Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-25T17:01:27.380Z Has data issue: false hasContentIssue false

Quantitative estimates of time-averaging in terebratulid brachiopod shell accumulations from a modern tropical shelf

Published online by Cambridge University Press:  08 April 2016

Monica Carroll
Affiliation:
Department of Geological Sciences, Virginia Polytechnic Institute and State University, Blacksburg, Virginia 24061
Michał Kowalewski
Affiliation:
Department of Geological Sciences, Virginia Polytechnic Institute and State University, Blacksburg, Virginia 24061. E-mail: michalk@vt.edu
Marcello G. Simões
Affiliation:
Instituto de Biociências, Universidade Estadual Paulista, Distrito de Rubião Junior, CP. 510, 18.610-000, Botucatu, SP, Brazil. E-mail: btsimoes@ibb.unesp.br
Glenn A. Goodfriend
Affiliation:
Department of Earth and Environmental Sciences, George Washington University, Washington, D.C. 20052. E-mail: glenng@gwu.edu

Abstract

Quantitative estimates of time-averaging in marine shell accumulations available to date are limited primarily to aragonitic mollusk shells. We assessed time-averaging in Holocene assemblages of calcitic brachiopod shells by direct dating of individual specimens of the terebratulid brachiopod Bouchardia rosea. The data were collected from exceptional (brachiopod-rich) shell assemblages, occurring surficially on a tropical mixed carbonate-siliciclastic shelf (the Southeast Brazilian Bight, SW Atlantic), a setting that provides a good climatic and environmental analog for many Paleozoic brachiopod shell beds of North America and Europe. A total of 82 individual brachiopod shells, collected from four shallow (5–25 m) nearshore (<2.5 km from the shore) localities, were dated by using amino acid racemization (D-alloisoleucine/L-isoleucine value) calibrated with five AMS-radiocarbon dates (r2 = 0.933). This is the first study to demonstrate that amino acid racemization methods can provide accurate and precise ages for individual shells of calcitic brachiopods.

The dated shells vary in age from modern to 3000 years, with a standard deviation of 690 years. The age distribution is strongly right-skewed: the young shells dominate the dated specimens and older shells are increasingly less common. However, the four localities display significant differences in the range of time-averaging and the form of the age distribution. The dated shells vary notably in the quality of preservation, but there is no significant correlation between taphonomic condition and age, either for individual shells or at assemblage level.

These results demonstrate that fossil brachiopods may show considerable time-averaging, but the scale and nature of that mixing may vary greatly among sites. Moreover, taphonomic condition is not a reliable indicator of pre-burial history of individual brachiopod shells or the scale of temporal mixing within the entire assemblage. The results obtained for brachiopods are strikingly similar to results previously documented for mollusks and suggest that differences in mineralogy and shell microstructure are unlikely to be the primary factors controlling the nature and scale of time-averaging. Environmental factors and local fluctuations in populations of shell-producing organisms are more likely to be the principal determinants of time-averaging in marine benthic shelly assemblages. The long-term survival of brachiopod shells is incongruent with the rapid shell destruction observed in taphonomic experiments. The results support the taphonomic model that shells remain protected below (but perhaps near) the surface through their early taphonomic history. They may be brought back up to the surface intermittently by bioturbation and physical reworking, but only for short periods of time. This model explains the striking similarities in time-averaging among different types of organisms and the lack of correlation between time-since-death and shell taphonomy.

