Hostname: page-component-76fb5796d-r6qrq Total loading time: 0 Render date: 2024-04-26T04:14:23.475Z Has data issue: false hasContentIssue false

Quantitative comparisons and models of time-averaging in bivalve and brachiopod shell accumulations

Published online by Cambridge University Press:  08 April 2016

Richard A. Krause Jr.
Affiliation:
Department of Geology and Geophysics, Yale University, New Haven, Connecticut 06520. E-mail: richard.krause@yale.edu
Susan L. Barbour
Affiliation:
Department of Geosciences and Natural Resources Management, Western Carolina University, Cullowhee, North Carolina 28723
Michał Kowalewski
Affiliation:
Department of Geosciences, Virginia Tech, Blacksburg, Virginia 24061
Darrell S. Kaufman
Affiliation:
Department of Geology, Northern Arizona University, Flagstaff, Arizona 86011
Christopher S. Romanek
Affiliation:
Department of Earth & Environmental Sciences, University of Kentucky, Lexington, Kentucky 40506
Marcello G. Simões
Affiliation:
Instituto de Biociências, Universidade Estadual Paulista, Botucatu, São Paulo, Brazil
John F. Wehmiller
Affiliation:
Department of Geology, University of Delaware, Newark, Delaware 19716

Abstract

The variation in time-averaging between different types of marine skeletal accumulations within a depositional system is not well understood. Here we provide quantitative data on the magnitude of time-averaging and the age structure of the sub-fossil record of two species with divergent physical and ecological characteristics, the brachiopod Bouchardia rosea and the bivalve Semele casali. Material was collected from two sites on a mixed carbonate-siliciclastic shelf off the coast of Brazil where both species are dominant components of the local fauna.

Individual shells (n = 178) were dated using amino acid racemization (aspartic acid) calibrated with 24 AMS radiocarbon dates. Shell ages range from modern to 8118 years b.p. for brachiopods, and modern to 4437 years for bivalves. Significant differences in the shape and central tendency of age-frequency distributions are apparent between each sample. Such differences in time-averaging magnitude confirm the assumption that taphonomic processes are subject to stochastic variation at all spatial and temporal scales. Despite these differences, each sample is temporally incomplete at centennial resolution and three of the four samples have similar right-skewed age-frequency distributions. Simulations of temporal completeness indicate that samples of both species from the shallow site are consistent with a more strongly right-skewed and less-complete age-frequency distribution than those from the deep site.

We conclude that intrinsic characteristics of each species exert less control on the time-averaging signature of these samples than do extrinsic factors such as variation in rates of sedimentation and taphonomic destruction. This suggests that brachiopod-dominated and bivalve-dominated shell accumulations may be more similar in temporal resolution than previously thought, and that the temporal resolution of multi-taxic shell accumulations may depend more on site-to-site differences than on the intrinsic properties of the constituent organisms.

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

Allmon, W. D. 1989. Paleontological completeness of the record of lower Tertiary mollusks, U.S. Gulf and Atlantic Coastal Plains: implications for phylogenetic studies. Historical Biology 3:141158.CrossRefGoogle Scholar
Anderson, L. C., Gupta, B. K. Sen, McBride, R. A., and Byers, M. R. 1997. Reduced seasonality of Holocene climate and pervasive mixing of Holocene marine section: northeastern Gulf of Mexico shelf. Geology 25:127130.2.3.CO;2>CrossRefGoogle Scholar
Angulo, R. J., and Lessa, G. C. 1997. The Brazilian sea-level curves: a critical review with emphasis on the curves from the Paranaguá and Cananéia regions. Marine Geology 140:141166.CrossRefGoogle Scholar
