Hostname: page-component-7c8c6479df-xxrs7 Total loading time: 0 Render date: 2024-03-28T17:08:04.976Z Has data issue: false hasContentIssue false

Taphonomy's contributions to paleobiology

Published online by Cambridge University Press:  08 April 2016

Abstract

Taphonomy established itself in paleontology primarily as a subdiscipline of paleoecology, but it has evolved into a much broader study of the ways in which preservation affects the fossil record. The past decade has seen a change in emphasis from descriptive taphonomic studies of fossil assemblages to more experimental, process-oriented investigations of necrolysis, stratification, and diagenesis of organic remains in modern environments. These actualistic studies are increasing the sophistication of taphonomic analysis in the fossil record by sharpening the diagnosis of bias in paleontological data and by providing a baseline for quantitative modeling of preservational patterns. The analysis of bias is also expanding into the evaluation of temporal resolution in the fossil record (sample acuity, stratigraphic completeness), and taphonomic research is thus contributing to broad-scale problems in evolution, biogeography, and biostratigraphy. In addition, taphonomic studies are providing new insights into paleoenvironmental reconstruction and into the direct paleobiological significance of post mortem processes such as the behavior of scavengers and the role of dead hardparts in structuring benthic communities. One of taphonomy's most promising new frontiers is comparative analysis applied to different taxonomic groups within assemblages and across environments, tectonic settings, and climatic regimes. All of this currently active research is contributing to a better understanding of the fossil record as the result of a dynamic, evolving, integrated system of biological and sedimentological processes that have both limited and enhanced knowledge of Earth history.

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

Literature Cited

Abel, O. 1912. Grundzüge der Palaeobiologie der Wirbeltiere. E. Schweizerbartsche Verlagsbuchhandlung Nagele; Stuttgart.CrossRefGoogle Scholar
Ahrens, T. J. and O'Keefe, J. D. 1983. Impact of an asteroid or comet in the ocean and extinction of terrestrial life. In: Boyton, W. V. and T. H. Abrams, eds. Proc. 13th Lunar and Planet. Sci. Conf., J. Geophys. Res. B88(Suppl.2):A799A806.Google Scholar
Aller, R. C. 1982. Carbonate dissolution in nearshore terrigenous muds: the role of physical and biological reworking. J. Geol. 90:7996.Google Scholar
Alvarez, L. W., Alvarez, W., Asaro, F., and Michel, H. V. 1980. Extraterrestrial cause for the Cretaceous-Tertiary extinction. Science. 208:10951108.Google Scholar
Andrews, P. and Nesbit Evans, E. 1983. Small mammal bone accumulations produced by mammalian carnivores. Paleobiology. 9:289307.Google Scholar
Badgley, C. E. 1982a. How much time is represented in the present? The development of time-averaged modern assemblages as models for the fossil record. Proc. 3d N. Am. Paleontol. Conv. 1:2328.Google Scholar
Badgley, C. E. 1982b. Community reconstruction of a Siwalik mammalian assemblage. Ph.D. Dissertation, Yale Univ. 364 pp.Google Scholar
Bambach, R. K. 1977. Species richness in marine benthic habitats through the Phanerozoic. Paleobiology. 3:152168.Google Scholar
Behrensmeyer, A. K. 1975. The taphonomy and paleoecology of Plio-Pleistocene vertebrate assemblages of Lake Rudolf, Kenya. Bull. Mus. Comp. Zool. 146:473578.Google Scholar
