Hostname: page-component-848d4c4894-wg55d Total loading time: 0 Render date: 2024-06-02T23:05:25.278Z Has data issue: false hasContentIssue false

Time resolution in fluvial vertebrate assemblages

Published online by Cambridge University Press:  08 February 2016

Anna K. Behrensmeyer*
Affiliation:
Department of Paleobiology, National Museum of Natural History, Smithsonian Institution, Washington, D.C. 20560

Abstract

Calibrating levels of time resolution that are accessible in the fossil record is important in understanding what evolutionary phenomena can be profitably studied using fossils. A model for attritional bone assemblage formation in fluvial deposits, based on observations of taphonomic processes in modern environments, provides order-of-magnitude estimates for time intervals represented in single unit, ‘contemporaneous' vertebrate samples. In order to form units with adequate material for analysis of morphological variation or paleoecological associations, it appears that bones must be spatially concentrated or stratigraphically condensed by sedimentary processes or biological agencies. In many cases this means that significant periods of time will be represented by single unit assemblages. According to predictions from modern environments, carcasses contributed through normal attrition can accumulate in the soil to ‘fossiliferous' densities over time intervals of 102–104 yrs. Attritional channel assemblages include bones from three sources: floodplain land surfaces, floodplain deposits, and the active channel, and represent time intervals on the order of 102–104 yrs. Given additional limitations on the composition of the fossil sample imposed by circumstances of preservation, outcrop availability and collecting strategy, attritional fluvial assemblages probably can be resolved only to 103 years even under the best conditions. Time intervals represented by fossils are not necessarily the same as those represented by sedimentary events in fluvial systems because bones can continue to accumulate and may be concentrated during times of erosion or non-deposition. Fluvial vertebrate assemblages of comparable taphonomic history can be used to document evolutionary changes over periods longer than their finest level of time resolution. While they may not be applicable to questions of punctuated or gradual transitions over shorter time scales, the longer-term patterns should have their own evolutionary significance.

Type
Articles
Copyright
Copyright © The Paleontological Society 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Literature Cited