Angulo, R. J., Giannini, P. C. F., Suguio, K., and Pessenda, L. C. R. 1999. Relative sea-level changes in the last 550 years in southern Brazil (Laguna-Imbituba Region, Santa Catarina State) based on vermetid 14C ages. Marine Geology 159:323339.CrossRefGoogle Scholar
Angulo, R. J., de Souza, M. C., Reimer, P. J., and Sasaoka, S. K. 2005. Reservoir effect of the Southern and Southeastern Brazilian Coast. Radiocarbon 47:6773.CrossRefGoogle Scholar
Angulo, R. J., Lessa, G. C., and de Souza, M. C. 2006. A critical review of mid- to late-Holocene sea-level fluctuations on the eastern Brazilian coastline. Quaternary Science Reviews 25:486586.Google Scholar
Ayers, G. P. 2001. Comment on regression analysis of air quality data. Atmospheric Environment 35:24232425.CrossRefGoogle Scholar
Wood, S. L. Barbour, Krause, R. A. Jr., Kowalewski, M., Wehmiller, J. F., and Simões, M. G. 2006. Aspartic acid racemization dating of Holocene brachiopods and bivalves from the Southern Brazilian Shelf, South Atlantic. Quaternary Research 66:323331.CrossRefGoogle Scholar
Behrensmeyer, A. K., Kidwell, S. M., and Gastaldo, R. A. 2000. Taphonomy and paleobiology. In Erwin, D. H. and Wing, S. L., eds. Deep time: Paleobiology's perspective Paleobiology 26(Suppl. to No. 4):103147.CrossRefGoogle Scholar
Behrensmeyer, A. K., Fürsich, F. T., Gastaldo, R. A., Kidwell, S. M., Kosnik, M. A., Kowalewski, M., Plotnick, R. E., Rogers, R. R., and Alroy, J. 2005. Are the most durable shelly taxa also the most common in the marine fossil record? Paleobiology 31:607623.CrossRefGoogle Scholar
Blackwell, B. A. B., Last, W. M., and Rutter, N. W. 2000. Biogeochemical diagenesis in recent mammalian bones from saline lakes in western Victoria, Australia. Pp. 88107 in Goodfriend, et al. 2000.Google Scholar
Boss, K. J. 1972. The genus Semele in the Western Atlantic. Johnsonia 5(49):132.Google Scholar
Braga, E. S., and Muller, T. J. 1998. Observation of regeneration of nitrate, phosphate and silicate during upwelling off Ubatuba, Brazil, 23 degrees south. Continental Shelf Research 18:915922.CrossRefGoogle Scholar
Brigham, J. K. 1983. Intrashell variations in amino acid concentrations and isoleucine epimerization ratios in fossil Hiatella arctica . Geology 11:509513.Google Scholar
Brunton, C. H. C. 1996. The functional morphology of the Recent brachiopod Bouchardia rosea . Acta Zoologica 77:233240.Google Scholar
Burnham, R. J. 1993. Time resolution in terrestrial macrofloras; guidelines from modern accumulations. Pp. 5778 in Kidwell, and Behrensmeyer, 1993.Google Scholar
Burone, L. 2002. Foraminíferos bentônicos e parâmetros físicoquímicos da enseada de Ubatuba, São Paulo: estudo ecológico em uma área com poluição orgânica. Ph.D. dissertation. University of São Paulo, Institute of Oceanography, São Paulo.Google Scholar
Burone, L., Muniz, P., Pires-Vanin, A. M. S., and Rodrigues, M. 2003. Spatial distribution of organic matter in the surface sediments of Ubatuba Bay (Southeastern Brazil). Anais da Academia Brasileira de Ciências 75:7790.Google Scholar
Bush, A. M., Powell, M. G., Arnold, W. S., Bert, T. M., and Daley, G. M. 2002. Time-averaging, evolution, and morphologic variation. Paleobiology 28:925.Google Scholar
Callender, W. R., Staff, G. M., Powell, E. N., and MacDonald, I. R. 1990. Gulf of Mexico hydrocarbon seep communities. V. Biofacies and shell orientation of autochthonous shell beds below storm wave base. Palaios 5:214.CrossRefGoogle Scholar
Callender, W. R., Powell, E. N., Staff, G. M., Davies, D. J., Beauchamp, B., and von Bitter, P. H. 1992. Distinguishing autochthony, parautochthony and allochthony using taphofacies analysis; can cold seep assemblages be discriminated from assemblages of the nearshore and continental shelf? Palaios 7:409421.CrossRefGoogle Scholar
Callender, W. R., Powell, E. N., and Staff, G. M. 1994. Taphonomic rates of molluscan shells placed in autochthonous assemblages on the Louisiana continental slope. Palaios 9:6073.CrossRefGoogle Scholar