Behrensmeyer, A. K. 1976a. Fossil assemblages in relation to sedimentary environments in the East Rudolf succession. Pp. 383401. In: Coppens, Y., Howell, F. C., Isaac, G. L., and Leakey, R. E., eds. Earliest Man and Environments in the Lake Rudolf Basin. Univ. Chicago Press; Chicago. 615 pp.Google Scholar
Behrensmeyer, A. K. 1976b. Taphonomy and paleoecology in the hominid fossil record. Yb. Phys. Anthropol. 19:3650.Google Scholar
Behrensmeyer, A. K. 1978. Taphonomic and ecologic information from bone weathering. Paleobiology. 4:150162.Google Scholar
Behrensmeyer, A. K. 1982a. Time resolution in fluvial vertebrate assemblages. Paleobiology. 8:211228.Google Scholar
Behrensmeyer, A. K. 1982b. Time sampling intervals in the vertebrate fossil record. Proc. 3d N. Am. Paleontol. Conv. 1:4145.Google Scholar
Behrensmeyer, A. K. and Schindel, D. 1983. Resolving time in paleobiology. Paleobiology. 9:18.Google Scholar
Behrensmeyer, A. K., Western, D., and Dechant Boaz, D. E. 1979. New perspectives in vertebrate paleoecology from a recent bone assemblage. Paleobiology. 5:1221.Google Scholar
Berger, W. H. 1976. Biogenesis in deep sea sediments: production, preservation and interpretation. In; Riley, J. P. and R. Chester, eds. Chem. Oceanogr. 3:265388.Google Scholar
Birkelund, T. and Hankasson, E. 1982. The terminal Cretaceous extinction in Boreal shelf seas—a multicausal event. Geol. Soc. Am. Spec. Pub. 190:373384.Google Scholar
Bonnefille, R. 1979. Méthode palynologique et reconstitutions paléoclimatiques au Cenozoïque dans le Rift Est Africain. Bull. Soc. Géol. France 21:331342.Google Scholar
Brain, C. K. 1958. The Transvaal Ape-Man-Bearing Cave Deposits. Transvaal Mus. Mem. 13:1125.Google Scholar
Brain, C. K. 1967. Hottentot food remains and their meaning in the interpretation of fossil bone assemblages. Sci. Pap. Namib. Desert Res. Sta. 32:111.Google Scholar
Brain, C. K. 1969. The contribution of Namib Desert Hottentots to an understanding of Australopithecine bone accumulations. Sci. Pap. Namib. Desert Res. Sta. 39:1322.Google Scholar
Brett, C. E. and Baird, G. C. 1984. Comparative taphonomy: a key to interpretation of sedimentary sequences using fossil preservation. GSA Abstracts (Southeastern and North-Central Sections) 16:126.Google Scholar
Brett, C. E., Speyer, S. E., and Baird, G. C. In press. Storm-generated sedimentary units: tempestite proximality and event stratification in the Middle Devonian Hamilton Group. In: Brett, C. E., ed. Stratigraphy and Depositional Environments of the Hamilton Group (Middle Devonian) in New York. New YorkState Mus. Sci. Serv. Bull.Google Scholar
Buzas, M. A., Koch, C. F., Culver, Stephen J., and Sohl, N. F. 1982. On the distribution of species occurrence. Paleobiology. 8:143150.Google Scholar
Cadée, G. C. 1968. Molluscan biocoenoses and thanatocoenoses in the Ria de Arosa, Galicia, Spain. Zool. Verh. Rijksmus. Nat. Hist. Leiden. 95:1121.Google Scholar
Cheetham, A. H. and Thomsen, E. 1981. Functional morphology of arborescent animals: strength and design of cheilostome bryozoan skeletons. Paleobiology. 7:355383.Google Scholar
Clark, J., Beerbower, J. R., and Kietzke, K. K. 1967. Oligocene sedimentation, stratigraphy and paleoclimatology in the Big Badlands of South Dakota. Fieldiana Geol. 5:1158.Google Scholar
Damuth, J. 1982. Analysis of the preservation of community structure in assemblages of fossil mammals. Paleobiology. 8:434446.CrossRefGoogle Scholar
Davis, M. B. 1969. Climatic changes in southern Connecticut recorded by pollen deposition at Rogers Lake. Ecology. 50:409422.Google Scholar
