Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-26T13:21:50.646Z Has data issue: false hasContentIssue false

Taphonomy and paleobiology

Published online by Cambridge University Press:  26 February 2019

Anna K. Behrensmeyer
Affiliation:
Department of Paleobiology, National Museum of Natural History, Smithsonian Institution, MRC 121, Washington, D.C. 20560. E-mail: Behrensmeyer.Kay@NMNH.SI.edu
Susan M. Kidwell
Affiliation:
Department of Geophysical Sciences, University of Chicago, 5734 South Ellis Avenue, Chicago, Illinois 60637. E-mail: skidwell@midway.uchicago.edu
Robert A. Gastaldo
Affiliation:
Department of Geology, Colby College, Waterville, Maine 04901-4799. E-mail: ragastal@colby.edu

Abstract

Taphonomy plays diverse roles in paleobiology. These include assessing sample quality relevant to ecologic, biogeographic, and evolutionary questions, diagnosing the roles of various taphonomic agents, processes and circumstances in generating the sedimentary and fossil records, and reconstructing the dynamics of organic recycling over time as a part of Earth history. Major advances over the past 15 years have occurred in understanding (1) the controls on preservation, especially the ecology and biogeochemistry of soft-tissue preservation, and the dominance of biological versus physical agents in the destruction of remains from all major taxonomic groups (plants, invertebrates, vertebrates); (2) scales of spatial and temporal resolution, particularly the relatively minor role of out-of-habitat transport contrasted with the major effects of time-averaging; (3) quantitative compositional fidelity; that is, the degree to which different types of assemblages reflect the species composition and abundance of source faunas and floras; and (4) large-scale variations through time in preservational regimes (megabiases), caused by the evolution of new bodyplans and behavioral capabilities, and by broad-scale changes in climate, tectonics, and geochemistry of Earth surface systems. Paleobiological questions regarding major trends in biodiversity, major extinctions and recoveries, timing of cladogenesis and rates of evolution, and the role of environmental forcing in evolution all entail issues appropriate for taphonomic analysis, and a wide range of strategies are being developed to minimize the impact of sample incompleteness and bias. These include taphonomically robust metrics of paleontologic patterns, gap analysis, equalizing samples via rarefaction, inferences about preservation probability, isotaphonomic comparisons, taphonomic control taxa, and modeling of artificial fossil assemblages based on modern analogues. All of this work is yielding a more quantitative assessment of both the positive and negative aspects of paleobiological samples. Comparisons and syntheses of patterns across major groups and over a wider range of temporal and spatial scales present a challenging and exciting agenda for taphonomy in the coming decades.

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

Allen, J. R. L., Spicer, R. A., and Behrensmeyer, A. K. 1990. Transport—hydrodynamics. Pp. 227235 in Briggs and Crowther 1990.Google Scholar
Allison, P. A. 1986. Soft-bodied animals in the fossil record: the role of decay in fragmentation during transport. Geology 14:979981.Google Scholar
Allison, P. A. 1988. Konservat-Lagerstätten: cause and classification. Paleobiology 14:331344.Google Scholar
Allison, P. A., and Briggs, D. E. G., eds. 1991a. Taphonomy, releasing the data locked in the fossil record. Plenum, New York.Google Scholar
Allison, P. A., and Briggs, D. E. G., eds. 1991b. The taphonomy of soft-bodied animals. Pp. 120140 in Donovan 1991.Google Scholar
Allison, P. A., and Briggs, D. E. G., eds. 1993a. Paleolatitudinal sampling bias, Phanerozoic species diversity, and the end-Permian extinction. Geology 21:6568.Google Scholar
Allison, P. A., and Briggs, D. E. G., eds. 1993b. Exceptional fossil record: distribution of soft-tissue preservation through the Phanerozoic. Geology 21:527530.Google Scholar
Allison, P. A., and Pye, K. 1994. Early diagenetic mineralization and fossil preservation in modern carbonate concretions. Palaios 9:561575.Google Scholar
Allison, P. A., Smith, C. R., Kukert, H., Deming, J. W., and Bennett, B. 1991. Deep-water taphonomy of vertebrate carcasses: a whale skeleton in the bathyal Santa Catalina Basin. Paleobiology 17:7889.Google Scholar
Alroy, J. 1996. Constant extinction, constrained diversification, and uncoordinated stasis in North American mammals. Palaeogeography, Palaeoclimatology, Palaeoecology 127:285311.Google Scholar
Alroy, J., Koch, P. L., and Zachos, J. C. 2000. Global climate change and North American mammalian evolution. In Erwin, D. H. and Wing, S. L., eds. Deep time: Paleobiology's perspective. Paleobiology 26(Suppl. to No. 4):259288.Google Scholar
Anderson, L. C., Sen Gupta, B. K., McBride, R. A., and Byrnes, M. R. 1997. Reduced seasonality of Holocene climate and pervasive mixing of Holocene marine section: northeastern Gulf of Mexico shelf. Geology 25:127130.Google Scholar
Anderson, L. C., McBride, R. A., Taylor, M. J., and Byrnes, M. R. 1998. Late Holocene record of community replacement preserved in time-averaged molluscan assemblages, Louisiana chenier plain. Palaios 13:488499.Google Scholar
Andrews, P. 1990. Owls, caves, and fossils. University of Chicago Press, Chicago.Google Scholar
Andrews, P., and Cook, J. 1989. Natural modifications to bones in a temperate setting. Man 20:675691.Google Scholar
Archer, M., Hand, S. J., and Godthelp, H. 1991. Riversleigh: the story of animals in ancient rainforests of inland Australia. Reed Books, Balgowlah, New South Wales.Google Scholar
Argast, S., Farlow, J. O., Gabet, R. M., and Brinkman, D. L. 1987. Transport-induced abrasion of fossil reptilian teeth: implications for the existence of Tertiary dinosaurs in the Hell Creek Formation, Montana. Geology 15:927930.Google Scholar
Aronson, R. B. 1992. Decline of the Burgess Shale fauna: ecologic or taphonomic restriction? Lethaia 25:225229.Google Scholar
Aslan, A., and Behrensmeyer, A. K. 1996. Taphonomy and time resolution of bone assemblages in a contemporary fluvial system: the East Fork River, Wyoming. Palaios 11:411421.Google Scholar
Ausich, W. I. 1997. Regional encrinites: a vanished lithofacies. Pp. 509519 in Brett, C. E. and Baird, G. C., eds. Paleontological events: stratigraphic, ecological and evolutionary implications. Columbia University Press, New York.Google Scholar
Ausich, W. I., and Sevastopulo, G. D. 1994. Taphonomy of Lower Carboniferous crinoids from the Hook Head Formation, Ireland. Lethaia 27:245256.Google Scholar
Bada, J. L., Wang, X. S., and Hamilton, H. 1999. Preservation of key biomolecules in the fossil record: current knowledge and future challenges. Philosophical Transactions of the Royal Society of London B 354:7787.Google Scholar
Badgley, C. E. 1982. How much time is represented in the present? The development of time-averaged modern assemblages as models for the fossil record. In Mamet, B. and Copeland, M. J., eds. Proceedings of the Third North American Paleontological Convention 1:2328.Google Scholar
Badgley, C. E. 1986. Counting individuals in mammalian fossil assemblages from fluvial environments. Palaios 1:328355.Google Scholar
Badgley, C. E., and Behrensmeyer, A. K. 1995. Preservational, paleoecological and evolutionary patterns in the Paleogene of Wyoming-Montana and the Neogene of Pakistan. Palaeogeography, Palaeoclimatology, Palaeoecology 115:319340.Google Scholar
Baird, G. C., Sroka, S. D., Shabica, C. W., and Kuecher, G. J. 1986. Taphonomy of Middle Pennsylvanian Mazon Creek area fossil localities, northeast Illinois: significance of exceptional fossil preservation in syngenetic concretions. Palaios 1:271285.Google Scholar
Bambach, R. K. 1977. Species richness in marine benthic habitats through the Phanerozoic. Paleobiology 3:152167.Google Scholar
Barry, J. C., Morgan, M. E., Winkler, A. J., Flynn, L. J., Lindsay, E. H., Jacobs, L. L., and Pilbeam, D. 1991. Faunal interchange and Miocene terrestrial vertebrates of southern Asia. Paleobiology 17:231245.Google Scholar
Bartels, C., Briggs, D. E. G., and Brassel, G. 1998. The fossils of the Hunsrück Slate: marine life in the Devonian. Cambridge University Press, Cambridge.Google Scholar
Barthel, K. W., Winburne, N. H. M., and Morris, S. Conway. 1990. Solnhofen: a study in Mesozoic paleontology. Cambridge University Press, Cambridge.Google Scholar
Bartley, J. K. 1996. Actualistic taphonomy of cyanobacteria: implications for the Precambrian fossil record. Palaios 11:571586.Google Scholar
Bartram, L. E. Jr., and Marean, C. W. 1999. Explaining the “Klasies pattern”: Kua ethnoarchaeology, the Die Kelders Middle Stone Age archaeofauna, long bone fragmentation and carnivore ravaging. Journal of Archaeological Science 26:929.Google Scholar
Baumiller, T. K., Llewellyn, G., Messing, C. G., and Ausich, W. E. 1995. Taphonomy of isocrinoid stalks: influence of decay and autonomy. Palaios 10:8795.Google Scholar
Behrensmeyer, A. K. 1978a. Taphonomic and ecological information from bone weathering. Paleobiology 4:150162.Google Scholar
Behrensmeyer, A. K. 1978b. Correlation in Plio-Pleistocene sequences of the northern Lake Turkana Basin: a summary of evidence and issues. Pp. 421440 in Bishop, W. W., ed. Geological background to fossil man. Scottish Academic Press, Edinburgh.Google Scholar
Behrensmeyer, A. K. 1982. Time resolution in fluvial vertebrate assemblages. Paleobiology 8:211227.Google Scholar
Behrensmeyer, A. K. 1987. Miocene fluvial facies and vertebrate taphonomy in northern Pakistan. In Ethridge, F. G., Flores, R. M., and Harvey, M. D., eds. Recent developments in fluvial sedimentology. SEPM Special Publication 39:169176.Google Scholar
Behrensmeyer, A. K. 1988. Vertebrate preservation in fluvial channels. Palaeogeography, Palaeoclimatology, Palaeoecology 63:183199.Google Scholar
Behrensmeyer, A. K. 1990. Transport/hydrodynamics of bones. Pp. 232235 in Briggs and Crowther 1990.Google Scholar
Behrensmeyer, A. K. 1991. Terrestrial vertebrate accumulation. Pp. 291335 in Allison and Briggs 1991.Google Scholar
Behrensmeyer, A. K. 1993. The bones of Amboseli: bone assemblages and ecological change in a modern African ecosystem. National Geographic Research 9:402421.Google Scholar
Behrensmeyer, A. K., and Chapman, R. E. 1993. Models and simulations of time-averaging in terrestrial vertebrate accumulations. Pp. 125149 in Kidwell and Behrensmeyer 1993.Google Scholar
Behrensmeyer, A. K., and Dechant Boaz, D. E. 1980. The Recent bones of Amboseli Park, Kenya, in relation to East African paleoecology. Pp. 7292 in Behrensmeyer, A. K. and Hill, A. P., eds. Fossils in the making. University of Chicago Press, Chicago.Google Scholar
Behrensmeyer, A. K., and Hook, R. W. 1992. Paleoenvironmental contexts and taphonomic modes. Pp. 15136 in Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A., Potts, R., Sues, H.-D., and Wing, S. L., eds. Terrestrial ecosystems through time. University of Chicago Press, Chicago.Google Scholar
Behrensmeyer, A. K., and Kidwell, S. M. 1985. Taphonomy's contributions to paleobiology. Paleobiology 11:105119.Google Scholar