Callender, W. R., Staff, G.M., Parsons-Hubbard, K. M., Powell, E. N., Rowe, G. T., Walker, S. E., Brett, C. E., Raymond, A., Carlson, D., White, S., and Heiss, E. A. 2002. Taphonomic trends along a forereef slope: Lee Stocking Island, Bahamas. I. Location and water depth. Palaios 17:5065.Google Scholar
Campos, E. J. D., Velhot, D., and da Silveira, I. C. A. 1995. Water mass characteristics and geostrophic circulation in the South Brazil Bight: summer of 1991. Journal of Geophysical Research 100:1853718550.CrossRefGoogle 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
Carroll, M., Kowalewski, M., Simões, M. G., and Goodfriend, G. A. 2003. Quantitative estimates of time-averaging in brachiopod shell accumulations from a modern tropical shelf. Paleobiology 29:381402.2.0.CO;2>CrossRefGoogle Scholar
Castelao, R. M., Campos, E. J. D., and Miller, J. L. 2004. A modeling study of coastal upwelling driven by wind and meanders of the Brazil Current. Journal of Coastal Research 20:662671.CrossRefGoogle Scholar
Cherns, L., and Wright, V. P. 2000. Missing mollusks as evidence of large-scale, early skeletal aragonite dissolution in a Silurian sea. Geology 28:791794.2.0.CO;2>CrossRefGoogle Scholar
Coan, E. V. 1988. Recent eastern Pacific species of the bivalve genus Semele . Veliger 31:142.Google Scholar
Cummins, H., Powell, E. N., Stanton, R. J. Jr., and Staff, G. 1986. The rate of taphonomic loss in modern benthic habitats: how much of the potentially preservable community is preserved? Palaeogeography Palaeoclimatology Palaeoecology 89:193204.Google Scholar
da Rocha, J., Milliman, J. D., Santana, C. I., and Vicalvi, M. A. 1975. Southern Brazil. In Milliman, J. D., and Summerhayes, C. P., eds. Upper continental margin sedimentation off Brazil. Contributions to Sedimentology No. 4:117150. Schweizerbart, Stuttgart.Google Scholar
Depetris, P. J., Kemp, S., Latif, M., and Mook, W. G. 1996. ENSO-controlled flooding in the Paraná River (1904–1991). Naturwissenschaften 83:127129.Google Scholar
Diaconis, P., and Efron, B. 1983. Computer intensive methods in statistics. Scientific American 248:116130.CrossRefGoogle Scholar
DiCiccio, T. J., and Romano, J. P. 1988. A review of bootstrap confidence intervals. Journal of the Royal Statistical Society B 50:338354.Google Scholar
Domaneschi, O. 1995. A comparative study of the functional morphology of Semele purpurascens (Gmelin, 1791) and Semele proficua . Veliger 38:323342.Google Scholar
Efron, B. 1981. Nonparametric standard errors and confidence intervals. Canadian Journal of Statistics 9:139172.CrossRefGoogle 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.CrossRefGoogle 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
Folk, R. L. 1959. Practical petrographic classification of limestones. American Association of Petroleum Geologists Bulletin 43:138.Google Scholar
Fürsich, F. T., and Aberhan, M. 1990. Significance of time averaging for paleocommunity analysis. Lethaia 23:143152.CrossRefGoogle 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.CrossRefGoogle 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. 1992. Rapid racemization of aspartic acid in mollusc shells and potential for dating over Recent centuries. Nature 357:399401.CrossRefGoogle 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.CrossRefGoogle 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.CrossRefGoogle Scholar
Goodfriend, G. A., Collins, M. J., Fogel, M. L., Macko, S. A., and Wehmiller, J. F., eds. 2000. Perspectives in amino acid and protein geochemistry. Oxford University Press, Oxford.Google Scholar
Greenwood, D. R. 1991. The taphonomy of plant macrofossils. Pp. 141169 in Donovan, S. K., ed. The processes of fossilization. Columbia University Press, New York.Google Scholar
Griffiths, D. 1998. Sampling effort, regression method, and the shape and slope of size-abundance relations. Journal of Animal Ecology 67:795804.CrossRefGoogle Scholar
Hall, P. 1992. Efficient bootstrap simulations. Pp. 127143 in Lepage, R. and Billard, L. L., eds. Exploring the limits of bootstrap. Wiley, New York.Google Scholar