Davis, M. B. 1975. Reconstructions of local and regional Holocene environments from the pollen and peat stratigraphies of some Driftless Area peat deposits. Ph.D. thesis, Univ. Wisconsin. 187 pp.Google Scholar
Davis, M. B. and Webb, T. III. 1975. The contemporary distribution of pollen in eastern North America: a comparison with the vegetation. Quat. Res. 5:435465.Google Scholar
Dingus, L. and Sadler, P. M. 1982. The effects of stratigraphic completeness on estimates of evolutionary rates. Syst. Zool. 31:400412.CrossRefGoogle Scholar
Efremov, I. A. 1940. Taphonomy: a new branch of paleontology. Pan-Am. Geol. 74:8193.Google Scholar
Efremov, I. A. 1953. Taphonomé et annates géologiques. Ann. Centre d'Etud. et de Doc. Paléontol. No. 4:1164. (Transl. by Ketchian, S. and Roger, J.)Google Scholar
Flessa, K. W. and Sepkoski, J. J. Jr. 1978. On the relationship between Phanerozoic diversity and changes in habitable area. Paleobiology. 4:359366.Google Scholar
Fursich, F. T. 1982. Rhythmic bedding and shell bed formation in the Upper Jurassic of East Greenland. Pp. 208222. In: Einsele, G. and Seilacher, A., eds. Cyclic and Event Stratification. Springer-Verlag; Berlin.Google Scholar
Gifford, D. P. 1981. Taphonomy and paleoecology: a critical review of archeology's sister discipline. Pp. 365438. In: Schiffer, M. B., ed. Advances in Archaeological Method and Theory, Vol. 4. Academic Press; New York.Google Scholar
Gingerich, P. D. 1982. Time resolution in mammalian evolution: sampling, lineages and faunal turnover. Proc. 3d N. Am. Paleontol. Conv. 1:205210.Google Scholar
Grayson, D. K. 1978. Reconstructing mammalian communities: a discussion of Shotwell's method of paleoecological analysis. Paleobiology. 4:7781.Google Scholar
Havinga, A. J. 1967. Palynology and pollen preservation. Rev. Palaeobot. Palynol. 2:8198.Google Scholar
Haynes, G. 1983. A guide for differentiating mammalian carnivore taxa responsible for gnaw damage to herbivore limb bones. Paleobiology. 9:164172.CrossRefGoogle Scholar
Hill, A. P. 1975. Taphonomy of contemporary and late Cenozoic East African vertebrates. Ph.D. dissertation, Univ. London.Google Scholar
Hill, A. P. 1978. Taphonomical background to fossil man-problems in paleoecology. Pp. 87102. In: Bishop, W. W., ed. Geological Background to Fossil Man. Univ. Toronto Press; Toronto. 585 pp.Google Scholar
Holtzman, R. C. 1979. Maximum likelihood estimation of fossil assemblage composition. Paleobiology. 5:7789.Google Scholar
Jablonski, D. 1980. Apparent versus real biotic effects of transgressions and regressions. Paleobiology. 6:397407.Google Scholar
Jablonski, D. 1983. Apparent versus real extinctions at the end of the Cretaceous Period. Geol. Soc. Am. Abstr. 15:602.Google Scholar
Jablonski, D. 1985. Causes and consequences of mass extinction: a comparative approach. In: Elliott, D. K., ed. Dynamics of Extinction. Wiley; New York.Google Scholar
Johnson, R. G. 1957. Experiments on the burial of shells. J. Geol. 65:527535.Google Scholar
Johnson, R. G. 1960. Models and methods for analysis of the mode of formation of fossil assemblages. Bull. Geol. Soc. Am. 71:10751086.Google Scholar
Johnson, R. G. 1962. Mode of formation of marine fossil assemblages of the Pleistocene Millerton Formation of California. Bull. Geol. Soc. Am. 73:113130.Google Scholar
Johnson, R. G. 1965. Pelecypod death assemblages in Tomales Bay, California. J. Paleontol. 39:8085.Google Scholar