Behrensmeyer, A. K., Todd, N. E., Potts, R., and McBrinn, G. E. 1997. Late Pliocene faunal turnover in the Turkana Basin, Kenya and Ethiopia. Science 278:15891594.Google Scholar
Bell, M. A., Sadagursky, M. S., and Baumgartner, J. V. 1987. Utility of lacustrine deposits for study of variation within fossil samples. Palaios 2:455466.Google Scholar
Bell, M. A., Wells, C. E., and Marshall, J. A. 1989. Mass-mortality layers of fossil stickleback fish: catastrophic kills of polymorphic schools. Evolution 43:607619.Google Scholar
Bengtson, S. 1994. Early life on Earth (Nobel Symposium No. 84). Columbia University Press, New York.Google Scholar
Bennington, J. B., and Rutherford, S. D. 1999. Precision and reliability in paleocommunity comparisons based on cluster-confidence intervals: how to get more statistical bang for your sampling buck. Palaios 14:506515.Google Scholar
Benton, M. J. 1998. The quality of the fossil record of the vertebrates. Pp. 269300 in Donovan, S. K. and Paul, C. R. C., eds. The adequacy of the fossil record. Wiley, New York.Google Scholar
Berner, R. A. 1991. A model for atmospheric CO2 over Phanerozoic time. American Journal of Science 291:339376.Google Scholar
Berner, R. A., and Canfield, D. E. 1989. A new model for atmospheric oxygen over Phanerozoic time. American Journal of Science 289:333361.Google Scholar
Best, M. M. R., and Kidwell, S. M. 1996. Bivalve shell taphonomy in tropical siliciclastic environments: preliminary experimental results. In Repetski, J. E., ed. Sixth North American paleontological convention, Abstracts of papers. Paleontological Society Special Publication 8:34.Google Scholar
Best, M. M. R., and Kidwell, S. M. 2000a. Bivalve taphonomy in tropical mixed siliciclastic-carbonate settings. I. Environmental variation in shell condition. Paleobiology 26:80102.Google Scholar
Best, M. M. R., and Kidwell, S. M. 2000b. Bivalve taphonomy in tropical mixed siliciclastic-carbonate settings. II. Effect of bivalve life habits and shell types. Paleobiology 26:103115.Google Scholar
Best, M. M. R., Kidwell, S. M., Ku, T. C. W., and Walter, L. M. 1999. The role of microbial iron reduction in the preservation of skeletal carbonate: bivalve taphonomy and porewater geochemistry in tropic siliciclastics vs. carbonates. Geological Society of America Abstracts with Programs 31:419420.Google Scholar
Bishop, J. D. D. 1989. Colony form and the exploitation of spatial refuges by encrusting bryozoa. Biological Reviews 64:197218.Google Scholar
Blob, R. W. 1997. Relative hydrodynamic dispersal potentials of soft-shelled turtle elements: implications for interpreting skeletal sorting in assemblages of non-mammalian terrestrial vertebrates. Palaios 12:151164.Google Scholar
Blumenschine, R. 1986. Carcass consumption sequences and the archaeological distinction of scavenging and hunting. Journal of Human Evolution 15:639659.Google Scholar
Blumenschine, R. 1988. An experimental model of the timing of hominid and carnivore influence on archaeological bone assemblages. Journal of Archaeological Science 15:483502.Google Scholar
Blumenschine, R. 1991. Hominid carnivory and foraging strategies and the socio-economic function of early archaeological sites. Philosophical Transactions of the Royal Society of London B 334:211221.Google Scholar
Bonnichsen, R., and Sorg, M. H. 1989. Bone modification. Institute for Quaternary Studies, University of Maine, Orono.Google Scholar
Bosence, D. W. J. 1979. Live and dead faunas from coralline algal gravels, Co. Galway. Palaeontology 22:449478.Google Scholar
Bottjer, D. J., and Jablonski, D. 1988. Paleoenvironmental patterns in the evolution of post-Paleozoic benthic marine invertebrates. Palaios 3:540560.Google Scholar
Boulton, A. J., and Boon, P. I. 1991. A review of methodology used to measure leaf litter decomposition in lotic environments: time to turn over an old leaf? Australian Journal of Marine and Freshwater Research 42:143.Google Scholar
Bown, T. M., and Beard, K. C. 1990. Systematic lateral variation in the distribution of fossil mammals in alluvial paleosols, lower Eocene Willwood Formation, Wyoming. Geological Society of America Special Paper 243:135151.Google Scholar
Brandt, D. S. 1986. Preservation of event beds through time. Palaios 1:9296.Google Scholar
Brandt, D. S., and Elias, R. J. 1989. Temporal variation in tempestite thickness may be a geologic record of atmospheric CO2 . Geology 17:951952.Google Scholar
Brett, C. E. 1995. Sequence stratigraphy, biostratigraphy, and taphonomy in shallow marine environments. Palaios 10:597616.Google Scholar
Brett, C. E., and Baird, G. C. 1986. Comparative taphonomy: a key to paleoenvironmental interpretation based on fossil preservation. Palaios 1:207227.Google Scholar
Brett, C. E., and Seilacher, A. 1991. Fossil Lagerstätten: a taphonomic consequence of event sedimentation. Pp. 283297 in Einsele, G., Ricken, W., and Seilacher, A., eds. Cycles and events in stratigraphy. Springer, Berlin.Google Scholar
Brett, C. E., Boucot, A. J., and Jones, B. 1993. Absolute depths of Silurian benthic assemblages. Lethaia 26:2540.Google Scholar
Brett, C. E., Whiteley, T. E., Allison, P. A., and Yochelson, E. L. 1999. The Walcott-Rust quarry: Middle Ordovician trilobite Konservat-Lagerstätten. Journal of Paleontology 73:288305.Google Scholar
Briggs, D. E. G. 1993. Fossil biomolecules. Bulletin of the Biochemical Society 15:812.Google Scholar
Briggs, D. E. G. 1995. Experimental taphonomy. Palaios 10:539550.Google Scholar
Briggs, D. E. G., and Clarkson, E. N. K. 1990. The late Palaeozoic radiation of malacostracan crustaceans. In Taylor, P. D. and Larwood, G. P., eds. Major evolutionary radiations. Systematics Association Special Volume 42:165186. Clarendon, Oxford.Google Scholar
Briggs, D. E. G., and Crowther, P. R., eds. 1990. Paleobiology, a synthesis. Blackwell Science, Oxford.Google Scholar
Briggs, D. E. G., and Kear, A. J. 1994a. Decay of Branchiostoma: Implications for soft-tissue preservation in conodonts and other primitive chordates. Lethaia 26:275287.Google Scholar
Briggs, D. E. G., and Kear, A. J. 1994b. Decay and mineralization of shrimps. Palaios 9:431456.Google Scholar
Briggs, D. J., Gilbertson, D. D., and Harris, A. L. 1990. Molluscan taphonomy in a braided river environment and its implications for studies of Quaternary cold-stage river deposits. Journal of Biogeography 17:623637.Google Scholar
Briggs, D. E. G., Wilby, P. R., Pérez-Moreno, B. P., Sanz, J. L., and Fregenal-Martínez, M. 1997. The mineralization of dinosaur soft tissue in the Lower Cretaceous of Las Hoyas, Spain. Journal of the Geological Society, London 154:587588.Google Scholar
Briggs, D. E. G., Stankiewicz, B. A., Meischner, D., Bierstedt, A., and Evershed, R. P. 1998. Taphonomy of arthropod cuticles from Pliocene late sediments, Willershausen, Germany. Palaios 13:386394.Google Scholar
Briggs, D. E. G., Evershed, R. P., and Lockheart, M. J. 2000. The molecular paleontology of continental fossils. In Erwin, D. H. and Wing, S. L., eds. Deep time: Paleobiology's perspective. Paleobiology 26(Suppl. to No. 4):169193.Google Scholar
Buatois, L. A., Mángano, A. G., Genise, J. F., and Taylor, T. N. 1998. The ichnologic record of the continental invertebrate invasion: evolutionary trends in environmental expansion, ecospace utilization, and behavioral complexity. Palaios 13:217240.Google Scholar
Budd, D. A., and Hiatt, E. E. 1993. Mineralogical stabilization of high-magnesium calcite: geochemical evidence for intracrystal recrystallization within Holocene porcellaneous foraminifera. Journal of Sedimentary Petrology 63:261274.Google Scholar
Burnham, R. J. 1989. Relationships between standing vegetation and leaf litter in a paratropical forest: implications for paleobotany. Review of Palaeobotany and Palynology 58:532.Google Scholar
Burnham, R. J. 1990. Paleobotanical implications of drifted seeds and fruits from modern mangrove litter, Twin Cays, Belize. Palaios 5:364370.Google Scholar
Burnham, R. J. 1993. Reconstructing richness in the plant fossil record. Palaios 8:376384.Google Scholar
Burnham, R. J. 1994. Patterns in tropical leaf litter and implications for angiosperm paleobotany. Review of Palaeobotany and Palynology 81:99113.Google Scholar
Burnham, R. J., and Spicer, R. A. 1986. Fossil litter preserved by volcanic activity at El Chicón, Mexico: a potentially accurate record of the pre-eruption vegetation. Palaios 1:158161.Google Scholar
Burnham, R. J., Wing, S. L., and Parker, G. G. 1992. The reflection of deciduous forest communities in leaf litter: implications for autochthonous litter assemblages from the fossil record: Paleobiology 18:3049.Google Scholar
Butterfield, N. J. 1990. Organic preservation of non-mineralizing organisms and the taphonomy of the Burgess Shale. Paleobiology 16:272286.Google Scholar
Butterfield, N. J. 1995. Secular distribution of Burgess Shale-type preservation. Lethaia 28:113.Google Scholar
Cadée, G. C. 1984. Macrobenthos and macrobenthic remains on the Oyster Ground, North Sea. Netherlands Journal of Sea Research 18:160178.Google Scholar
Cadée, G. C. 1991. The history of taphonomy. Pp. 321 in Donovan 1991.Google Scholar
Cadée, G. C. 1994. Eider, shelduck, and other predators, the main producers of shell fragments in the Wadden Sea—paleoecological implications. Palaeontology 37:181202.Google Scholar
Cadée, G. C. 1999. Bioerosion of shells by terrestrial gastropods. Lethaia 32:253260.Google Scholar
Calleja, M., Rossignol-Strick, M., and Duzer, D. 1993. Atmospheric pollen content off West Africa. Review of Palaeobotany and Palynology 79:335368.Google Scholar
Callender, W. R., Powell, E. N., and Staff, G. M. 1994. Taphonomic rates of molluscan shells placed in authochthonous assemblages on the Louisiana contental slope. Palaios 9:6073.Google Scholar
Carroll, R. L. 1997. Patterns and processes of vertebrate evolution. Cambridge University Press, Cambridge.Google Scholar
Cate, A. S., and Evans, I. 1994. Taphonomic significance of the biomechanical fragmentation of live molluscan shell material by a bottom-feeding fish (Pogonias cromis) in Texas coastal bays. Palaios 9:254274.Google Scholar
Chafetz, H., and Buczynski, C. 1992. Bacterially induced lithification of microbial mats. Palaios 7:277293.Google Scholar
Chatterton, B. D. E., and Speyer, S. E. 1997. Ontogeny. Pp. 173247 in Whittington, H. B. et al. Arthropoda 1, Trilobita, revised. Part O of R. C. Moore and C. Teichert, eds. Treatise on invertebrate paleontology. Geological Society of America and University of Kansas, Boulder, Colo.Google Scholar
Claassen, C. 1998. Shells. Cambridge Manuals in Archaeology. Cambridge University Press, Cambridge.Google Scholar
Clark, G. R. II. 1999. Organic matrix taphonomy in some molluscan shell microstructures. Palaeogeography, Palaeoclimatology, Palaeoecology 149:305312.Google Scholar
Clyde, W. C., and Gingerich, P. D. 1998. Mammalian community response to the latest Paleocene thermal maximum: an isotaphonomic study in the northern Bighorn Basin, Wyoming. Geology 26:10111014.Google Scholar