Hearty, P. J., 1987. New data on the Pleistocene of Mallorca. Quaternary Science Reviews 6:245257.CrossRefGoogle Scholar
Hendry, J. P., Ditchfield, P. W., and Marshall, J. D. 1995. Two stage metamorphism of Jurassic aragonitic bivalves: implications for early diagenesis. Journal of Sedimentary Research A65:214224.Google Scholar
Hendry, J. P., Trewin, N. H., and Fallick, A. E. 1996. Low Mgcalcite marine cement in Cretaceous turbidites: origin, spatial distribution, and relationship to sea-water chemistry. Sedimentology 43:877900.Google Scholar
Hughen, K. A., Baillie, M. G. L., Bard, E., Bayliss, A., Bertrand, C. J. H., Blackwell, P. G., Buck, C. E., Burr, G. S., Cutler, K. B., Damon, P. E., Edwards, R. L., Fairbanks, R. G., Friedrich, M., Guilderson, T. P., Kromer, B., McCormac, F. G., Manning, S. W., Bronk Ramsey, C., Reimer, P. J., Reimer, R. W., Remmele, S., Southon, J. R., Stuiver, M., Talamo, S., Taylor, F. W., van der Plicht, J., and Weyhenmeyer, C. E. 2004. Marine04 marine radiocarbon age calibration, 26–0 ka BP. Radiocarbon 46:10591086.Google Scholar
Ikemoto, T., and Takai, K. 2000. A new linearized formula for the law of total effective temperature and the evaluation of line-fitting methods with both variables subject to error. Environmental Entomology 29:671682.CrossRefGoogle Scholar
Kaufman, D. S. 2000. Amino acid racemization in ostracodes. Pp. 145160 in Goodfriend, et al. 2000.Google Scholar
Kaufman, D. S., and Manley, W. F. 1998. A new procedure for determining DL amino acid ratios in fossils using reverse phase liquid chromatography. Quaternary Science Reviews 17:9871000.Google Scholar
Keil, R. G., Tsamakis, E., and Hedges, J. I. 2000. Early diagenesis of particulate amino acids in marine systems. Pp. 6982 in Goodfriend, et al. 2000.Google Scholar
Kidwell, S. M. 1993. Patterns of time averaging in the shallow marine fossil record. Pp. 275300 in 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.CrossRefGoogle Scholar
Kidwell, S. M., and Behrensmeyer, A. K., eds. 1993. Taphonomic approaches to time resolution in fossil assemblages. Short Courses in Paleontology 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. 116188 in Allison, P. A. and Briggs, D. E. G., eds. Taphonomy: releasing the data locked in the fossil record. Plenum Press, New York.Google Scholar
Kidwell, S. M., and Brenchley, P. J. 1994. Patterns in bioclastic accumulation through the Phanerozoic: Changes in input or in destruction? Geology 22:11391143.2.3.CO;2>CrossRefGoogle 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., Best, M. M. R., and Kaufman, D. S. 2005. Taphonomic trade-offs in tropical marine death assemblages: differential time averaging, shell loss, and probable bias in siliciclastic vs. carbonate facies. Geology 33:729732.Google Scholar
Kimber, R. W. L., and Griffin, C. V. 2000. Interpretation of D/L amino acid data from agricultural soils. Pp. 195201 in Goodfriend, et al. 2000.Google Scholar
Kosnik, M. A., and Kaufman, D. S. 2008. Identifying outliers and assessing the accuracy of amino acid racemization measurements for use in geochronology. II. Data screening. Quaternary Geochronology 3:328341.CrossRefGoogle Scholar
Kosnik, M. A., Hua, Q., Jacobsen, G. E., Kaufman, D. S., and Wüst, R. A. 2007. Sediment mixing and stratigraphic disorder revealed by the age-structure of Tellina shells in Great Barrier Reef sediment. Geology 35:811814.CrossRefGoogle Scholar
Kosnik, M. A., Kaufman, D. S., and Hua, Q. 2008. Identifying outliers and assessing the accuracy of amino acid racemization measurements for use in geochronology. I. Age calibration curves. Quaternary Geochronology 3:308327.Google Scholar
Kosnik, M. A., Hua, Q., Kaufman, D. S., and Wüst, R. A. 2009. Taphonomic bias and time-averaging in tropical molluscan death assemblages: differential shell half-lives in Great Barrier Reef sediment. Paleobiology 35:565586.CrossRefGoogle Scholar
Kowalewski, M. 1996a. Time-averaging, overcompleteness, and the geological record. Journal of Geology 104:317326.Google Scholar