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

Abreu, J. 1980. Distribuição e ecologia dos Decapoda numa área estuarina de Ubatuba, SP. Boletim do Instituto Oceanográfico 29:13.Google Scholar
Allison, P. A., and Briggs, D. E. G., eds. 1991. Taphonomy: releasing the data locked in the fossil record. Topics in Geobiology, Vol. 9, Plenum, New York.Google Scholar
Anderson, L. C., Gupta, B. K., 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
Angulo, R. J., Giannini, P. C. F., Suguio, K., and Pessenda, L. C. R. 1999. Relative sea-level changes in the last 5500 years in southern Brazil (Laguna-Imbituba region, Santa Catarina State) based on vermetid 14C ages. Marine Geology 159:323339.Google Scholar
Baker, R. G. V., Haworth, R. J., and Flood, P. G. 2001. Warmer or cooler late Holocene marine palaeoenvironments? Interpreting Southeast Australian and Brazilian sea-level changes using fixed biological indicators and their δ18O composition. Palaeogeography, Palaeoclimatology, Palaeoecology 168:249272.Google Scholar
Be, A. W. H., and Hutson, W. H. 1977. Ecology of planktonic foraminifera and biogeographic pattern of life and fossil assemblages in the Indian Ocean. Micropaleontology 23:369414.Google Scholar
Behrensmeyer, A. K., and Kidwell, S. M. 1985. Taphonomy's contributions to paleobiology. Paleobiology 11:105119.Google Scholar
Behrensmeyer, A. K., Kidwell, S. M., and Gastaldo, R. A. 2000. Taphonomy and paleobiology. In Erwin, D. H. and Wing, S. L., eds. Dee time: Paleobiology‘s perspective. Paleobiology 26(Suppl. to No. 4):103147.Google Scholar
Benigni, C. 1987. Shell microstructure of Mediterranean terebratulids from Pliocene to Recent and its diagnostic significance. Bolletino del Museo Regionale de Scienza Naturale de Torino 15:211217.Google Scholar
Bowman, S. 1990. Radiocarbon dating. University of California Press, Berkeley.Google Scholar
Brandt, D. S. 1989. Taphonomic grades as a classification for fossiliferous assemblages and implications for paleoecology. Palaios 4:303309.Google Scholar
Brett, C. E., and Baird, G. C. 1993. Taphonomic approaches to temporal resolution in stratigraphy: examples from Paleozoic marine mudrocks. Pp. 250274in Kidwell, and Behrensmeyer, 1993.Google Scholar
Brunton, C. H. C. 1996. The functional morphology of the Recent brachiopod Bouchardia rosea. Acta Zoologica 77:233240.Google Scholar
Campos, E. J. D., Gonçalves, J. E., and Ikeda, Y. 1995. Water mass characteristics and geostrophic circulation in the South Brazil Bight: summer of 1991. Journal of Geophysical Research 100:1853718550.Google Scholar
Campos, E. J. D., Velhot, D., and da Silveira, I. C. A. 2000. Shelf break upwelling driven by Brazil Current cyclonic meanders. Geophysical Research Letters 27:751754.Google Scholar
Carter, R. M., Abbott, S. T., Fulthorpe, T., Craig, S., Haywick, D. W., and Henderson, R. A. 1991. Application of global sea-level and sequence-stratigraphic models in Southern Hemisphere Neogene strata from New Zealand. In MacDonald, D. I. M., ed. Sedimentation, tectonics and eustasy: sea-level changes at active margins. Special Publication of the International Association of Sedimentologists 12:4165.Google Scholar
Clifton, H. E. 1971. Orientation of empty pelecypod shells and shell fragments in quiet water. Journal of Sedimentary Petrology 41:671682.Google Scholar
Collins, M. J. 1986. Post mortality strength loss in shells of the Recent articulate brachiopod Terebratulina retusa (L.) from the west coast of Scotland. In Racheboeuf, P. R. and Emig, C. C., eds. Les brachiopodes fossiles et actuels. Biostratigraphie du Paleozoique 4:209218.Google Scholar