Collinson, M. E. 1983. Accumulations of fruits and seeds in three small sedimentary environments in southern England and their palaeoecological implications. Annals of Botany 52:583592.Google Scholar
Cook, E. 1995. Taphonomy of two non-marine Lower Cretaceous bone accumulations from southeastern England. Palaeogeography, Palaeoclimatology, Palaeoecology 116:263270.Google Scholar
Crowley, S. S., Dufek, D. A., Stanton, R. W., and Ryer, T. A. 1994. The effects of volcanic ash disturbances on a peat-forming environment: environmental disruption and taphonomic consequences. Palaios 9:158174.Google Scholar
Cruz-Uribe, K., and Klein, R. G. 1998. Hyrax and hare bones from modern South African eagle roosts and the detection of eagle involvement in fossil bone assemblages. Journal of Archaeological Science 25:135147.Google Scholar
Cummins, H. 1994. Taphonomic processes in modern fresh-water molluscan death assemblages: implications of the fresh-water fossil record. Palaeogeography, Palaeoclimatology, Palaeoecology 108:5573.Google Scholar
Cummins, H., Powell, E. N., Stanton, R. J., and Staff, G. M. 1986a. The size frequency distribution in palaeoecology: effects of taphonomic processes during formation of molluscan death assemblages in Texas bays. Palaeontology 29:495518.Google Scholar
Cummins, H., Powell, E. N., Stanton, R. J., and Staff, G. M. 1986b. The rate of taphonomic loss in modern benthic habitats: how much of the potentially preservable community is preserved? Palaeogeography, Palaeoclimatology, Palaeoecology 52:291320.Google Scholar
Cunningham, C. R., Feldman, H. R., Franseen, E. K., Gastaldo, R. A., Mapes, G., Maples, C. G., and Schultze, H.-P. 1993. The Upper Carboniferous Hamilton fossil Lagerstätte in Kansas: a valley fill, tidally influenced deposit. Lethaia 26:225236.Google Scholar
Cutler, A. H. 1989. Shells survive—loss, persistence and accumulation of hardparts in shallow marine sediments. Geological Society of America Abstracts with Programs 21:A71.Google Scholar
Cutler, A. H. 1991. Nested faunas and extinction in fragmented habitats. Conservation Biology 5:496505.Google Scholar
Cutler, A. H. 1993. Mathematical models of temporal mixing in the fossil record. Pp. 169187 in Kidwell and Behrensmeyer 1993.Google Scholar
Cutler, A. H., and Behrensmeyer, A. K. 1996. Models of vertebrate mass mortality events at the K/T Boundary. In Ryder, G., Fastovsky, D., and Gartner, S., eds. The Cretaceous-Tertiary Event and other catastrophes in earth history. Geological Society of America Special Paper 307:375380.Google Scholar
Cutler, A. H., and Flessa, K. W. 1990. Fossils out of sequence: computer simulations and strategies for dealing with stratigraphic disorder. Palaios 5:227235.Google Scholar
Cutler, A. H., and Flessa, K. W. 1995. Bioerosion, dissolution and precipitation as taphonomic agents at high and low latitudes. Senckenbergiana Maritima 25:115121.Google Scholar
Cutler, A. H., Behrensmeyer, A. K., and Chapman, R. E. 1999. Environmental information in a recent bone assemblage: roles of taphonomic processes and ecological change. Palaeogeography, Palaeoclimatology, Palaeoecology 149:359372.Google Scholar
Daley, G. M. 1993. Passive deterioration of shelly material: a study of the Recent eastern Pacific articulate brachiopod Terebratalia transversa Sowerby. Palaios 8:226232.Google Scholar
Daley, R. L., and Boyd, D. W. 1996. The role of skeletal microstructure during selective silicification of brachiopods. Journal of Sedimentary Research 66:155162.Google Scholar
Damuth, J. 1982. Analysis of the preservation of community structure in assemblages of fossil mammals. Paleobiology 8:434446.Google Scholar
Davies, D. J. 1993. Taphonomic analysis as a tool for long-term community baseline delineation: taphoanalysis in an environmental impact statement (EIS) for proposed human seafloor disturbances, Alabama continental shelf. Geological Society of America Abstracts with Programs 25:A459.Google Scholar
Davies, D. J., Powell, E. N., and Stanton, R. J. 1989. Taphonomic signature as a function of environmental process: shells and shell beds in a hurricane-influenced inlet on the Texas coast. Palaeogeography, Palaeoclimatology, Palaeoecology 72:317356.Google Scholar
Davies-Vollum, K. S., and Wing, S. L. 1998. Sedimentological, taphonomic, and climatic aspects of Eocene swamp deposits (Willwood Formation, Bighorn Basin, Wyoming). Palaios 13:2840.Google Scholar
Davis, P. G., and Briggs, D. E. G. 1998. The impact of decay and disarticulation on the preservation of fossil birds. Palaios 13:313.Google Scholar
Dawson, J. W. 1882. On the results of recent explorations of erect trees containing animal remains in the coal formation of Nova Scotia. Philosophical Transactions of the Royal Society of London B 173:621659.Google Scholar
Dechant Boaz, D. 1994. Taphonomy and the fluvial environment. Pp. 377413 in Corruccini, R. S. and Ciochon, R. L., eds. Integrative paths to the past: paleoanthropological advances in honor of F. Clark Howell. Prentice-Hall, Englewood Cliffs, N.J. Google Scholar
Demko, T. M., and Gastaldo, R. A. 1992. Paludal environments of the Lower Mary Lee coal zone, Pottsville Formation, Alabama: stacked clastic swamps and peat mires. International Journal of Coal Geology 20:2347.Google Scholar
Demko, T. M., Dubiel, R. F., and Parrish, J. T. 1998. Plant taphonomy in incised valleys: implications for interpreting paleoclimate from fossil plants. Geology 26:11191122.Google Scholar
Dent, S. R. 1995. A taphofacies model of the Recent South Florida continental shelf: a new perspective for a classic, exposed carbonate environment. Ph.D. dissertation. University of Cincinnati, Cincinnati, Ohio.Google Scholar
D'Hondt, S., Donaghay, P., Zachos, J. C., Luttenberg, D., and Lindinger, M. 1998. Organic carbon fluxes and ecological recovery from the Cretaceous-Tertiary mass extinction. Science 282:276279.Google Scholar
DiMichele, W. A., and Aronson, R. B. 1992. The Pennsylvanian-Permian vegetational transition: a terrestrial analogue to the onshore-offshore hypothesis. Evolution 46:807824.Google Scholar
DiMichele, W. A., and Hook, R. W. 1992. Paleozoic terrestrial ecosystems. Pp. 205325 in Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A., Potts, R., Sues, H.-D., and Wing, S. L., eds. Terrestrial ecosystems through time. University of Chicago Press, Chicago.Google Scholar
DiMichele, W. A., Mamay, S. H., Chaney, D. S., Hook, R. W., and Nelson, W. J. In press. An early Permian flora with Late Permian and Mesozoic affinities from north-central Texas. Journal of Paleontology.Google Scholar
Dodson, P. 1987. Microfaunal studies of dinosaur paleoecology, Judith River Formation of southern Alberta (Canada). Palaeogeography, Palaeoclimatology, Palaeoecology 10:2174.Google Scholar
Dominguez-Rodrigo, M. 1999. Flesh availability and bone modifications in carcasses consumed by lions: palaeoecological relevance in hominid foraging patterns. Palaeogeography, Palaeoclimatology, Palaeoecology 149:373388.Google Scholar
Donovan, S. K., ed. 1991. The processes of fossilization. Columbia University Press, New York.Google Scholar
Donovan, S. K., and Paul, C. R. C., eds. 1998. The adequacy of the fossil record. Wiley, New York.Google Scholar
Downing, K. F., and Park, L. E. 1998. Geochemistry and early diagenesis of mammal-bearing concretions from the Sucker Creek Formation (Miocene) of southeastern Oregon. Palaios 13:1427.Google Scholar
Doyle, P., and Macdonald, D. I. M. 1993. Belemnite battlefields. Lethaia 26:6580.Google Scholar
Droser, M. L., and Bottjer, D. J. 1993. Trends and patterns of Phanerozoic ichnofabrics. Annual Review of Earth and Planetary Sciences 21:205225.Google Scholar
Duncan, I. J., Briggs, D. E. G., and Archer, M. 1998. Three-dimensionally mineralized insects and millipedes from the Tertiary of Riversleigh, Queensland, Australia. Palaeontology 41:835851.Google Scholar
Eberth, D. A. 1990. Stratigraphy and sedimentology of vertebrate microfossil localities in uppermost Judith River Formation (Campanian) of Dinosaur Provincial Park, south-central Alberta, Canada. Palaeogeography, Palaeoclimatology, Palaeoecology 78:136.Google Scholar
Eberth, D., Rogers, R. and Fiorillo, T., convenors. 1999. Bonebeds: genesis, analysis, and paleoecological significance (program for a symposium). Journal of Vertebrate Paleontology 19(Suppl. to No. 3):7A.Google Scholar
Efremov, J. A. 1940. Taphonomy: new branch of paleontology. Pan American Geologist 74:8193.Google Scholar
Elder, R. L., and Smith, G. R. 1984. Fish taphonomy and paleoecology. Geobios Mémoire Spécial 8:287291.Google Scholar
Elias, S. A., Van Devender, T. R., and Baca, R. De. 1995. Insect fossil evidence of late glacial and Holocene environments in the Bolson de Mapimi, Chihuahuan Desert, Mexico: comparisons with the paleobotanical record. Palaios 10:454464.Google Scholar
Emig, C. C. 1990. Examples of post-mortality alteration in Recent brachiopod shells and (paleo)ecological consequences. Marine Biology 104:233238.Google Scholar
Evans, S., and Todd, J. A. 1997. Late Jurassic soft-bodied wood epibionts preserved by bioimmuration. Lethaia 30:185189.Google Scholar
Farley, M. B. 1987. Palynomorphs from surface water of the eastern and central Caribbean Sea. Micropaleontology 33:254262.Google Scholar
Fedonkin, M. A. 1994. Vendian body fossils and trace fossils. Pp. 370388 in Bengtson 1984.Google Scholar
Feige, A., and Fürsich, F. T. 1991. Taphonomy of the Recent molluscs of Bahía la Choya (Gulf of California, Sonora, Mexico). Zitteliana 18:89113.Google Scholar
Feldmann, R. M., Villamil, T., and Kauffman, E. G. 1999. Decapod and stomatopod crustaceans from mass mortality Lagerstätten: Turonian (Cretaceous) of Colombia. Journal of Paleontology 73:91101.Google Scholar
Ferguson, D. K. 1995. Plant part processing and community reconstruction. Eclogae Geologicae Helvetiae 88:627641.Google Scholar
Fernández-Jalvo, Y., Denys, C., Andrews, P., Williams, T., Dauphin, Y. and Humphreys, L. 1998. Taphonomy and paleoecology of Olduvai Bed-I (Pleistocene, Tanzania). Journal of Human Evolution 34:137172.Google Scholar
Fernández-López, S. 2000. Ammonite taphocycles in carbonate epicontinental platforms. Fifth international symposium on the Jurassic System (Vancouver, B.C.). GeoResearch Forum 6:293300.Google Scholar
Fiorillo, A. R. 1988. Taphonomy of Hazard Homestead Quarry (Ogalalla Group), Hitchcock County, Nebraska. Contributions to Geology University of Wyoming 26:5797.Google Scholar
Fiorillo, A. R. 1989. An experimental study of trampling: implications for the fossil record. Pp. 6172 in Bonnichsen, R. and Sorg, M. H., eds. Bone modification. Institute for Quaternary Studies, University of Maine, Orono.Google Scholar
Fiorillo, A. R. 1991. Taphonomy and depositional setting of Careless Creek Quarry (Judith River Formation), Wheatland County, Montana. Palaeogeography, Palaeoclimatology, Paleoecology 81:281311.Google Scholar
Flessa, K. W. 1998. Well-traveled cockles: shell transport during the Holocene transgression of the southern North Sea. Geology 26:187190.Google Scholar
Flessa, K. W., and Kowalewski, M. 1994. Shell survival and time-averaging in nearshore and shelf environments: estimates from the radiocarbon literature. Lethaia 27:153165.Google Scholar