Kowalewski, M. 1996b. Taphonomy of a living fossil: the lingulide brachiopod Glottidia palmeri Dall from Baja California, Mexico. Palaios 11:244265.Google Scholar
Kowalewski, M. 1997. The reciprocal taphonomic model. Lethaia 30:8688.CrossRefGoogle Scholar
Kowalewski, M., and Bambach, R. K. 2003. The limits of paleontological resolution. Pp. 148 in Harries, P. J. ed. Approaches in high-resolution stratigraphic paleontology. Kluwer Academic/Plenum Publishers, New York.Google Scholar
Kowalewski, M., and Rimstidt, J. D. 2003. Average lifetime and age spectra of detrital grains: toward a unifying theory of sedimentary particles. Journal of Geology 111:427439.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
Kowalewski, M., Serrano, G. E. A., Flessa, K. W., and Goodfriend, G. A. 2000. 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 brachiopods on a tropical upwelling-influenced shelf (Southeast Brazilian Bight, South Atlantic). Palaios 17:277286.Google Scholar
Kowalewski, M., Carroll, M., Casazza, L., Gupta, N., Hannisdal, B., Hendy, A., Krause, R. A. Jr., LaBarbera, M., Lazo, D. G., Messina, C., Puchalski, S., Rothfus, T. A., Sälgeback, J., Stempien, J., Terry, R. C., and Tomašových, A. 2003. Quantitative fidelity of brachiopod-mollusk assemblages from modern subtidal environments of San Juan Islands, USA. Journal of Taphonomy 1:4365.Google Scholar
Lajoie, K. R., Wehmiller, J. F., and Kennedy, G. L. 1980. Inter- and intrageneric trends in apparent racemization kinetics of amino acids in Quaternary mollusks. Pp. 305340 in Hare, P. E., Hoering, T. C., and King, K. Jr., eds. Biogeochemistry of amino acids. Wiley, New York.Google Scholar
Lee, D. E. 1991. Aspects of the ecology and distribution of the living Brachiopoda of New Zealand. Pp. 273279 in MacKinnon, D. L., Lee, D. L., and Campbell, J. D., eds. Brachiopods through time. Balkema, Rotterdam.Google Scholar
Lever, J., Kessler, A., Van Overbeeke, A. P., and Thijssen, R. 1961. Quantitative beach research. II. The “hole effect.” Netherlands Journal of Sea Research 1:339358.Google Scholar
Mahiques, M. M. d. 1995. Sedimentary dynamics of the bays off Ubatuba, state of Sao Paolo. Boletim do Institute Oceanografico, São Paulo 43:111122.CrossRefGoogle Scholar
Mahiques, M. M. d., 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
Mahiques, M. M., Tessler, M. G., Ciotti, A. M., da Silveira, I. C. A., Sousa, S. H. M., Figueria, R. C. L., Tassinari, C. C. G., Furtado, V. V., and Passos, R. F. 2004. Hydrodynamically-driven patterns of recent sedimentation in the shelf and upper slope off southeast Brazil. Continental Shelf Research 24:16851697.CrossRefGoogle Scholar
Manceñido, 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 Geosciências 18:201211.CrossRefGoogle Scholar
Manley, W. F., Miller, G. H., and Czywezynski, J. 2000. Kinetics of aspartic acid racemization in Mya and Hiatella: modeling age and paleotemperature of high-latitude Quaternary mollusks. Pp. 202218 in Goodfriend, et al. 2000.Google Scholar
Manly, B. F. J. 1997. Randomization and Monte Carlo methods in biology, 2d ed. Chapman and Hall, London.Google Scholar
Mantelatto, F. L. M., and Fransozo, A. 1999. Characterization of the physical and chemical parameters of Ubatuba Bay, northern coast of São Paulo State, Brazil. Revista Brasileira de Biologia 59:2331.CrossRefGoogle Scholar
Martin, L., Bittencourt, A. C. S. P., Dominguez, J. M. L., Flexor, J., and Suguio, K. 1998. Oscillations or not oscillations, that is the question: Comment on Angulo, R.J., and Lessa, G.C. “The Brazilian sea-level curves: a critical review with emphasis on the curves from the Paranaguá and Cananéia regions” [Mar. Geol. 140, 141–166]. Marine Geology 150:179187.Google Scholar
Martin, L., Dominguez, J. M. L., and Bittencourt, A. C. S. P. 2003. Fluctuating Holocene sea levels in eastern and southeastern Brazil: evidence from multiple fossil and geometric indicators. Journal of Coastal Research 19:101124.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, Bahía la Choya, Sonora, Mexico (northern Gulf of California): taphonomic grades and temporal resolution. Paleobiology 22:8090.Google Scholar
