's manual: intertidal invertebrates of the central California coast. University of California Press, Berkeley.Google Scholar
Soto, L. A. 1991. Faunal zonation of the deep-water brachyuran crabs in the straits of Florida. Bulletin of Marine Science 49:623637.Google Scholar
Strauss, D., and Sadler, P. M. 1989. Classical confidence intervals and the Bayesian probability estimates for the ends of local taxon ranges. Mathematical Geology 21:411427.Google Scholar
Tait, R. V., and Dipper, F. A. 1998. Elements of marine ecology. Butterworth-Heinemann, Oxford.Google Scholar
Thompson, B. E., Jones, G. F., Laughlin, J. D., and Tsukada, D. T. 1987. Distribution, abundance, and size composition of echinoids from basin slopes off southern California. Bulletin of Southern California Academy of Sciences 86:113125.Google Scholar
Van Wagoner, J. C., Mitchum, R. M., Campion, K. M., and Rahmanian, V. D. 1990. Siliciclastic sequence stratigraphy in well logs, cores, and outcrops. American Association of Petroleum Geologists Methods in Exploration Series, No. 7. Tulsa, Okla.Google Scholar
Vinogradova, N. G. 1962. Vertical zonation in the distribution of deep-sea benthic fauna in the ocean. Deep-Sea Research 8:245250.Google Scholar
Walton, W. R. 1955. Ecology of living benthonic foraminifera, Todos Santos Bay, Baja California. Journal of Paleontology 29:9521018.Google Scholar
Whittaker, R. H. 1970. Communities and ecosystems. Macmillan, New York.Google Scholar
Ziegler, A. M., Cocks, R. M., and Bambach, R. K. 1968. The composition and structure of Lower Silurian marine communities. Lethaia 1:127.Google Scholar
Zong, Y., and Horton, B. P. 1999. Diatom-based tidal-level transfer functions as an aid in reconstructing Quaternary history of sea-level movements in the UK. Journal of Quaternary Science 14:153167.Google Scholar