Flessa, K. W., Cutler, A. H., and Meldahl, K. H. 1993. Time and taphonomy: quantitative estimates of time-averaging and stratigraphic disorder in a shallow marine habitat. Paleobiology 19:266286.Google Scholar
Foote, M. 1996. On the probability of ancestors in the fossil record. Paleobiology 22:141151.Google Scholar
Foote, M. 1997. Sampling, taxonomic description, and our evolving knowledge of morphological diversity. Paleobiology 23:181206.Google Scholar
Foote, M. 2000. Origination and extinction components of taxonomic diversity: general problems. In Erwin, D. H. and Wing, S. L., eds. Deep time: Paleobiology's perspective. Paleobiology 26(Suppl. to No. 4):74102.Google Scholar
Foote, M., and Raup, D. M. 1996. Fossil preservation and the stratigraphic ranges of taxa. Paleobiology 22:121140.Google Scholar
Foote, M., Hunter, J. P., Janis, C. M., and Sepkoski, J. J. Jr. 1999. Evolutionary and preservational constraints on origins of biologic groups: divergence times of Eutherian mammals. Science 283:13101314.Google Scholar
Franzen, J. L. 1985. Exceptional preservation of Eocene vertebrates in the lake deposits of Grube Messel (West Germany). Philosophical Transactions of the Royal Society of London B 311:181186.Google Scholar
Frison, G. C., and Todd, L. C. 1986. The Colby Mammoth Site: taphonomy and archaeology of a Clovis kill in northern Wyoming. University of New Mexico Press, Albuquerque.Google Scholar
Fürsich, F. T. 1978. The influence of faunal condensation and mixing on the preservation of fossil benthic communities. Lethaia 11:243250.Google Scholar
Fürsich, F. T., and Aberhan, M. 1990. Significance of time-averaging for paleocommunity analysis. Lethaia 23:143152.Google Scholar
Fürsich, F. T., and Oschmann, W. 1993. Shell beds as tools in basin analysis: the Jurassic of Kachchh, western India. Journal of the Geological Society, London 150:169185.Google Scholar
Gastaldo, R. A. 1988. A conspectus of phytotaphonomy. In DiMichele, W. A. and Wing, S. L., eds. Methods and applications of plant paleoecology: notes for a short course. Paleontological Society Special Publication 3:1428.Google Scholar
Gastaldo, R. A. 1992. Taphonomic considerations for plant evolutionary investigations. The Palaeobotanist 41:211223.Google Scholar
Gastaldo, R. A. 1994. The genesis and sedimentation of phytoclasts with examples from coastal environments. Pp. 103127 in Traverse, A., ed. Sedimentation of organic particles. Cambridge University Press, Cambridge.Google Scholar
Gastaldo, R. A., and Ferguson, D. K. 1998. Reconstructing Tertiary plant communities: introductory remarks. Review of Paleobotany and Palynology 101:36.Google Scholar
Gastaldo, R. A., and Huc, A. Y. 1992. Sediment facies, depositional environments, and distribution of phytoclasts in the Recent Mahakam River delta, Kalimantan, Indonesia. Palaios 7:574591.Google Scholar
Gastaldo, R. A., and Staub, J. R. 1999. A mechanism to explain the preservation of leaf litter lenses in coals derived from raised mires. Palaeogeography, Palaeoclimatology, Palaeoecology 149:114.Google Scholar
Gastaldo, R. A., Douglass, D. P., and McCarroll, S. M. 1987. Origin, characteristics, and provenance of plant macrodetritus in a Holocene crevasse splay, Mobile Delta, Alabama. Palaios 2:229240.Google Scholar
Gastaldo, R. A., Demko, T. M., and Liu, Y. 1993a. Application of sequence and genetic stratigraphic concepts to Carboniferous coal-bearing strata: an example from the Black Warrior Basin, USA. Geologische Rundschau 82:212226.Google Scholar
Gastaldo, R. A., Allen, G. P., and Huc, A. Y. 1993b. Detrital peat formation in the tropical Mahakam River delta, Kalimantan, eastern Borneo: formation, plant composition, and geochemistry. In Cobb, J. C. and Cecil, C. B., eds. Modern and ancient coal-forming environments. Geological Society of America Special Paper 286:107118.Google Scholar
Gastaldo, R. A., Walther, H., Rabold, J., and Ferguson, D. 1996. Criteria to distinguish parautochthonous leaves in Cenophytic alluvial channel-fills. Review of Palaeobotany and Palynology 91:121.Google Scholar
Gastaldo, R. A., Riegel, W., Püttmann, W., Linnemann, U. H., and Zetter, R. 1998. A multidisciplinary approach to reconstruct the Late Oligocene vegetation in central Europe. Review of Palaeobotany and Palynology 101:7194.Google Scholar
Gehling, J. G. 1999. Microbial mats in terminal Proterozoic siliciclastics: Ediacaran death masks. Palaios 14:4057.Google Scholar
Gifford-Gonzalez, D. 1991. Bones are not enough: analogues, knowledge, and interpretive strategies in zooarchaeology. Journal of Anthropological Archaeology 10:215254.Google Scholar
Gilinsky, N. L., and Bennington, J. B. 1994. Estimating numbers of whole individuals from collections of body parts: a taphonomic limitation of the paleontological record. Paleobiology 20:245258.Google Scholar
Gilinsky, N. L., and Signor, P. W., eds. 1991. Analytical paleobiology. Short Courses in Paleontology No. 4. Paleontological Society, Knoxville, Tenn.Google Scholar
Glover, C. P., and Kidwell, S. M. 1993. Influence of organic matrix on the post-mortem destruction of molluscan shells. Journal of Geology 101:729747.Google Scholar
Goldstein, S. T., Cappellen, P. Van, Roychoudhury, A., and Koretsky, C. 1997. Preservation of salt-marsh foraminifera in experimental arrays deployed below the sediment-water interface, Sapelo Island, Georgia (USA). Geological Society of America Abstracts with Programs 30:A383.Google Scholar
Grayson, D. K. 1989. Bone transport, bone destruction, and reverse utility curves. Journal of Archaeological Science 16:643652.Google Scholar
Greenstein, B. J. 1989. Mass mortality of the West-Indian echinoid Diadema antillarum (Echinodermata: Echinoidea): a natural experiment in taphonomy. Palaios 4:487492.Google Scholar
Greenstein, B. J. 1991. An integrated study of echinoid taphonomy: predictions for the fossil record of four echinoid families. Palaios 6:519540.Google Scholar
Greenstein, B. J. 1992. Taphonomic bias and the evolutionary history of the family Cidaridae (Echinodermata: Echinoidea). Paleobiology 18:5079.Google Scholar
Greenstein, B. J. 1993. Is the fossil record of regular echinoids really so poor a comparison of living and subfossil assemblages? Palaios 8:587601.Google Scholar
Greenstein, B. J. 1999. Taphonomy of reef-building corals II: shallow and deep reef environments of the tropical western Atlantic. Geological Society of America Abstracts with Programs 31:A420.Google Scholar
Greenstein, B. J., and Moffat, H. A. 1996. Comparative taphonomy of modern and Pleistocene corals, San Salvador, Bahama. Palaios 11:5763.Google Scholar
Greenstein, B. J., and Pandolfi, J. M. 1997. Preservation of community structure in modern reef coral life and death assemblages of the Florida Keys: implications for the Quaternary fossil record of coral reefs. Bulletin of Marine Science 61:431452.Google Scholar
Greenstein, B. J., Harris, L. A., and Curran, H. A. 1998. Comparison of Recent coral life and death assemblages to Pleistocene reef communities: implications for rapid faunal replacement of recent reefs. Carbonates and Evaporites 13:2331.Google Scholar
Grotzinger, J. P. 1994. Trends in Precambrian carbonate sediments and their implication for understanding evolution. Pp. 245258 in Bengtson 1984.Google Scholar
Hadly, E. A. 1999. Fidelity of terrestrial vertebrate fossils to a modern ecosystem. Palaeogeography, Palaeoclimatology, Palaeoecology 149:389410.Google Scholar
Hageman, S. J., James, N. P., and Bone, Y. 2000. Cool-water carbonate production from epizoic bryozoans on ephemeral substrates. Palaios 15:3348.Google Scholar
Haglund, W. D., and Sorg, M. H., eds. 1997. Forensic taphonomy, the post-mortem fate of human remains. CRC Press, New York.Google Scholar
Haynes, G. 1985. On watering holes, mineral licks, death, and predation. Pp. 5371 in Meltzer, D. and Mead, J. I., eds. Environments and extinctions in late glacial North America. Center for the Study of Early Man, University of Maine, Orono.Google Scholar
Haynes, G. 1988. Mass deaths and serial predation: comparative taphonomic studies of modern large-mammal deathsites. Journal of Archaeological Science 15:219235.Google Scholar
Haynes, G. 1991. Mammoths, mastodonts and elephants. Cambridge University Press, Cambridge.Google Scholar
Henderson, S. W., and Frey, R. W. 1986. Taphonomic redistribution of mollusk shells in a tidal inlet channel, Sapelo Island, Georgia. Palaios 1:316.Google Scholar
Henwood, A. A. 1992a. Exceptional preservation of dipteran flight muscle and the taphonomy of insects in amber. Palaios 7:203212.Google Scholar
Henwood, A. A. 1992b. Soft-part preservation of beetles in Tertiary amber from the Dominican Republic. Palaeontology 35:901912.Google Scholar
Holland, S. M. 1995. The stratigraphic distribution of fossils. Paleobiology 21:92109.Google Scholar
Holland, S. M. 1996. Recognizing artifactually generated coordinated stasis: implications of numerical models and strategies for field tests. Palaeogeography, Palaeoclimatology, Palaeoecology 127:147156.Google Scholar
Holland, S. M. 2000. The quality of the fossil record: a sequence stratigraphic perspective. In Erwin, D. H. and Wing, S. L., eds. Deep time: Paleobiology's perspective. Paleobiology 26(Suppl. to No. 4):148168.Google Scholar
Holland, S. M., and Patzkowsky, M. E. 1999. Models for simulating the fossil record. Geology 27:491494.Google Scholar
Hudson, J. 1993. From bones to behavior. Occasional Paper No. 21. Center for Archaeological Investigations, Southern Illinois University, Carbondale.Google Scholar
Hughes, N. C., and Cooper, D. L. 1999. Paleobiologic and taphonomic aspects of the “granulosa” trilobite cluster, Kope Formation (Upper Ordovician, Cincinnati region). Journal of Paleontology 73:306319.Google Scholar
Hughes, N. C., and Labandeira, C. C. 1995. The stability of species in taxonomy. Paleobiology 21:401403.Google Scholar
Hunter, J. 1994. Lack of a high body count at the K-T boundary. Journal of Paleontology 68:1158.Google Scholar
Jablonski, D. J. 1988. Estimates of species durations. Science 240:969.Google Scholar
Jablonski, D. J. 1999. The future of the fossil record. Science 284:21142116.Google Scholar
Jablonski, D. J. 2000. Micro- and macroevolution: scale and hierarchy in evolutionary biology and paleontology. In Erwin, D. H. and Wing, S. L., eds. Deep time: Paleobiology's perspective. Paleobiology 26(Suppl. to No. 4):1552.Google Scholar
Jablonski, D. J., Lidgard, S., and Taylor, P. D. 1997. Comparative ecology of bryozoan radiations: origin of novelties in cyclostomes and cheilostomes. Palaios 12:505523.Google Scholar
Jackson, S. T. 1989. Postglacial vegetational change along an elevational gradient in the Adirondack Mountains (New York): a study of plant macrofossils. New York State Museum Bulletin 465.Google Scholar
Jackson, S. T. 1994. Pollen and spores in Quaternary lake sediments as sensors of vegetation composition: theoretical models and empirical evidence. Pp. 253286 in Traverse, A., ed. Sedimentation of organic particles. Cambridge University Press, Cambridge.Google Scholar