Daley, G. M. 1993. Passive deterioration of shelly material: a study of the recent Eastern Pacific articulate brachiopod Terebratulia transversa Sowerby. Palaios 8:226232.Google Scholar
Davies, D. J., Powell, E. N., and Stanton, R. J. Jr. 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 72:317352.Google Scholar
Driscoll, E. G. 1970. Selective bivalve shell destruction in marine environments: a field study. Journal of Sedimentary Petrology 40:898905.Google Scholar
Emig, C. C. 1990. Examples of postmortem alteration in recent brachiopod shells and (paleo)ecological consequences. Marine Biology 104:233238.Google Scholar
Flessa, K. W. 1993. Time-averaging and temporal resolution in Recent marine shelly faunas. Pp. 933in Kidwell, and Behrensmeyer, 1993.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
Flessa, K. W., Hallman, D. P., Goodfriend, G. A., and Kowalewski, M. 1996. Why is the taphonomic clock such a poor timekeeper? In Repetski, J. E., ed. Sixth North American Paleontological Convention, Abstracts, of papers. Paleonotological Society Special Publication 8:121.Google Scholar
Forneris, L. 1969. Faunas bentônicas da Baía do Flamengo, Ubatuba, SP: aspectos ecológicos. Ph.D. dissertation. Universidade de São Paulo, Sao Paulo.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 Flessa, K. W. 1987. Taphonomy of tidal flat molluscs in the northern Gulf of California: paleoenvironmental analysis despite the perils of preservation. Palaios 2:543559.Google Scholar
Fürsich, F. T., and Flessa, K. W. 1991. The origin and interpretation of Bahía la Choya (northern Gulf of California) taphocoenoses: implications for paleoenvironmental analysis. Zitteliana 18:165169.Google Scholar
Gaspard, D. 1989. Quelques aspects de la biodegradation des coquilles de brachiopodes; consequences sur leur fossilisation. Bulletin de la Société Géologique de France, 8e série 5:12071216.Google Scholar
Gaspard, D. 1996. Taphonomy of some Cretaceous and Recent Brachiopods. Pp. 95102in Copper, P. and Jin, J., eds. Brachiopods. Proceedings of the Third International Brachiopod Congress, Sudbury, Canada.Google Scholar
Goodfriend, G. A. 1987. Chronostratigraphic studies of sediments in the Negev Desert, using amino acid epimerization analysis of land snail shells. Quaternary Research 28:374392.Google Scholar
Goodfriend, G. A. 1989. Complementary use of amino acid epimerization and radiocarbon analysis for dating mixed-age fossil assemblages. Radiocarbon 31:10411047.Google Scholar
Goodfriend, G. A. 1991. Patterns of racemization and epimerization of amino acids in land snail shells over the course of the Holocene. Geochimica et Cosmochimica Acta 55:293302.Google Scholar
Goodfriend, G. A. 1992. Rapid racemization of aspartic acid in mollusk shells and potential for dating over recent centuries. Nature 357:399401.Google Scholar
Goodfriend, G. A., and Stanley, D. J. 1996. Reworking and discontinuities in Holocene sedimentation in the Nile Delta: documentation from amino acid racemization and stable isotopes in mollusk shells. Marine Geology 129:271283.Google Scholar
Goodfriend, G. A., Brigham-Grette, J., and Miller, G. H. 1996. Enhanced age resolution of the marine Quaternary record in the Arctic using aspartic acid racemization dating of bivalve shells. Quaternary Research 45:176187.Google Scholar
Goodfriend, G. A., Flessa, K. W., and Hare, P. H. 1997. Variation in amino acid epimerization rates and amino acid composition among shell layers in the bivalve Chione from the Gulf of California. Geochimica et Cosmochimica Acta 61:14871493.Google Scholar
Hare, P. E., and Mitterer, R. M. 1969. Laboratory simulations of amino-acid diagenesis in fossils. Carnegie Institution of Washington Yearbook 67:205208.Google Scholar