Allen, J. R. L. 1974. Studies in fluviatile sedimentation: implication of pedogenic carbonate units, lower Old Red Sandstone, Anglo-Welsh outcrop. Geol. J. 9:181208.CrossRefGoogle Scholar
Badgley, C. E. and Behrensmeyer, A. K. 1980. Paleoecology of middle Siwalik sediments and faunas, northern Pakistan. Palaeogeogr., Palaeoclimat., Palaeoecol. 30:133155.CrossRefGoogle 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. 1978. Taphonomic and ecologic information from bone weathering. Paleobiology. 4:150162.CrossRefGoogle Scholar
Behrensmeyer, A. K. 1982. Time sampling intervals in the vertebrate fossil record. Vol. 1. Proc. Third N. Am. Paleontol. Conv. (Montreal, Aug. 1982).Google Scholar
Behrensmeyer, A. K. and Dechant Boaz, D. E. 1980. The recent bones of Amboseli National Park, Kenya, in relation to East African paleoecology. Pp. 7293. In: Behrensmeyer, A. K. and Hill, A., eds. Fossils in the Making. 338 pp.Univ. Chicago Press; Chicago.Google Scholar
Behrensmeyer, A. K. and Tauxe, L. 1982. Isochronous fluvial systems in Miocene deposits of northern Pakistan. Sedimentology. 29:331335.CrossRefGoogle 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.CrossRefGoogle Scholar
Bookstein, F. L., Gingerich, P. D., and Kluge, A. G. 1978. Hierarchical linear modeling of the tempo and mode of evolution. Paleobiology. 4:120134.CrossRefGoogle Scholar
Bown, T. M. 1979. Geology and mammalian paleontology of the Sand Creek facies, lower Willwood Formation (lower Eocene), Washakie County, Wyoming. Geol. Surv. Wyoming Mem. 2:1151.Google Scholar
Bown, T. M. and Kraus, M. J. 1981a. Lower Eocene alluvial paleosols (Willwood Formation, northwest Wyoming, U.S.A.) and their significance for paleoecology, paleoclimatology, and basin analysis. Palaeogeogr., Palaeoclimatol., Palaeoecol. 34:130.CrossRefGoogle Scholar
Bown, T. M. and Kraus, M. J. 1981b. Vertebrate fossil-bearing paleosol units (Willwood Formation, lower Eocene, northwest Wyoming, U.S.A.): implications for taphonomy, biostratigraphy and assemblage analysis. Palaeogeogr., Palaeoclimatol., Palaeoecol. 34:3156.CrossRefGoogle Scholar
Brain, C. K. 1980. Some criteria for the recognition of bone-collecting agencies in African caves. Pp. 108130. In: Behrensmeyer, A. K. and Hill, A., eds. Fossils in the Making. 338 pp.Univ. Chicago Press; Chicago.Google Scholar
Bridge, J. S. and Leeder, M. R. 1979. A simulation model of alluvial stratigraphy. Sedimentology. 26:617644.CrossRefGoogle Scholar
Clark, J., Beerbower, J. R., and Kietzke, K. K. 1967. Oligocene sedimentation, stratigraphy, paleoecology and paleoclimatology. Fieldiana: Geol. Mem. 5:1158.Google Scholar
Cornwall, I. W. 1958. Soils for the Archeologist. MacMillan; New York. 230 pp.Google Scholar
Dodson, P. 1971. Sedimentology and taphonomy of the Oldman Formation (Campanian) Dinosaur Provincial Park, Alberta (Canada). Palaeogeogr., Palaeoclimatol., Palaeoecol. 10:2174.CrossRefGoogle Scholar
Dodson, P., Behrensmeyer, A. K., Bakker, R. T., and McIntosh, J. S. 1980. Taphonomy and paleoecology of the dinosaur beds of the Jurassic Morrison Formation. Paleobiology. 6:208232.CrossRefGoogle Scholar
Dodson, P. and Wexlar, D. 1979. Taphonomic investigations of owl pellets. Paleobiology. 5:275284.CrossRefGoogle Scholar
Frey, R. W. and Basan, P. B. 1981. Taphonomy of relict Holocene salt marsh deposits, Cabretta Island, Georgia. Senckenbergiana Marit. 13:111155.Google Scholar
Fürsich, F. T. 1978. The influence of faunal condensation and mixing on the preservation of fossil benthic communities. Lethaia. 11:243250.CrossRefGoogle Scholar
Gifford, D. P. 1977. Observations of modern human settlements as an aid to archeological interpretation. Ph.D. dissertation, Dept. Anthropol., Univ. California, Berkeley. 434 pp.Google Scholar
Gifford, D. P. 1981. Taphonomy and paleoecology: a critical review of archaeology's sister disciplines. Pp. 365438. In: Schiffer, M. B., ed. Advances in Archaeological Method and Theory, Vol. 4. Academic Press; New York.CrossRefGoogle Scholar
Gingerich, P. D. 1974. Stratigraphic record of early Eocene Hyopsodus and the geometry of mammalian phylogeny. Nature. 248:107109.CrossRefGoogle Scholar
Gingerich, P. D. 1976. Paleontology and phylogeny: patterns of evolution at the species level in early Tertiary mammals. Am. J. Sci. 276:128.CrossRefGoogle Scholar
Gingerich, P. D. 1982. Time resolution in mammalian evolution: sampling, lineages, and faunal turnover. Vol. 1. Proc. Third N. Am. Paleontol. Conv. (Montreal, Aug. 1982).Google Scholar
Gordon, C. C. and Baikstra, J. E. 1981. Soil pH, bone preservation, and sampling bias at mortuary sites. Am. Antiq. 46:566571.CrossRefGoogle Scholar
Gould, S. J. and Eldridge, N. 1977. Punctuated equilibria: the tempo and mode of evolution reconsidered. Paleobiology. 3:115151.CrossRefGoogle Scholar
Gradzinski, R. 1969. Results of the Polish-Mongolian palaeontological expeditions, Pt. II: sedimentation of the dinosaur-bearing upper cretaceous deposits of the Nemegt Basin, Gobi Desert. Palaeontologia Polonica. 21:147229.Google Scholar
Hanson, C. B. 1980. Fluvial taphonomic processes: models and experiments. Pp. 156181. In: Behrensmeyer, A. K. and Hill, A., eds. Fossils in the Making. 338 pp.Univ. Chicago Press; Chicago.Google Scholar