Jackson, S. T., and Overpeck, J. T. 2000. Responses of plant populations and communities to environmental changes of the late Quaternary. In Erwin, D. H. and Wing, S. L., eds. Deep time: Paleobiology's perspective. Paleobiology 26(Suppl. to No. 4):194220.Google Scholar
Jackson, S. T., and Whitehead, D. R. 1991. Holocene vegetation patterns in the Adirondack Mountains. Ecology 72:641653.Google Scholar
Janzen, D. H. 1977. Why fruits rot, seeds mold, and meat spoils. American Naturalist 111:691713.Google Scholar
Jodry, M. A., and Stanford, D. J. 1992. Stewart's Cattle Guard Site: an analysis of bison remains in a Folsom kill-butchery campsite. Pp. 101168 in Stanford, D. J. and Day, J. S., eds. Ice Age hunters of the Rockies. Denver Museum of Natural History and University Press of Colorado, Denver. 378 pp.Google Scholar
Johnson, R. G. 1960. Models and methods for analysis of the mode of formation of fossil assemblages. Geological Society of America Bulletin 71:1051086.Google Scholar
Kah, L. C., and Knoll, A. H. 1996. Microbenthic distribution of Proterozoic tidal flats: environmental and taphonomic considerations. Geology 24:7982.Google Scholar
Kendricks, P., and Crane, P. R. 1997. The origin and early diversification of land plants: a cladistic study. Smithsonian Institution Press, Washington, D.C. Google Scholar
Kennish, M. J., and Lutz, R. A. 1999. Calcium carbonate dissolution rates in deep-sea bivalve shells on the East pacific Rise at 21°N: results of an 8-year in-situ experiment. Palaeogeography, Palaeoclimatology, Palaeoecology 154:293299.Google Scholar
Kerbis Peterhans, J. C., Wrangham, R. W., Carter, M. L., and Hauser, M. D. 1993. A contribution to tropical rain forest taphonomy: retrieval and documentation of chimpanzee remains from Kibale Forest, Uganda. Journal of Human Evolution 25:485514.Google Scholar
Kershaw, P. J., Swift, D. J., and Denoon, D. C. 1988. Evidence of recent sedimentation in the eastern Irish Sea. Marine Geology 85:114.Google Scholar
Kidston, R., and Lang, W. H. 1920. Old Red Sandstone plants showing structure, from the Rhynie Chert bed, Aberdeenshire, Part 2. Transactions of the Royal Society of Edinburgh 52:603627.Google Scholar
Kidwell, S. M. 1991. The stratigraphy of shell concentrations. Pp. 211290 in Allison and Briggs 1991.Google Scholar
Kidwell, S. M. 1993. Taphonomic expressions of sedimentary hiatus: field observations on bioclastic concentrations and sequence anatomy in low, moderate and high subsidence settings. Geologische Rundschau 82:189202.Google Scholar
Kidwell, S. M. 1998. Time-averaging in the marine fossil record: overview of strategies and uncertainties. Geobios 30:977995.Google Scholar
Kidwell, S. M. 1999. High fidelity of species relative abundances in marine molluscan death assemblages. Geological Society of America Abstracts with Programs 31:A419.Google Scholar
Kidwell, S. M. In press. Ecological fidelity of molluscan death assemblages. In Woodin, S. A., Aller, J. Y., and Aller, R. C., eds. Organism-sediment interactions. Belle Baruch Institute Volume. University of South Carolina Press, Columbia.Google Scholar
Kidwell, S. M., and Aigner, T. 1985. Sedimentary dynamics of complex shell beds: implications for ecologic and evolutionary patterns. Pp. 382395 in Bayer, U. and Seilacher, A., eds. Sedimentary and evolutionary cycles. Springer, Berlin.Google Scholar
Kidwell, S. M., and Baumiller, T. 1990. Experimental disintegration of regular echinoids: roles of temperature, oxygen and decay thresholds. Paleobiology 16:247271.Google Scholar
Kidwell, S. M., and Behrensmeyer, A. K., eds. 1993. Taphonomic approaches to time resolution in fossil assemblages. Short Courses in Paleontology No. 6. Paleontological Society, Knoxville, Tenn.Google Scholar
Kidwell, S. M., and Bosence, D. W. J. 1991. Taphonomy and time-averaging of marine shelly faunas. Pp. 115209 in Allison and Briggs 1991.Google Scholar
Kidwell, S. M., and Brenchley, P. J. 1994. Patterns in bioclastic accumulation through the Phanerozoic: changes in input or in destruction? Geology 22:11391143.Google Scholar
Kidwell, S. M., and Brenchley, P. J. 1996. Evolution of the fossil record: thickness trends in marine skeletal accumulations and their implications. Pp. 290336 in Jablonski, D., Erwin, D. H., and Lipps, J. H., eds. Evolutionary paleobiology. University of Chicago Press, Chicago.Google Scholar
Kidwell, S. M., and Flessa, K. W. 1995. The quality of the fossil record: populations, species, and communities. Annual Review of Ecology and Systematics 26:269299.Google Scholar
Kidwell, S. M., and Jablonski, D. 1983. Taphonomic feedback: ecological consequences of shell accumulation. Pp. 195248 in Tevesz, M. J. S. and McCall, P. L., eds. Biotic interactions in recent and fossil benthic communities. Plenum, New York.Google Scholar
Knoll, A. H. 1985. Exceptional preservation of photosynthetic organisms in silicified carbonates and silicified peats. Philosophical Transactions of the Royal Society of London B 311:111122.Google Scholar
Knoll, A. H., and Sergeev, V. N. 1995. Taphonomic and evolutionary changes across the Mesoproterozoic-Neoproterozoic transition. Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 195:289302.Google Scholar
Knoll, A. H., Niklas, K., Gensel, P. G., and Tiffney, B. 1984. Character diversification and patterns of evolution in early vascular plants. Paleobiology 10:3447.Google Scholar
Knoll, A. H., Fairchild, I. J., and Swett, K. 1993. Calcified microbes in Neoproterozoic carbonates: implications for our understanding of the Proterozoic/Cambrian transition. Palaios 8:512525.Google Scholar
Kondo, Y., Abbott, S. T., Katamura, A., Kamp, P. J. J., Naish, T. R. E., Kamataki, T., and Saul, G. S. 1998. The relationship between shellbed type and sequence architecture: examples from Japan and New Zealand. Sedimentary Geology 122:109127.Google Scholar
Kotier, E., Martin, R. E., and Liddell, W. D. 1992. Experimental analysis of abrasion and dissolution resistance of modern reef-dwelling Foraminifera: implications for the preservation of biogenic carbonate. Palaios 7:244276.Google Scholar
Kowalewski, M. 1996a. Taphonomy of a living fossil: the lingulide brachiopod Glottidia palmeri Dall from Baja California, Mexico. Palaios 11:244265.Google Scholar
Kowalewski, M. 1996b. Time-averaging, overcompleteness, and the geological record. Journal of Geology 104:317326.Google Scholar
Kowalewski, M. 1997. The reciprocal taphonomic model. Lethaia 30:8688.Google Scholar
Kowalewski, M., and Flessa, K. W. 1996. Improving with age: the fossil record of lingulide brachiopods and the nature of taphonomic megabiases. Geology 24:977980.Google Scholar
Kowalewski, M., Goodfriend, G. A., and Flessa, K. W. 1998. High-resolution estimates of temporal mixing within shell beds: the evils and virtues of time-averaging. Paleobiology 24:287304.Google Scholar
Kranz, P. M. 1974. The anastrophic burial of bivalves and its palaeoecological significance. Journal of Geology 82:237265.Google Scholar
Kristensen, E., Ahmed, S. I., and Devol, A. H. 1995. Aerobic and anaerobic decomposition of organic matter in marine sediment: which is fastest? Limnology and Oceanography 40:14301437.Google Scholar
Labandeira, C. C. 1998. Early history of arthropod and vascular plant associations. Annual Review of Earth Planetary Sciences 28:153193.Google Scholar
Labandeira, C. C., and Smith, D. M. 1999. Forging a future for fossil insects: thoughts on the First International Congress of Paleoentomology. Paleobiology 25:154157.Google Scholar
Labandeira, C. C., Johnson, K. R., and Lang, P. In press. Insect herbivory across the Cretaceous/Tertiary boundary: major extinction and minimum rebound. In Hartman, J. H., Johnson, K. R., and Nichols, D. J., eds. The Hell Creek Formation and the Cretaceous-Tertiary boundary in the northern Great Plains—an integrated Continental record at the end of the Cretaceous. Geological Society of America Special Paper.Google Scholar
Lask, P. B. 1993. The hydrodynamic behavior of sclerites from the trilobite Flexicalymene meeki. Palaios 8:219225.Google Scholar
Lescinsky, H. L. 1993. Taphonomy and paleoecology of epibionts on the scallops Chlamys hastata (Sowerby 1843) and Chlamys rubida (Hinds 1845). Palaios 8:267277.Google Scholar
Lescinsky, H. L. 1995. The life orientation of concavo-convex brachiopods: overturning the paradigm. Paleobiology 21:520551.Google Scholar
Li, X., and Droser, M. L. 1997. Nature and distribution of Cambrian shell concentrations: evidence from the Basin and Range province of western United States (California, Nevada and Utah). Palaios 12:111126.Google Scholar
Llewellyn, G., and Messing, C. G. 1993. Compositional and taphonomic variations in modern crinoid-rich sediments from the deep-water margin of a carbonate bank. Palaios 8:554573.Google Scholar
Llona, A., Pinto, C., and Andrews, P. 1999. Amphibian taphonomy and its application to the fossil record of Dolina (middle Pleistocene, Atapuerca, Spain). Palaeogeography, Palaeoclimatology, Palaeoecology 149:411430.Google Scholar
Lupia, R. 1995. Paleobotanical data from fossil charcoal: an actualistic study of seed plant reproductive structures. Palaios 10:465477.Google Scholar
Lyell, C., and Dawson, J. W. 1853. On the remains of a reptile (Dendrerpeton acadianum, Wyman and Owen), and of a land shell discovered in the interior of an erect fossil tree in the coal measures of Nova Scotia. Quarterly Journal of the Geological Society of London IX:5863.Google Scholar
Lyman, R. L. 1985. Bone frequencies: differential transport, in situ destruction, and the MGUI. Journal of Archaeological Science 12:221236.Google Scholar
Lyman, R. L. 1994. Vertebrate taphonomy. Cambridge Manuals in Archaeology, Cambridge University Press, Cambridge.Google Scholar
Lyman, R. L., and Fox, G. L. 1989. A critical evaluation of bone weathering as an indication of bone assemblage formation. Journal of Archaeological Science 16:293317.Google Scholar
Malinky, J. M., and Heckel, P. H. 1998. Paleoecology and taphonomy of faunal assemblages in gray “core” (offshore) shales in Midcontinent Pennsylvanian cyclothems. Palaios 13:311334.Google Scholar
Maliva, R. G., Knoll, A. H., and Siever, R. 1989. Secular change in chert distribution: a reflection of evolving biological participation in the silica cycle. Palaios 4:519532.Google Scholar
Marean, C. W. 1991. Measuring the post-depositional destruction of bone in archaeological assemblages. Journal of Archaeological Science 18:677694.Google Scholar
Marean, C. W. 1992. Captive hyaena bone choice and destruction, the Schlepp effect and Olduvai archaeofaunas. Journal of Archaeological Science 19:101121.Google Scholar
Markwick, P. J. 1998. Crocodilian diversity in space and time: the role of climate in paleoecology and its implications for understanding K/T extinctions. Paleobiology 24:470497.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 20:459460.Google Scholar
Martill, D. M. 1985. The preservation of marine vertebrates in the Lower Oxford Clay (Jurassic) of central England. Philosophical Transactions of the Royal Society of London B 311:155165.Google Scholar
Martill, D. M. 1988. Preservation of fish in the Cretaceous Santana Formation of Brazil. Palaeontology 31:118.Google Scholar
Martill, D. M. 1990. Macromolecular resolution of fossilized muscle tissue from an elopomorph fish. Nature 346:171172.Google Scholar