Hearty, P. J. 1987. New data on the Pleistocene of Mallorca. Quaternary Science Reviews 6:245257.Google Scholar
Hearty, P. J., Miller, G. H., Stearns, C. E., and Szabo, B. J. 1986. Aminostratigraphy of Quaternary shorelines in the Mediterranean basin. Geological Society of America Bulletin 97:850858.Google Scholar
Kidwell, S. M. 1990. Phanerozoic evolution of macroinvertebrate shell accumulations: preliminary data from the Jurassic of Great Britain. In Miller, W. M. III, ed. Paleocommunity temporal dynamics. Paleontological Society Special Publication 5:309327. Knoxville, Tenn.Google Scholar
Kidwell, S. M. 1991. The stratigraphy of shell concentrations. Pp. 211290in Allison, and Briggs, 1991.Google Scholar
Kidwell, S. M. 1993. Patterns of time-averaging in the shallow marine fossil record. Pp. 275300in Kidwell, and Behrensmeyer, 1993.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., and Behrensmeyer, A. K., eds. 1993. Taphonomic approaches to time resolution in the fossil assemblages. Short Courses in Paleontology No. 6. Paleontological Society, Knoxville, Tenn.Google Scholar
Kidwell, S. M., and Best, M. M. R. 2001. Tropical time-averaging: disparate absolute ages and taphonomic clocks in bivalve assemblages from modern subtidal siliciclastic and carbonate facies. Seventh North American Paleontological Convention, Abstracts. PaleoBios 21(Suppl. to No. 2):79.Google Scholar
Kidwell, S. M., and Bosence, D. W. J. 1991. Taphonomy and time-averaging of marine shelly faunas. Pp. 115209in Allison, P. A., and Briggs, D. E. G., eds. Taphonomy: releasing data locked in the fossil record. Topics in Geobiology 9, Plenum Press, New York.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. 209336in 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., Rothfus, T. A., and Best, M. M. R. 2001. Sensitivity of taphonomic signatures to sample size, sieve size, damage scoring system, and target taxa. Palaios 16:2652.Google Scholar
Kowalewski, M. 1996. 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 Bambach, R. K. 2003. The limits of paleontological resolution. In Harris, P. J., and Geary, D. H., eds. High resolution approaches in paleontology. Topics in Geobiology Series. Kluwer Academic/Plenum, New York.Google Scholar
Kowalewski, M., Flessa, K. W., and Aggen, J. A. 1994. Taphofacies analysis of Recent shelly cheniers (beach ridges), northeastern Baja California, Mexico. Facies 31:209242.Google Scholar
Kowalewski, M., Goodfriend, G. A., and Flessa, K. W. 1998. The high-resolution estimates of temporal mixing in shell beds: the evils and virtues of time-averaging. Paleobiology 24:287304.Google Scholar
Kowalewski, M., Avila Serrano, G. E., Flessa, K. W., and Goodfriend, G. A. 2000. A dead delta's former productivity: Two trillion shells at the mouth of the Colorado River. Geology 28:10591062.Google Scholar
Kowalewski, M., Simões, M. G., Carroll, M., and Rodland, D. L. 2002. Abundant articulated brachiopods on a tropical, upwelling-influenced shelf (Southeast Brazilian Bight, South Atlantic). Palaios 17:277286.Google Scholar
Lessa, G. C., Angulo, R. J., Giannini, P. C., and Araujo, A. D. 2000. Stratigraphy and Holocene evolution of a regressive barrier in South Brazil. Marine Geology 165:87108.Google Scholar
Magliocca, A., and Kurtner, A. S. 1964. Conteúdo orgânico dos sedimentos de fundo de Cananéia, São Paulo. Contribuições do Instituto Oceanográfico 195:265273.Google Scholar
Magliocca, A., and Kurtner, A. S. 1965. Sedimentos de fundo da Enseada do Flamengo, Ubatuba. Contribuições do Instituto Oceanográfico 198:115.Google Scholar
Mahiques, M. M. de. 1995. Sedimentary dynamics of the bays off Ubatuba, state of Sao Paulo. Boletim do Instituto Oceanográfico, São Paulo 43:111122.Google Scholar