Johnson, R. G. 1972. Conceptual models of benthic marine communities. Pp. 148159. In: Schopf, T. J. M., ed. Models in Paleobiology. W. H. Freeman; San Francisco.Google Scholar
Jones, D. S. 1980. Annual cycle of shell growth increment formation in two continental shelf bivalves and its paleoecological significance. Paleobiology. 6:331340.Google Scholar
Kidwell, S. M. 1981. Stratigraphic patterns of fossil accumulation and their paleobiologic significance. Geol. Soc. Am. Abstr. 13:486.Google Scholar
Kidwell, S. M. 1982. Time scales of fossil accumulation: patterns from Miocene benthic assemblages. Proc. 3d N. Am. Paleontol. Conv. 1:295300.Google Scholar
Kidwell, S. M. 1983. Stratigraphic taphonomy: predicting occurrence, time scales, and modes of formation of fossil deposits. Geol. Soc. Am. Abstr. 15:612.Google Scholar
Kidwell, S. M. and Aigner, T. 1985. Sedimentary dynamics of complex skeletal accumulations: paleoecological and evolutionary implications. In: Bayer, U. and Seilacher, A., eds. Cycles in Sedimentation and Evolution. Springer-Verlag; Berlin.Google Scholar
Kidwell, S. M. and Jablonski, D. 1983. Taphonomic feedback: ecological consequences of shell accumulation. Pp. 195248. In: Tevesz, J. S. and McCall, P. L., eds. Biotic Interactions in Recent and Fossil Benthic Communities. Plenum; New York.CrossRefGoogle Scholar
Koch, C. F. and Sohl, N. F. 1983. Preservational effects in paleoecological studies: Cretaceous mollusc examples. Paleobiology. 9:2634.Google Scholar
Larson, D. W. and Rhoads, D. C. 1983. The evolution of infaunal communities and sedimentary fabrics. Pp. 627646. Biotic Interactions and Fossil Benthic Communities. Plenum; New York.CrossRefGoogle Scholar
Lasker, H. 1976. Effects of differential preservation on the measurement of taxonomic diversity. Paleobiology. 2:8493.CrossRefGoogle Scholar
Lasker, H. 1978. The measurement of taxonomic evolution: preservational consequences. Paleobiology. 4:135149.Google Scholar
Lawrence, D. R. 1968. Taphonomy and information losses in fossil communities. Bull. Geol. Soc. Am. 79:13151330.Google Scholar
Lawrence, D. R. 1971. The nature and structure of paleoecology. J. Paleontol. 45:593607.Google Scholar
Lawrence, D. R. 1979. Taphonomy. Pp. 793799. In: Fairbridge, R. and Jablonski, D., eds. Encyclopedia of Paleontology. Vol. 7. Dowden, Hutchinson & Ross; Stroudsburg, Pa.Google Scholar
Levinton, J. S. and Bambach, R. K. 1975. A comparative study of Silurian and Recent deposit-feeding bivalve communities. Paleobiology. 1:97124.Google Scholar
MacCarthy, B. 1977. Selective preservation of mollusc shells in a Permian beach environment, Sydney Basin, Australia. N. Jahrb. Geol. Paläontol., Mh. 8:466474.Google Scholar
MacDonald, K. B. 1976. Paleocommunities: toward some confidence limits. Pp. 87106. In: Scott, R. W. and West, R. R., eds. Structure and Classification of Paleocommunities. Dowden, Hutchinson, & Ross; Stroudsburg, Pa.Google Scholar
Miller, M. F. 1984. Distribution of biogenic structures in Paleozoic nonmarine and marine-margin sequences: an actualistic model. J. Paleontol. 58:550570.Google Scholar
Olson, E. C. 1952. The evolution of a Permian vertebrate chronofauna. Evol. 6:181196.Google Scholar
Olson, E. C. 1958. Fauna of the Vale and Choza: 14. Summary, review, and integration of the geology and the faunas. Fieldiana Geol. 10:397448.Google Scholar
Olson, E. C. 1980. Taphonomy: its history and role in community evolution. Pp. 519. In: Behrensmeyer, A. K. and Hill, A., eds. Fossils in the Making. Univ. Chicago Press; Chicago.Google Scholar
Peterson, C. H. 1976. Relative abundances of living and dead molluscs in two California lagoons. Lethaia. 9:137148.Google Scholar