Martin, R. E. 1993. Time and taphonomy: actualistic evidence for time-averaging of benthic foraminiferal assemblages. Pp. 3456 in Kidwell and Behrensmeyer 1993.Google Scholar
Martin, R. E. 1995. Cyclic and secular variation in microfossil biomineralization: clues to the biogeochemical evolution of Phanerozoic oceans. Global and Planetary Change 11:123.Google Scholar
Martin, R. E. 1999. Taphonomy, a process approach. Cambridge University Press, Cambridge.Google Scholar
Martin, R. E., Wehmiller, J. F., Harris, M. S., and Liddell, W. D. 1996. Comparative taphonomy of bivalves and foraminifera from Holocene tidal flat sediments, Bahía la Choya, Sonora, Mexico (Northern Gulf of California): taphonomic grades and temporal resolution. Paleobiology 22:8090.Google Scholar
Martin, R. E., Patterson, R. T., Goldstein, S. T., and Kumar, A., eds. 1999a. Taphonomy as a tool in paleoenvironmental reconstruction and environmental assessment. Palaeogeography, Palaeoclimatology, Palaeoecology 149(special issue).Google Scholar
Martin, R. E., Hippensteel, S. P., Nikitina, D., and Pizzuto, J. E. 1999b. Artificial time-averaging and the recovery of ecological signals preserved in the subfossil record: linking the temporal scales of ecology and paleoecology. Geological Society of America Abstracts with Programs: A356.Google Scholar
Martinez-Delclos, X., and Martinell, J. 1993. Insect taphonomy experiments: their application to the Cretaceous outcrops of lithographic limestones from Spain. Kaupia (Darmstädter Beiträge zur Naturgeschichte) 2:133144.Google Scholar
Massé, H. L. 1999. Les carbonates associés a la macrofaune des sables fins littoraux en Méditerranée nord-occidentale. Oceanologica Acta 22:413420.Google Scholar
Maxwell, W. D., and Benton, M. J. 1990. Historical tests of the absolute completeness of the fossil record of tetrapods. Paleobiology 16:322335.Google Scholar
McGree, H. 1984. On food and cooking: the science and lore of the kitchen. Scribner, New York.Google Scholar
McIlroy, D., and Logan, G. A. 1999. The impact of bioturbation on infaunal ecology and evolution during the Proterozoic-Cambrian transition. Palaios 14:5872.Google Scholar
McKinney, F. K. 1996. Encrusting organisms on co-occurring disarticulated valves of two marine bivalves: comparison of living assemblages and skeletal residues. Paleobiology 222:534567.Google Scholar
McKinney, M. L. 1991. Completeness of the fossil record: an overview. Pp. 6683 in Donovan 1991.Google Scholar
Meldahl, K. H. 1990. Sampling, species abundance, and the stratigraphic signature of mass extinction: a test using Holocene tidal flat molluscs. Geology 18:890893.Google Scholar
Meldahl, K. H., and Flessa, K. W. 1990. Taphonomic pathways and comparative biofacies and taphofacies in a Recent intertidal / shallow shelf environment. Lethaia 23:4360.Google Scholar
Meldahl, K. H., Scott, D., and Carney, K. 1995. Autochthonous leaf assemblages as records of deciduous forest communities: an actualistic study. Lethaia 28:383394.Google Scholar
Meldahl, K. H., Flessa, K. W., and Cutler, A. H. 1997a. Time-averaging and postmortem skeletal survival in benthic fossil assemblages: quantitative comparisons among Holocene environments. Paleobiology 23:209229.Google Scholar
Meldahl, K. H., Yajimovich, O. G., Empedocles, C. D., Gustafson, C. S., Hidalgo, M. M., and Reardon, T. W. 1997b. Holocene sediments and molluscan faunas of Bahía Concepcíon: a modern analog to Neogene rift basins of the Gulf of California. Geological Society of America Special Paper 318:3956.Google Scholar
Meyer, D. L., and Meyer, K. B. 1986. Biostratinomy of Recent crinoids (Echinodermata) at Lizard Island, Great Barrier Reef, Australia. Palaios 1:294302.Google Scholar
Meyer, D. L., Ausich, W. I., and Terry, R. E. 1989. Comparative taphonomy of echinoderms in carbonate facies: Fort Payne Formation (Lower Mississippian) of Kentucky and Tennessee. Palaios 4:533552.Google Scholar
Miller, A. I. 1988. Spatial resolution in subfossil molluscan remains: implications for paleobiological analyses. Paleobiology 14:91103.Google Scholar
Miller, A. I. 1997. Dissecting global diversity patterns: examples from the Ordovician radiation. Annual Review of Ecology and Systematics 28:85104.Google Scholar
Miller, A. I., and Cummins, H. 1990. A numerical model for the formation of fossil assemblages: estimating the amount of post-mortem transport along environmental gradients. Palaios 5:303316.Google Scholar
Miller, A. I., and Cummins, H. 1993. Using numerical models to evaluate the consequences to time-averaging in marine fossil assemblages. Pp. 150168 in Kidwell and Behrensmeyer 1993.Google Scholar
Miller, A. I., and Foote, M. 1996. Calibrating the Ordovician radiation of marine life: implications for Phanerozoic diversity trends. Paleobiology 22:304309.Google Scholar
Miller, A. I., Llewellyn, G., Parsons, K. M., Cummins, H., Boardman, M. R., Greenstein, B. J., and Jacobs, D. K. 1992. Effect of Hurricane Hugo on molluscan skeletal distributions, Salt River Bay, St. Croix, U. S. Virgin Islands. Geology 20:2326.Google Scholar
Mirsky, S. 1998. I shall return. Earth 7:4853.Google Scholar
Morales, M., ed. 1996. The continental Jurassic. Museum of Northern Arizona Bulletin 60.Google Scholar
Nebelsick, J. H. 1992. Echinoid distribution by fragment identification in the northern Bay of Safaga, Red Sea, Egypt. Palaios 7:316328.Google Scholar
Nebelsick, J. H. 1995. Comparative taphonomy of clypeasteroids. Eclogae Geologica Helvetiae 88:685693.Google Scholar
Nebelsick, J. H. 1999. Taphonomy of Clypeaster fragments: preservation and taphofacies. Lethaia 32:241252.Google Scholar
Noe-Nygaard, N. 1987. Taphonomy in archaeology, with special emphasis on man as a biasing factor. Journal of Danish Archaeology 6:762.Google Scholar
Norris, R. D. 1986. Taphonomic gradients in shelf fossil assemblages: Pliocene Purisima Formation, California. Palaios 1:256270.Google Scholar
Oliver, J. S., and Graham, R. W. 1994. A catastrophic kill of ice-trapped coots: time-averaged versus scavenger-specific disarticulation patterns. Paleobiology 20:229244.Google Scholar
Oliver, J. S., Sikes, N. E., and Stewart, K. M., eds. 1994. Early hominid behavioural ecology. Academic Press, London.Google Scholar
Olson, E. C. 1966. Community evolution and the origin of mammals. Ecology 47:291302.Google Scholar
Olszewski, T. D. 1999. Taking advantage of time-averaging. Paleobiology 25:226238.Google Scholar
Olszewski, T. D., and West, R. R. 1997. Influence of transportation and time-averaging in fossil assemblages from the Pennsylvanian of Oklahoma. Lethaia 30:315329.Google Scholar
Oost, A. P., and de Boer, P. L. 1994. Tectonic and climatic setting of lithographic limestone basins. Geobios Mémoire Spécial 16:321330.Google Scholar
Orr, P. J., Briggs, D. E. G., and Kearns, S. L. 1998. Cambrian Burgess Shale animals replicated in clay minerals. Science 281:11731175.Google Scholar
Palaios. 1999. Unexplored microbial worlds (theme issue). Vol. 141 Palaios. 1999. Unexplored microbial worlds (theme issue). Vol. 14, No. 1.Google Scholar
Palmqvist, P. 1991. Differences in the fossilization potential of bivalve and gastropod species related to their life sites and trophic resources. Lethaia 24:287288.Google Scholar
Palmqvist, P. 1993. Trophic levels and the observational completeness of the fossil record. Revista Española de Paleontología 8:3336.Google Scholar
Pandolfi, J. M., and Greenstein, B. J. 1997a. Taphonomic alteration of reef coral: effects of reef environment and coral growth form. I. The Great Barrier Reef. Palaios 12:2742.Google Scholar
Palmqvist, P. 1997b. Preservation of community structure in death assemblages of deep-water Caribbean reef corals. Limnology and Oceanography 42:15051516.Google Scholar
Parsons, K. M. 1989. Taphonomy as an indicator of environment: Smuggler's Cove, St. Croix, U.S.V.I. In Hubbard, D. K., ed. Terrestrial and marine ecology of St. Croix, U.S. Virgin Islands. West Indies Laboratory Special Publication 8:135143.Google Scholar
Parsons, K. M., and Brett, C. E. 1991. Taphonomic processes and biases in modern marine environments: an actualistic perspective on fossil assemblage preservation. Pp. 2265 in Donovan 1991.Google Scholar
Parsons, K. M., Brett, C. E., and Miller, K. B. 1988. Taphonomy and depositional dynamics of Devonian shell-rich mudstones. Palaeogeography, Palaeoclimatology, Palaeoecology 63:109140.Google Scholar
Parsons-Hubbard, K. M., Callender, W. R., Powell, E. N., Brett, C. E., Walker, S. E., Raymond, A. L., and Staff, G. M. 1999. Rates of burial and disturbance of experimentally-deployed molluscs: implications for preservation potential. Palaios 14:337351.Google Scholar
Paul, C. R. C. 1992. How complete does the fossil record have to be? Revista Española Paleontología 7:127133.Google Scholar
Paul, C. R. C. 1998. Adequacy, completeness and the fossil record. Pp. 122 in Donovan, S. K. and Paul, C. R. C., eds. The adequacy of the fossil record. Wiley, New York.Google Scholar
Perry, C. T. 1996. The rapid response of reef sediments to changes in community composition: implications for time averaging and sediment accumulation. Journal of Sedimentary Research 66:459467.Google Scholar
Perry, C. T. 1999. Reef framework preservation in four contrasting modern reef environments, Discovery Bay, Jamaica. Journal of Coastal Research 15:796812.Google Scholar
Peterson, C. H. 1977. The paleoecological significance of undetected short-term temporal variability. Journal of Paleontology 51:976981.Google Scholar
Plummer, T., and Kinyua, A. M. 1994. Provenancing of hominid and mammalian fossils from Kanjera, Kenya, using EDXRF. Journal of Archaeological Science 21:553563.Google Scholar
Potts., R. 1986. Temporal span of bone accumulations at Olduvai Gorge and implications for early hominid foraging behavior. Paleobiology 12:2531.Google Scholar
Potts., R. 1988. Early hominid activities at Olduvai. Aldyne de Gruyter, New York.Google Scholar
Potts, R., Behrensmeyer, A. K., and Ditchfield, P. 1999. Paleolandscape variation and early Pleistocene hominid activities: Members 1 and 7, Olorgesailie Formation. Journal of Human Evolution 37:747788.Google Scholar
Poulicek, M., Goffinet, G., Jeuniaux, C., Simon, A., and Voss-Foucart, M. F. 1988. Early diagenesis of skeletal remains in marine sediments: a 10 years study. Actes Colloque Recherches Océanographiques en Mer Méditerranée, Université Etat, Liege 107–124.Google Scholar
Powell, E. N., and Davies, D. J. 1990. When is an “old” shell really old? Journal of Geology 98:823844.Google Scholar
Powell, E. N., Staff, G., Davies, D. J., and Callender, W. R. 1989. Macrobenthic death assemblages in modern marine environments: formation, interpretation and application. CRC Critical Reviews in Aquatic Science 1:555589.Google Scholar
Powell, E. N., Stanton, R. J. Jr., Logan, A., and Craig, M. A. 1992. Preservation of Mollusca in Copano Bay, Texas. The long-term record. Palaeogeography, Palaeoclimatology, Palaeoecology 95:209228.Google Scholar
Prager, E. J., Southard, J. B., and Vivoni-Gallart, E. R. 1996. Experiments on the entrainment threshold of well-sorted and poorly sorted carbonate sands. Sedimentology 43:3340.Google Scholar