Mahiques, M. M. de, Tessler, M. G., and Furtado, V. V. 1998. Characterization of energy gradient in enclosed bays of Ubatuba region, Southeastern Brazil. Estuarine, Coastal and Shelf Science 47:431446.Google Scholar
Mañcenido, M., and Griffin, M. 1988. Distribution and palaeoenvironmental significance of the genus Bouchardia (Brachiopoda, Terebratellidina): its bearing on the Cenozoic evolution of the South Atlantic. Revista Brasileira de Geociências 18(2):201211.Google Scholar
Mantelatto, F. L. M., and Fransozo, A. 1999. Characterization of the physical and chemical parameters of Ubatuba Bay, northern coast of Sao Paulo State, Brazil. Revista Brasileira de Biologia 59:2331.Google Scholar
Martin, L., Suguio, K., Flexor, J. M., Dominguez, J. M. L., and Bittencourt, C. S. P. 1996. Quaternary sea-level history and variation in dynamics along the central Brazilian coast: consequences on coastal plain construction. Anais da Academia Brasileira de Ciências 68:303354.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 Liddel, W. D. 1996. Comparative taphonomy of bivalves and foraminifera from Holocene tidal flat sediments, Bahia la Choya, Sonora, Mexico (northern Gulf of California): taphonomic grades and temporal resolution. Paleobiology 22:8090.Google Scholar
Meldahl, K. H., Flessa, K. W., and Cutler, A. H. 1997. Time-averaging and postmortem skeletal survival in benthic fossil assemblages: quantitative comparisons among Holocene environments. Paleobiology 23:207229.Google Scholar
Miller, B. B., and Hare, P. E. 1980. Amino acid geochronology: integrity of the carbonate matrix and potential of molluscan fossils. Pp. 415443in Hare, P. E. et al., eds. Biogeochemistry of amino acids. Wiley, New York.Google Scholar
Miller, B. B., McCoy, W. D., and Bleuer, N. K. 1987. Stratigraphic potential of amino acid ratios in Pleistocene terrestrial gastropods: an example from West-Central Indiana. Boreas 16:133138.Google Scholar
Muir-Wood, H. M., Stehli, F. G., Elliott, G. F., and Hatai, K. 1965. Terebratulida. Pp. H728H857in Williams, A. et al. Brachiopoda 2. Part H of Moore, R. C., ed. Treatise on invertebrate paleontology. Geological Society of America, New York, and University of Kansas, Lawrence.Google Scholar
Negreiros-Fransozo, M. L., Fransozo, A., Pinheiro, M. A. A., Mantelatto, F. L. M., and Santos, S. 1991. Caracterização física e química da enseada de Fortaleza, Ubatuba, SP. Revista Brasileira de Geociências 21:114120.Google Scholar
Olszewski, T. 1999. Taking advantage of time-averaging. Paleobiology 25:226238.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
Peterson, C. H. 1977. The paleontological significance of undetected short-term temporal variability. Journal of Paleontology 51:976981.Google Scholar
Powell, E. N., and Davies, D. J. 1990. When is an ‘old’ shell really old? Journal of Geology 98:823844.Google Scholar
Richardson, J. R. 1981. Brachiopods and pedicles. Paleobiology 7:8795.Google Scholar
Rudwick, M. J. S. 1970. Living and fossil brachiopods. Hutchison, London.Google Scholar
Sejrup, H. P., and Haugen, J.-E. 1994. Amino acid diagenesis in the marine bivalve Arctica islandica Linné from northwest European sites: only time and temperature? Journal of Quaternary Science 9:301309.Google Scholar
Simões, M. G., Kowalewski, M., Mello, L. H., Rodland, D. L., and Carroll, M. 2000. Present-day terebratulid brachiopods from the southern Brazilian shelf: paleontological and biogeographic implications. Geological Society of America Abstracts with Programs 32:A14.Google Scholar