Pollack, J. B., Toon, O. B., Ackerman, T. P., McKay, C. T., and Turco, R. P. 1983. Elemental effects of an impact-generated dustcloud: implications for the Cretaceous-Tertiary extinctions. Science. 219:287289.Google Scholar
Raup, D. M. 1976. Species diversity in the Phanerozoic: a tabulation. Paleobiology. 2:279288.Google Scholar
Raup, D. M. 1979. Biases in the fossil record of species and genera. Bull. Carnegie Mus. Nat. Hist. 13:8591.Google Scholar
Raup, D. M. 1982. Large body impacts and terrestrial evolution meeting, October 19–22, 1981. Paleobiology. 8:13.CrossRefGoogle Scholar
Retallack, G. 1984. Completeness of the rock and fossil record: some estimates using fossil soils. Paleobiology 10:5978.Google Scholar
Richter, R. 1928. Aktuopaläontologie und Paläobiologie, eine Abgrenzung. Senckenbergiana. 10:285292.Google Scholar
Sadler, P. M. 1981. Sediment accumulation rates and the completeness of stratigraphic sequences. J. Geol. 89:569584.Google Scholar
Sadler, P. M. and Dingus, L. 1982. Expected completeness of sedimentary sections: estimating a time-scale dependent, limiting factor in the resolution of the fossil record. Proc. 3d N. Am. Paleontol. Conv. 2:461464.Google Scholar
Sanders, H. L. 1968. Marine benthic diversity: a comparative study. Amer. Nat. 102:243282.CrossRefGoogle Scholar
Schäfer, W. 1972. Ecology and Palaeoecology of Marine Environments. Univ. Chicago Press; Chicago. 568 pp.Google Scholar
Schindel, D. E. 1980. Microstratigraphic sampling and the limits of paleontologic resolution. Paleobiology. 6:408426.CrossRefGoogle Scholar
Schindel, D. E. 1982. Resolution analysis: a new approach to the gaps in the fossil record. Paleobiology. 8:340353.CrossRefGoogle Scholar
Schopf, T. J. M. 1978. Fossilization potential of an intertidal fauna: Friday Harbor, Washington. Paleobiology. 4:261270.Google Scholar
Schopf, T. J. M., Collier, K. O., and Bach, B. O. 1980. Relation of the morphology of stick-like bryozoans at Friday Harbor, Washington, to bottom currents, suspended matter and depth. Paleobiology. 6:466476.Google Scholar
Scott, D. B. and Medioli, F. S. 1980. Living vs. total foraminiferal populations: their relative usefulness in paleoecology. J. Paleontol. 54:814831.Google Scholar
Scott, R. W. and West, R. R. 1976. Structure and Classification of Paleocommunities. Dowden, Hutchinson & Ross; Stroudsburg, Pa. 291 pp.Google Scholar
Seilacher, A. 1970. Begriff und Bedeutung der Fossillagerstätten. N. Jahrb. Geol. Paläontol. Abh. 1970:3439.Google Scholar
Seilacher, A. 1976. Allgemeiner Überblick, Bericht 1970–1975 des Sonderforschungsbereichs 53 Palökologie. Zentralblatt für Geologic und Paläontologie, Teil II:206210.Google Scholar
Sepkoski, J. J. Jr. 1975. Stratigraphic biases in the analysis of taxonomic survivorship. Paleobiology. 1:343355.Google Scholar
Sepkoski, J. J. Jr. 1976. Species diversity in the Phanerozoic: species-area effects. Paleobiology. 2:298303.Google Scholar
Sepkoski, J. J. Jr. 1978. Taphonomic factors influencing the lithologic occurrence of fossils in Dresbachian (Upper Cambrian) shaley facies. Geol. Soc. Am. Abstr. 10:490.Google Scholar
Sepkoski, J. J. Jr. 1982. Flat–pebble conglomerates, storm deposits, and the Cambrian bottom fauna. Pp. 371385. In: Einsele, G. and Seilacher, A., eds. Cyclic and Event Stratification. Springer-Verlag; Berlin.Google Scholar
Sepkoski, J. J. Jr., Bambach, R. K., Raup, D. M., and Valentine, J. W. 1981. Phanerozoic marine diversity and the fossil record. Nature. 293:435437.Google Scholar