Pratt, B. R. 1998. Probable predation on Upper Cambrian trilobites and its relevance for the extinction of soft-bodied Burgess Shale-type animals. Lethaia 31:7388.Google Scholar
Rasmussen, K. A., and Brett, C. E. 1985. Taphonomy of Holocene cryptic biotas from St. Croix, Virgin Islands: information loss and preservation biases. Geology 13:551553.Google Scholar
Raup, D. M. 1975. Taxonomic diversity estimation using rarefaction. Paleobiology 1:333342.Google Scholar
Raymond, A., and Metz, C. 1995. Laurussian land-plant diversity during the Silurian and Devonian: mass extinction, sampling bias, or both? Paleobiology 21:7491.Google Scholar
Retallack, G. J. 1990. Soils of the past: an introduction to paleopedology. Unwin Hyman, London.Google Scholar
Retallack, G. J., and Dilcher, D. L. 1986. Reconstructions of selected seed ferns. Annals of the Missouri Botanical Garden 75:10101057.Google Scholar
Rich, F. J. 1989. A review of the taphonomy of plant remains in lacustrine sediments. Review of Palaeobotany and Palynology 58:3346.Google Scholar
Richmond, D. R., and Morris, T. H. 1996. The dinosaur deathtrap of the Cleveland-Lloyd Dinosaur Quarry, Emery County, Utah. Pp. 533546 in Morales 1996.Google Scholar
Rigby, J. K. Jr., Newman, K. R., Smit, J., Van Der Kaars, S., Sloan, R. E., and Rigby, J. K. 1987. Dinosaurs from the Paleocene part of the Hell Creek Formation, McCone County, Montana. Palaios 2:296302.Google Scholar
Rivas, P., Aguirre, J., and Braga, J. C. 1997. Entolium beds: hiatal shell concentrations in starved pelagic settings (middle Liaassic, SE Spain). Eclogae Geologica Helvetiae 90:293301.Google Scholar
Robinson, J. M. 1990. Lignin, land plants, and fungi. Biological evolution affecting Phanerozoic oxygen balance. Geology 15:607610.Google Scholar
Robinson, J. M. 1991. Phanerozoic atmospheric reconstructions: a terrestrial perspective. Palaeogeography, Palaeoclimatology, Palaeoecology 97:5162.Google Scholar
Rogers, R. R. 1990. Taphonomy of three dinosaur bone beds in the Upper Cretaceous Two Medicine Formation of northwestern Montana: evidence for drought-related mortality. Palaios 5:394413.Google Scholar
Rogers, R. R. 1993. Systematic patterns of time-averaging in the terrestrial vertebrate record: a Cretaceous case study. Pp. 228249 in Kidwell, S. M. and Behrensmeyer, A. K., eds. Taphonomic approaches to time resolution in fossil assemblages. Short Courses in Paleontology 6 (Paleontological Society).Google Scholar
Rogers, R. R., and Kidwell, S. M. 2000. Associations of vertebrate skeletal concentrations and discontinuity surfaces in continental and shallow marine records: a test in the Cretaceous of Montana. Journal of Geology 108:131154.Google Scholar
Rolfe, W. D. I., Clarkson, E. N. K., and Panchen, A. L., eds. 1994 (1993). Volcanism and early terrestrial biotas. Transactions of the Royal Society of Edinburgh (Earth Sciences) 84.Google Scholar
Roopnarine, P. D. 1999. Breaking the enigma of stratophenetic series: a computational approach to the analysis of microevolutionary mode. Geological Society of America Abstracts with Programs 31(7):A42.Google Scholar
Roopnarine, P. D., Byars, G., and Fitzgerald, P. 1999. Anagenetic evolution, stratophenetic patterns, and random walk models. Paleobiology 25:4157.Google Scholar
Roy, K., Valentine, J. W., Jablonski, D., and Kidwell, S. M. 1996. Scales of climatic variability and time averaging in Pleistocene biotas: implications for ecology and evolution. Trends in Ecology and Evolution 11:458463.Google Scholar
Sadler, P. M. 1981. Sediment accumulation and the completeness of stratigraphic sections. Journal of Geology 89:569584.Google Scholar
Sander, P. M. 1989. Early Permian depositional environments and pond bonebeds in central Archer County, Texas. Palaeogeography, Palaeoclimatology, Palaeoecology 69:121.Google Scholar
Sanders, H. L. 1968. Marine benthic diversity: a comparative study. American Naturalist 102:243282.Google Scholar
Schaal, S., and Ziegler, W., eds. 1992. Messel: an insight into the history of life and of the Earth. Translated by M. Shaffer-Fehre. Clarendon, Oxford.Google Scholar
Schäfer, W. 1972. Ecology and paleoecology of marine environments. University of Chicago Press, Chicago.Google Scholar
Scheihing, M. 1980. Reduction of wind velocity by the forest canopy and the rarity of non-arborescent plants in the Upper Carboniferous fossil record. Augumenta Palaeobotanica 6:133138.Google Scholar
Scheihing, M. H., and Pfefferkorn, H. W. 1984. The taphonomy of land plants in the Orinoco Delta: a model for the incorporation of plant parts in clastic sediments of late Carboniferous age of Euramerica. Review of Palaeobotany and Palynology 41:205240.Google Scholar
Schmude, D. E., and Weege, C. J. 1996. Stratigraphic relationship, sedimentology, and taphonomy of Meilyn, a dinosaur quarry in the basal Morrison Formation of Wyoming. Pp. 547554 in Morales 1996.Google Scholar
Schopf, T. J. M. 1978. Fossilization potential of an intertidal fauna: Friday Harbor, Washington. Paleobiology 4:261270.Google Scholar
Schubert, J. K., Kidder, D. L., and Erwin, D. H. 1997. Silica-replaced fossils through the Phanerozoic. Geology 25:10311034.Google Scholar
Scoffin, T. P. 1992. Taphonomy of coral reefs: a review. Coral Reefs 11:5777.Google Scholar
Scott, A. C. 1990. Anatomical preservation of fossil plants. Pp. 263266 in Briggs and Crowther 1990.Google Scholar
Scott, R. W. 1978. Approaches to trophic analysis of paleocommunities. Lethaia 11:114.Google Scholar
Seilacher, R. W. 1984. Late Precambrian and Early Cambrian metazoa: preservational or real extinctions? Pp. 159168 in Holland, H. D. and Trendall, A. F., eds. Patterns of change in earth evolution. Springer, Berlin.Google Scholar
Seilacher, R. W. 1985. The Jeram Model: event condensation in a modern intertidal environment. Pp. 336341 in Bayer, U. and Seilacher, A., eds. Sedimentary and evolutionary cycles. Springer, Berlin.Google Scholar
Seilacher, R. W. 1994. Early multicellular life: late Proterozoic fossils and the Cambrian extinction. Pp. 389400 in Bengtson 1984.Google Scholar
Seilacher, A., Reif, W. E., and Westphal, F. 1985. Sedimentological, ecological, and temporal patterns of fossil Lagerstätten. Philosophical Transactions of the Royal Society of London B 311:523.Google Scholar
Sepkoski, J. J. Jr. 1978. A kinetic model of Phanerozoic taxonomic diversity: I. Analysis of marine orders. Paleobiology 4:223251.Google Scholar
Sepkoski, J. J. Jr. 1990. The taxonomic structure of periodic extinction. In Sharpton, V. and Ward, P., eds. Global catastrophes in earth history. Geological Society of America Special Paper 247:3344.Google Scholar
Sepkoski, J. J. Jr. 1993. Ten years in the library: new data confirm paleontological patterns. Paleobiology 19:4351.Google Scholar
Sepkoski, J. J. Jr., and Koch, C. F. 1996. Evaluating paleontologic data relating to bio-events. Pp. 2134 in Walliser, O. H., ed. Global events and event stratigraphy in the Phanerozoic. Springer, 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
Sepkoski, J. J. Jr., Bambach, R. K., and Droser, M. L. 1991. Secular changes in Phanerozoic event bedding and the biological overprint. Pp. 298312 in Einsele, G., Ricken, W., and Seilacher, A., eds. Cycles and events in stratigraphy. Springer, Berlin.Google Scholar
Sept, J. M. 1994. Bone distribution in a semi-arid riverine habitat in eastern Zaire: implications for the interpretation of faunal assemblages at early archaeological sties. Journal of Archaeological Science 21:217235.Google Scholar
Silva de Echols, C. M. H. M. 1993. Diatom infestation of Recent crinoid ossicles in temperate waters, Friday harbor Laboratories, Washington: implications for biodegradation of skeletal carbonates. Palaios 8:278288.Google Scholar
Simões, M. G., Marques, A. C., Mello, L. H. C., and Ghilardi, R. P. 2000a. The role of taphonomy in cladistic analysis: a case study in Permian bivalves. Revista Española de Paleontología 15:153164.Google Scholar
Simões, M. G., Kowalewski, M., Torello, F. F., Ghilardi, R. P., and Mello, L. H. C. 2000b. Early onset of modern-style shell beds in the Permian sequences of the Paraná Basin: implications for the Phanerozoic trend in bioclastic accumulations. Revista Brasileira de Geociěncias 30:495499.Google Scholar
Smith, A. M., and Nelson, C. S. 1994. Selectivity in sea-floor processes: taphonomy of bryozoans. Pp. 177180 in Hayward, P. J., Ryland, J. S., and Taylor, P. D., eds. Biology and paleobiology of bryozoans. Proceedings of the ninth international bryozoology conference, 1992. Olsen and Olsen, Fredensborg, Denmark.Google Scholar
Smith, D. M. 2000. Beetle taphonomy in a recent ephemeral lake, southeastern Arizona. Palaios 15:152160.Google Scholar
Smith, G. R., Stearley, R. F., and Badgley, C. E. 1988. Taphonomic bias in fish diversity from Cenozoic floodplain environments. Palaeogeography, Paleoclimatology, Palaeoecology 63:263273.Google Scholar
Smith, R. M. H. 1993. Vertebrate taphonomy of Late Permian floodplain deposits in the southwestern Karoo Basin of South Africa. Palaios 8:4567.Google Scholar
Smith, R. M. H. 1995. Changing fluvial environments across the Permian-Triassic boundary in the Karoo Basin, South Africa and possible causes of tetrapod extinctions. Palaeogeography, Palaeoclimatology, Palaeoecology 117:81104.Google Scholar
Smith, R. M. H., and Kitching, J. 1996. Sedimentology and vertebrate taphonomy of the Tritylodon Acme Zone: a reworked palaeosol in the Lower Jurassic Elliot Formation, Karoo Supergroup, South Africa. Pp. 531532 in Morales 1996.Google Scholar
Solomon, S., Davidson, I., and Watson, D. 1990. Problem solving in taphonomy. Tempus: archaeology and material culture studies in anthropology, Vol. 2. University of Queensland, St. Lucia, Queensland, Australia.Google Scholar
Speyer, S. E., and Brett, C. E. 1986. Trilobite taphonomy and Middle Devonian taphofacies. Palaios 1:312327.Google Scholar
Speyer, S. E., and Brett, C. E. 1991. Taphofacies controls: background and episodic processes in fossil assemblage preservation. Pp. 501545 in Allison and Briggs 1991.Google Scholar
Spicer, R. A. 1980. The importance of depositional sorting to the biostratigraphy of plant megafossils. Pp. 171183 in Dilcher, D. L. and Taylor, T. N., eds. Biostratigraphy of fossil plants. Dowdon, Hutchinson, and Ross, New York.Google Scholar
Spicer, R. A. 1991. Plant taphonomic processes. Pp. 71113 in Allison and Briggs 1991.Google Scholar
Stachowitsch, M. 1984. Mass mortality in the Gulf of Trieste: the course of community destruction. Marine Ecology (Publicazioni della Stazione Zoologica di Napoli) 5:243264.Google Scholar
Staff, G. M., Powell, E. N., Stanton, R. J. Jr., and Cummins, H. 1985. Biomass: is it a useful tool in paleocommunity reconstruction? Lethaia 18:209232.Google Scholar
Staff, G. M., Stanton, R. J. Jr., Powell, E. N., and Cummins, H. 1986. Time averaging, taphonomy and their impact on paleocommunity reconstruction: death assemblages in Texas bays. Geological Society of America Bulletin 97:428443.Google Scholar