Heinzelin, J. de. 1972. Omo Research Expedition 1967–1971. Africa-Tervuren. 19:6774.Google Scholar
Heinzelin, J. de, Haesaerts, P., and Howell, F. C. 1976. Plio-Pleistocene formations of the lower Omo Basin, with particular reference to the Shungura Formation. Pp. 2449. In: Coppens, Y., Howell, F. C., Isaac, G. L. I., and Leakey, R. E., eds. Earliest Man and Environments in the Lake Rudolf Basin. 615 pp.Univ. Chicago Press; Chicago.Google Scholar
Hill, A. P. 1975. Taphonomy of contemporary and late Cenozoic East African vertebrates. Ph.D. Dissertation, Univ. London. 331 pp.Google Scholar
Hunt, R. M. 1978. Depositional settings of a Miocene mammal assemblage, Sioux County, Nebraska (U.S.A.). Palaeogeogr., Palaeoclimatol., Palaeoecol. 24:152.CrossRefGoogle Scholar
Johnson, G. D. 1977. Paleopedology of Ramapithecus-bearing sediments, North India. Geol. Rundsch. 66:192216.CrossRefGoogle Scholar
Johnson, R. G. 1965. Pelecypod death assemblages in Tomales Bay, California. J. Paleontol. 39:8085.Google Scholar
Klein, R. G. 1981. Ungulate mortality and sedimentary facies in the late Tertiary Varwater Formation, Langebaanweg, South Africa. Ann. South African Mus. 84:233254.Google Scholar
Korth, W. W. 1979. Taphonomy of microvertebrate fossil assemblages. Ann. Carnegie Museum Nat. Hist. 48:235285.CrossRefGoogle Scholar
Langbein, W. B. and Leopold, L. B. 1968. River channel bars and dunes—theory of kinematic waves. U.S. Geol. Surv. Prof. Paper. 422-L:120.Google Scholar
Leeder, M. R. 1973. Fluviatile fining-upward cycles and the magnitude of paleochannels. Geol. Mag. 110:265276.CrossRefGoogle Scholar
Leeder, M. R. 1975. Pedogenic carbonates and flood sediment accretion rates: a quantitative model for arid zone lithofacies. Geol. Mag. 112:257270.CrossRefGoogle Scholar
Lisle, T. E. 1976. Components of flow resistance in an alluvial channel. Ph.D. Dissertation, Univ. Calif.; Berkeley. 60 pp.Google Scholar
Mellet, J. S. 1974. Scatological origins of microvertebrate fossil accumulations. Science 185:349350.CrossRefGoogle Scholar
Millar, C. E., Turk, L. M., and Foth, H. D. 1968. Fundamentals of soil science. 491 pp. John Wiley and Sons; New York.Google Scholar
Peterson, C. H. 1977. The paleoecological significance of undetected short-term variability. J. Paleontol. 51:976981.Google Scholar
Raup, D. M., and Crick, R. E. 1981. Evolution of single characters in the Jurassic ammonite Kosmoceras. Paleobiology. 7:200215.CrossRefGoogle Scholar
Sadler, P. H. 1981. Sediment accumulation rates and the completeness of stratigraphic sections. J. Geol. 89:569584.CrossRefGoogle Scholar
Sadler, P. M. and Dingus, L. W. 1982. Expected completeness of sedimentary sections: estimating a time-scale dependent, limiting factor in the resolution of the fossil record. Vol. 2. Proc. Third N. Am. Paleontol. Conv. (Montreal, Aug. 1982).Google Scholar
Schindel, D. E. 1980. Microstratigraphic sampling and the limits of paleontologic resolution. Paleobiology. 6:408426.CrossRefGoogle Scholar
Schindel, D. E. 1982. Time resolution in cyclic Pennsylvanian strata: implications for evolutionary patterns in Glabrocingulum (Mollusca: Archaeogastropoda). Vol. 2. Proc. Third N. Am. Paleontol. Conv. (Montreal, Aug. 1982).Google Scholar
Schopf, T. M. 1981. Punctuated equilibrium and evolutionary stasis. Paleobiology. 7:156166.CrossRefGoogle Scholar
Shipman, P. 1981. Life History of a Fossil. 222 pp. Harvard Univ. Press; Cambridge, Mass.Google Scholar
Shotwell, J. A. 1958. Inter-community relationships in Hemphillian (mid-Pliocene) mammals. Ecology. 39:271282.CrossRefGoogle Scholar
Stanley, S. M. 1979. Macroevolution. 332 pp. W. H. Freeman; San Francisco.Google Scholar
Van Ent, D. W. and Hesse, B. C. 1981. The chemical content of archaeological bone: notes toward the measurement of differential attrition in faunal samples. Proc. Soc. Am.Arch. Annual Meeting, 1981. Pp. 111.Google Scholar
Voorhies, M. R. 1969. Taphonomy and population dynamics of an early Pliocene vertebrate fauna, Knox County, Nebraska. Univ. Wyo. Contrib. Geol. Sp. Pap. 1. 69 pp.Google Scholar
Voorhies, M. R. 1970. Sampling difficulties in reconstructing late Tertiary mammalian communities. Proc. N. Am. Paleontol. Conv. Evolution of Communities, Pt. E. Pp. 454468.Google Scholar
Walker, K. R. and Bambach, R. K. 1971. The significance of fossil assemblages from fine-grained sediments: time averaged communities. Abstracts with programs, Geol. Soc. Am. Annual meeting, Washington, D.C. pp. 783784.Google Scholar
Warme, J. E., Ekdale, A. A., Ekdale, S. F., and Peterson, C. H. 1976. Raw material of the fossil record. Pp. 143169. In: Scott, R. W. and West, R. R., eds. Structure and Classification of Paleocommunities. 291 pp.Dowden, Hutchinson, Ross; Stroudsburg, Pa.Google Scholar
Western, D. 1979. Size, life history and ecology in mammals. Afr. J. Ecol. 17:185204.CrossRefGoogle Scholar
Western, D. 1980. Linking the ecology of past and present mammal communities. Pp. 4156. In: Behrensmeyer, A. K. and Hill, A., eds. Fossils in the Making. 338 pp.Univ. Chicago Press; Chicago.Google Scholar
Wood, W. R. and Johnson, D. L. 1978. A survey of disturbance processes in archeological site formation. Pp. 315381. In: Schiffer, M. B., ed. Advances in Archeological Method and Theory. Vol. 1. 426 pp.Academic Press; New York.CrossRefGoogle Scholar