Simões, M. G., Kowalewski, M., Mello, L. H., Rodland, D. L., and Carroll, M. 2001. Taphonomy of Bouchardia rosea (Brachiopoda, Rhynchonelliformea) from sandy beaches and bays of the northwest coast of Sao Paulo State. Reunião Anual da Sociedade Brasileira de Paleontologia, Paleo-2001, Rio Claro, SP, Boletim de Resumos, p. 15.Google Scholar
Stuiver, M., and Reimer, P. J. 1993. Extended C-14 data-base and revised CALIB 3.0 C-14 age calibration program. Radiocarbon 35:215230.Google Scholar
Stuiver, M., Reimer, P. J., and Braziunas, T. F. 1998. High-precision radiocarbon age calibration for terrestrial and marine samples. Radiocarbon 40:11271151.Google Scholar
Tommasi, L. R. 1967. Observações preliminares sobre a fauna bêntica de sedimentos moles da Baía de Santos e regiões vizinhas. Boletim do Instituto Oceanográfico 16:4365.Google Scholar
Tommasi, L. R. 1970a. Sobre o braquiópode Bouchardia rosea (Mawe, 1823). Boletim do Instituto Oceanográfico 19:3342.Google Scholar
Tommasi, L. R. 1970b. Observações sobre a fauna bêntica do complexo estuarino-lagunar de Cananéia, SP. Boletim do Instituto Oceanográfico 19:4356.Google Scholar
Torello, F. F., Simões, M. G., and Passos, J. R. 2000. Taphonomic tumbling: destroying shells to understand the nature of fossil record. Reunião Anual da Sociedade Brasileira de Paleontologia, Paleo-2000, Botucatu, Resumos, p. 23.Google Scholar
Torello, F. F., Simões, M. G., and Passos, J. R. 2001. The Taphonomic tumbling barrel: design, construction and applications. 17 Congresso Brasileiro de Paleontologia, Rio Branco, AC, Resumos, p. 32.Google Scholar
Torello, F. F., Simões, M. G., and Passos, J. R. 2002. The taphonomic tumbling barrel: a methodological review to understand preservational biases in the fossil record. First International Paleontological Congress, 2002, Sydney, Abstracts 68:285286.Google Scholar
Torres, T., García-Alonso, P., Canoira, L., Llamas, J., Coello, F. J., García-González, L., Nestares, T., Peláez, A., and Rodríguez-Alto, N. 1997. Racemizatión de los aminoácidos de braquiópodos y pelecípodos de la sección de Cuesta Colorada (Almería, SE de España). Geogaceta 21:207210.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, S. E., and Voight, J. R. 1994. Paleoecologic and taphonomic potential of deep sea gastropods. Palaios 9:4859.Google Scholar
Wehmiller, J. F. 1977. Amino acid studies of the Del Mer, California, midden site: apparent rate constants, ground temperature models, and chronological implications. Earth and Planetary Science Letters 37:184196.Google Scholar
Wehmiller, J. F. 1982. A review of amino acid racemization studies in Quaternary mollusks: stratigraphic and chronologic applications in coastal and interglacial sites, Pacific and Atlantic coasts, United States, United Kingdom, Baffin Island and tropical islands. Quaternary Science Reviews 1:83120.Google Scholar
Wehmiller, J. F., York, L. L., and Bart, M. L. 1995. Amino-acid racemization geochronology of reworked Quaternary mollusks on U.S. Atlantic coast beaches: implications for chronostratigraphy, taphonomy, and coastal sediment transport. Marine Geology 124:303337.Google Scholar
Williams, A., and Rowell, A. J. 1965. Morphology. Pp. H57H155in Williams, A. et al. Brachiopoda 1. Part H of Moore, R. C., ed. Treatise on invertebrate paleontology. Geological Society of America, New York, and University of Kansas, Lawrence.Google Scholar
Williams, A., James, M. A., Emig, C. C., MacKay, S., and Rhodes, M. C. 1997. Anatomy. Pp. 7188in Williams, A. et al. Brachiopoda 1. Part H (revised) of Kaesler, R. L., ed. Treatise on invertebrate paleontology. Geological Society of America, Boulder, Colo., and University of Kansas, Lawrence.Google Scholar
Wilson, M. V. H. 1988. Taphonomic processes: information loss and information gain. Geoscience Canada 15:131148.Google Scholar