Sheehan, P.M. and Raup, D. M. 1977. Species diversity in the Phanerozoic: a reflection of labor by systematists? or systematists follow the fossils. Paleobiology. 3:325329.Google Scholar
Shotwell, J. A. 1955. An approach to the paleoecology of mammals. Ecology. 36:327337.Google Scholar
Shotwell, J. A. 1958. Inter-community relationships in Hemphillian (mid-Pliocene) mammals. Ecology. 39:271282.Google Scholar
Signor, P. W. 1978. Species richness in the Phanerozoic: an investigation of sampling effects. Paleobiology. 4:394406.CrossRefGoogle Scholar
Simpson, G. G. 1960. The history of life. Pp. 117180. In: Tax, S., ed. Evolution after Darwin. Vol. 1. Univ. Chicago Press; Chicago.Google Scholar
Sliter, W. V. 1975. Foraminiferal life and residue assemblages from Cretaceous slope deposits. GSA Bull. 86:897906.Google Scholar
Speyer, F. E. and Brett, C. E. 1984. Comparative taphonomy of Middle Devonian trilobite beds. GSA Abstr. (Southeastern and North-Central Sections). 16:198.Google Scholar
Thayer, C. W. 1983. Sediment-mediated biological disturbance and the evolution of the marine benthos. Pp. 480625. In: Tevesz, M. J. S. and McCall, P. L., eds. Biotic Interactions in Recent and Fossil Benthic Communities. Plenum; New York.Google Scholar
Van Valen, L. M. 1984. Review of Silver, L. T. and P. H. Schultz, eds. Geological Implications of Impacts of Large Asteroids and Comets on the Earth. Paleobiology. 10:121137.Google Scholar
Velbel, D. B. 1984. Sedimentology and taphonomy in a clastic Ordovician sea. Geol. Soc. Am. Abstr. 16:204.Google Scholar
Vermeij, G. J. 1977. The Mesozoic marine revolution: evidence from snails, predators and grazers. Paleobiology. 3:245258.Google Scholar
Vermeij, G. J., Zipser, E., and Dudley, E. C. 1980. Predation in time and space: peeling and drilling in terebrid gastropods. Paleobiology. 6:352364.Google Scholar
Voorhies, M. 1969. Taphonomy and population dynamics of an early Pliocene verrebrate fauna, Knox County, Nebraska. Contrib. Geol., Spec. Paper No. 1. Univ. Wyoming; Laramie.Google Scholar
Walker, K. R. and Parker, W. C. 1976. Population structure of a pioneer and a later stage species in an Ordovician ecological succession. Paleobiology. 2:191201.CrossRefGoogle Scholar
Warme, J. E. 1969. Live and dead molluscs in a coastal lagoon. J. Paleontol. 43:141150.Google Scholar
Warme, J. E. 1971. Paleoecological aspects of a modern coastal lagoon. Univ. Cal. Publ. Geol. 87:1131.Google Scholar
Warme, J. E., Ekdale, A. A., Ekdale, S. F., and Peterson, C. H. 1976. Raw material of the fossil record. Pp. 169243. In: Scott, R. W. and West, R. R., eds. Structure and Classification of Paleocommunities. Dowden, Hutchinson, & Ross; Stroudsburg, Pa.Google Scholar
Wasmund, E. 1926. Biocoenose und thanatocoenose. Arch. Hydrobiol. 17:1116.Google Scholar
Watkins, R. and Hurst, J. M. 1977. Community relations of Silurian crinoids at Dudley, England. Paleobiology. 3:207217.Google Scholar
Weigelt, J. 1927. Resente Wirbeltierleichen und ihre Paläobiologische Bedeutung. Verlag von Max Weg, Leipzig. 227 pp.Google Scholar
Westermann, G. E. G. and Ward, P. 1980. Septum morphology and bathymetry in cephalopods. Paleobiology. 6:4850.CrossRefGoogle Scholar
Wing, S. L. 1984. Relation of paleovegetation to geometry and cyclicity of some fluvial carbonaceous deposits. J. Sedimentol. Petrol. 54:5266.Google Scholar
Zangerl, R. and Richardson, E. S. Jr. 1963. The paleoecological history of two Pennsylvanian black shales. Fieldiana—Geol. Mem. 4:1352.Google Scholar