Stankiewicz, B. A., Poinar, H. N., Briggs, D. E. G., Evershed, R. P., and Poinar, G. O. Jr. 1998. Chemical preservation of plants and insects in natural resins. Proceedings of the Royal Society of London B 265:641647.Google Scholar
Stanley, S. E., and Hardie, L. A. 1998. Secular oscillations in the carbonate mineralogy of reef-building and sediment-producing organism driven by tectonically forced shifts in seawater chemistry. Palaeogeography, Palaeoclimatology, Palaeoecology 144:319.Google Scholar
Stewart, K. M., Lebranc, L., Matthiesen, D. P. and West, J. 1999. Microfaunal remains from a modern east African raptor roost: patterning and implications for fossil bone scatters. Paleobiology 25:483503.Google Scholar
Stoner, A. W., and Ray, M. 1996. Shell remains provide clues to historical distribution and abundance patterns in a large sea-grass-associated gastropod (Stombus gigas). Marine Ecology Progress Series 135:101108.Google Scholar
Strauss, D., and Sadler, P. M. 1989. Classical confidence intervals and Bayesian probability estimates for ends of local taxon ranges. Mathematical Geology 21:411427.Google Scholar
Sutcliffe, A. J. 1990. Rates of decay of mammalian remains in the permafrost environment of the Canadian High Arctic. Pp. 161186 in Harrington, C. R., ed. Canada's missing dimension: science and history in the Canadian Arctic Islands, Vol. 1. Canadian Museum of Nature, Ottawa.Google Scholar
Tappan, M. J. 1994a. Savanna ecology and natural bone deposition: implications for early hominid site formation, hunting and scavenging. Current Anthropology 36:223260.Google Scholar
Tappan, M. J. 1994b. Bone weathering in the tropical rain forest. Journal of Archaeological Science 21:667673.Google Scholar
Tardy, Y., N'Kounkou, R., and Probst, J.-L. 1989. The global water cycle and continental erosion during Phanerozoic time. American Journal of Science 289:455483.Google Scholar
Taylor, P. D. 1990. Preservation of soft-bodied and other organisms by bioimmuration: a review. Palaeontology 33:117.Google Scholar
Taylor, P. D. 1994. Evolutionary paleoecology of symbioses between bryozoans and hermit crabs. Historical Biology 9:157205.Google Scholar
Thayer, C. W. 1983. Sediment-mediated biological disturbance and the evolution of marine benthos. Pp. 479625 in Tevesz, M. J. S. and McCall, P. L., eds. Biotic interactions in recent and fossil benthic communities. Plenum, New York.Google Scholar
Thomasson, J. R. 1991. Sediment-borne “seeds” from Sand Creek, northwestern Kansas: taphonomic significance and paleoecological and paleoenvironmental implications. Palaeogeography, Palaeoclimatology, Palaeoecology 85:213225.Google Scholar
Traverse, A. 1990. Studies of pollen and spores in rivers and other bodies of water, in terms of source-vegetation and sedimentation, with special reference to Trinity River and Bay, Texas. Review of Palaeobotany Palynology 64:297303.Google Scholar
Traverse, A., ed. 1994. Sedimentation of organic particles. Cambridge University Press, Cambridge.Google Scholar
Trueman, C. N. 1999. Rare earth element geochemistry and taphonomy of terrestrial vertebrate assemblages. Palaios 14:555568.Google Scholar
Trueman, C. N., and Benton, M. J. 1997. A geochemical method to trace the taphonomic history of reworked bones in sedimentary settings. Geology 25:263266.Google Scholar
Underwood, C. J., and Bottrell, S. M. 1994. Diagenetic controls on multiphase pyritization of graptolites. Geological Magazine 131:315327.Google Scholar
Valentine, J. W. 1989. How good was the fossil record? Clues from the Californian Pleistocene. Paleobiology 15:8394.Google Scholar
Valentine, J. W., Jablonski, D., and Erwin, D. H. 1999. Fossils, molecules, and embryos: new perspectives on the Cambrian explosion. Development 126:851859.Google Scholar
Vaughan, A., and Nichols, G. 1995. Controls on the deposition of charcoal: implications for sedimentary accumulations of fusain. Journal of Sedimentary Research A 65:129135.Google Scholar
Vermeij, G. J. 1977. The Mesozoic marine revolution: evidence from snails, predators and grazers. Paleobiology 2:245258.Google Scholar
Vermeij, G. J. 1987. Evolution and escalation: an ecological history of life. Princeton University Press, Princeton, N.J. Google Scholar
Voorhies, M. R. 1992. Ashfall: life and death at a Nebraska waterhole ten million years ago. University of Nebraska State Museum, Museum Notes 81.Google Scholar
Vrba, E. S. 1995. On the connections between paleoclimate and evolution. Pp. 2445 in Vrba, E. S., Denton, G. H., Partridge, T. C., and Burckle, L. H., eds. Paleoclimate and evolution, with emphasis on human origins. Yale University Press, New Haven, Conn.Google Scholar
Wagner, P. J. 2000a. The quality of the fossil record and the accuracy of phylogenetic inferences about sampling and diversity. Systematic Biology 49:6586.Google Scholar
Wagner, P. J. 2000b. Phylogenetic analyses and the fossil record: tests and inferences, hypotheses and models. In Erwin, D. H. and Wing, S. L., eds. Deep time: Paleobiology's perspective. Paleobiology 26(Suppl. to No. 4):341371.Google Scholar
Waisfeld, B. G., Sanchez, T. M., and Carrera, M. G. 1999. Biodiversification patterns in the Early Ordovician of Argentina. Palaios 14:198214.Google Scholar
Walker, K. R., and Bambach, R. K. 1971. The significance of fossil assemblages from fine-grained sediments: time-averaged communities. Geological Society of America Abstracts with Programs 3:783784.Google Scholar
Walker, K. R., and Diehl, W. W. 1985. The role of marine cementation in the preservation of lower Paleozoic assemblages. Philosophical Transactions of the Royal Society of London B 311:143153.Google Scholar
Walker, S. E. 1988. Taphonomic significance of hermit crabs (Anomura: Paguridea): epifaunal hermit crab—infaunal gastropod example. Palaeogeography, Palaeoclimatology, Palaeoecology 63:4571.Google Scholar
Walker, S. E. 1989. Hermit crabs as taphonomic agents. Palaios 4:439452.Google Scholar
Walker, S. E. 1995. Taphonomy of modern and fossil intertidal gastropod associations from Isla Santa Cruz and Isla Santa Fe, Galapagos Islands. Lethaia 28:371382.Google Scholar
Walker, S. E., and Voight, J. R. 1994. Paleoecologic and taphonomic potential of deepsea gastropods. Palaios 9:4858.Google Scholar
Walker, S. E., Cappellen, P. Van, Roychoudhury, A., and Koretsky, C. 1997. Preservation of experimentally deployed molluscan carbonate below the sediment-water interface. Geological Society of America Abstracts with Programs 30:266.Google Scholar
Walker, S. E., Parsons-Hubbard, K., Powell, E. N., and Brett, C. E. 1998. Bioerosion or bioaccumulation? Shelf-slope trends for epi- and endobionts on experimentally deployed gastropod shells. Historical Biology 13:6172.Google Scholar
Walter, M. R., Veevers, J. J., Calver, C. R., and Grey, K. 1995. Neoproterozoic stratigraphy of the Centralian Superbasin, Australia. Precambrian Research 73:173195.Google Scholar
Walters, L. J., and Wethey, D. S. 1991. Settlement, refuges, and adult body form in colonial marine invertebrates: a field experiment. Biological Bulletin 180:112118.Google Scholar
Warren, R. E. 1991. Ozarkian fresh-water mussels (Unionoidea) in the upper Eleven Point River, Missouri. American Malacological Bulletin 8:131137.Google Scholar
Wayne, R. K., Leonard, J. A., and Cooper, A. 1999. Full of sound and fury: the recent history of ancient DNA. Annual Review of Ecology and Systematics 30:457477.Google Scholar
Webb, T. III. 1993. Constructing the past from late-Quaternary pollen data: temporal resolution and a zoom lens space-time perspective. In Kidwell and Behrensmeyer 1993.Google Scholar
Wehmiller, J. F., York, L. L., and Bart, M. L. 1995. Amino acid racemization geochronology of reworked Quaternary mollusks on US Atlantic coast beaches: implications for chronostratigraphy, taphonomy, and coastal sediment transport. Marine Geology 124:303337.Google Scholar
Westall, F., Boni, L., and Guerzoni, E. 1995. The experimental silicification of microorganisms. Palaeontology 38:495528.Google Scholar
Whittington, H. B., and Morris, S. Conway, eds. 1985. Extraordinary fossil biotas: their ecological and evolutionary significance. Philosophical Transactions of the Royal Society of London B 311:1192.Google Scholar
Wilby, P. R., and Briggs, D. E. G. 1997. Taxonomic trends in the resolution of detail preserved in fossil phosphatized soft tissues. Geobios Mémoire Spécial 20:493502.Google Scholar
Wilby, P. R., Briggs, D. E. G., Bernier, P., and Gaillard, C. 1996. Role of microbial mats in the fossilization of soft tissues. Geology 24:787790.Google Scholar
Wilf, P., and Labandeira, C. 1999. Response of plant-insect associations to Paleocene-Eocene warming. Science 284:21532156.Google Scholar
Wilf, P., Beard, K. C., Davies-Vollum, K. S., and Norejko, J. W. 1998. Portrait of a Late Paleocene (Early Clarkforkian) terrestrial ecosystem: Big Multi Quarry and associated strata, Washakie Basin, southwestern Wyoming. Palaios 13:514532.Google Scholar
Williams, M. E. 1994. Catastrophic versus noncatastrophic extinction of the dinosaurs: testing, falsifiability, and the burden of proof. Journal of Paleontology 68:183190.Google Scholar
Wilson, M. V. H. 1988a. Taphonomic processes: information loss and information gain. Geoscience Canada 15:131148.Google Scholar
Wilson, M. V. H. 1988b. Reconstruction of ancient lake environments using both autochthonous and allochthonous fossils. Palaeogeography, Palaeoclimatology, Palaeoecology 62:609623.Google Scholar
Wilson, M. V. H. 1988c. Predation as a source of fish fossils in Eocene lake sediments. Palaios 2:497504.Google Scholar
Wilson, M. V. H. 1993. Calibration of Eocene varves at Horsefly, British Columbia, Canada, and temporal distribution of specimens of the Eocene fish Amyzon aggregatum Wilson. Kaupia (Darmstadter Beitrage zur Naturgeschichte) 2:2738.Google Scholar
Wilson, M. V. H., and Barton, D. G. 1996. Seven centuries of taphonomic variation in Eocene freshwater fishes preserved in varves: paleoenvironments and temporal averaging. Paleobiology 22:535542.Google Scholar
Wing, S. L., and DiMichele, W. A. 1995. Conflict between local and global changes in plant diversity through geologic time. Palaios 10:551564.Google Scholar
Wing, S. L., Hickey, L. J., and Swisher, C. C. 1993. Implications of an exceptional fossil flora for Late Cretaceous vegetation. Nature 363:342344.Google Scholar
Wolff, R. G. 1975. Sampling and sample size in ecological analyses of fossil mammals. Paleobiology 1:195204.Google Scholar
Xiao, S., and Knoll, A. H. 1999. Fossil preservation in the Neoproterozoic Doushantuo phosphorite Lagerstätte, South China. Lethaia 32:219240.Google Scholar
Yang, H., and Yang, S. 1994. The Shanwang fossil biota in eastern China: a Miocene Konservat-Lagerstätte in lacustrine deposits. Lethaia 27:345354.Google Scholar
Zuschin, M., and Hohenegger, J. 1998. Subtropical coral-reef associated sedimentary facies characterized by molluscs (northern Bay of Safaga, Red Sea, Egypt). Facies 38:229254.Google Scholar
Zuschin, M., Hohenegger, J., and Steininger, F. F. 2000. A comparison of living and dead molluscs on coral reef associated hard substrata in the northern Red Sea—implications for the fossil record. Palaeogeography, Palaeoclimatology, Palaeoecology 159:167190.Google Scholar