McCormac, F. G., Hogg, A. G., Blackwell, P. G., Buck, C. E., Higham, T. F. G., and Reimer, P. J. 2004. SHCa104 Southern hemisphere calibration 0–11.0 cal kyr BP. Radiocarbon 46:10871092.Google Scholar
McKittrick, M. A. 1987. Experiments on the settling of gastropod and bivalve shells: biostratinomic implications. In Flessa, K. W., ed. Paleoecology and taphonomy of Recent to Pleistocene intertidal deposits Gulf of California. Paleontological Society Special Publication 2:150163. Washington, D.C. CrossRefGoogle 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. 415443 in Hare, P. E., Hoering, T. C., and King, K. Jr., eds. Biogeochemistry of amino acids. Wiley, New York.Google Scholar
Mitterer, R. M., and Kriausakul, N. 1989. Calculation of amino acid racemization ages based on apparent parabolic kinetics. Quaternary Science Reviews 8:353357.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.CrossRefGoogle Scholar
Morse, J. W., Zullig, J. J., Bernstein, L. D., Millero, F. J., Milne, P., Mucci, A., and Choppin, G. R. 1985. Chemistry of calcium-rich shallow water sediments in the Bahamas. American Journal of Science 285:147185.Google Scholar
Narchi, W., and Domaneschi, O. 1977. Semele casali – Jurado, 1949 (Mollusca, Bivalvia) in the Brazilian littoral. Studies on Neotropical Fauna and Environment 12:263272.Google Scholar
Olszewski, T. D. 1999. Taking advantage of time-averaging. Paleobiology 25:226238.CrossRefGoogle Scholar
Olszewski, T. D. 2004. Modeling the influence of taphonomic destruction, reworking, and burial, on time-averaging in fossil accumulations. Palaios 19:3950.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.CrossRefGoogle Scholar
Parsons-Hubbard, K. M., Powell, E. N., Staff, G. M., Callender, W. R., Brett, C. E., and Walker, S. E. 2001. The effect of burial on shell preservation and epibiont cover in Gulf of Mexico and Bahamas shelf and slope environments after two years: an experimental approach. In Aller, J. Y., Woodin, S. A., and Aller, R. C., eds. Organism-sediment interactions. Belle W. Baruch Library in Marine Science No. 21:297314. University of South Carolina Press, Columbia.Google Scholar
Pohlo, R. 1982. Evolution of the Tellinacea (Bivalvia). Journal of Molluscan Studies 48:245256.CrossRefGoogle Scholar
Powell, E. N., Parsons-Hubbard, K. M., Callender, W. R, Staff, G. M., Rowe, G. T., Brett, C. E., Walker, S. E., Raymond, A., Carlson, D. D., White, S., and Heise, E. A. 2002. Taphonomy on the continental shelf and slope: two-year trends—Gulf of Mexico and Bahamas. Palaeogeography, Palaeoclimatology, Palaeoecology 184:135.Google Scholar
Powell, E. N., Callender, W. R, Staff, G. M., Parsons-Hubbard, K. M., Brett, C. E., Walker, S. E., Raymond, A., and Ashton-Alcox, K. A. 2008. Molluscan shell condition after eight years on the sea floor — Taphonomy in the Gulf of Mexico and Bahamas. Journal of Shellfish Research 27:191225.CrossRefGoogle Scholar
Richardson, J. R. 1981. Brachiopods and pedicles. Paleobiology 7:8795.Google Scholar
Rodland, D. L., Kowalewski, M., Carroll, M., and Simões, M. G. 2004. Colonization of a “lost world” encrustation patterns in modern subtropical brachiopod assemblages. Palaios 19:381395.Google Scholar
Rodland, D. L., Kowalewski, M., Carroll, M., and Simões, M. G. 2006. The temporal resolution of epibiont assemblages; are they ecological snapshots or overexposures? Journal of Geology 114:313324.Google Scholar
Rodrigues, S. C. 2006. Tafonomia de moluscos bivalves e braquiópodes das enseadas de Ubatuba e Picinguaba, norte do Estado de São Paulo: implicações do uso de assinaturas tafonômicas no reconhecimento de gradientes ambientais. Ph.D. dissertation. University of São Paulo, Institute of Geosciences, São Paulo.Google Scholar
Rodrigues, S. C., Simóes, M. G., Kowalewski, M., Petti, M. A. V., Nonato, E. F., Martinez, S., and Del Rio, C. J. Biotic interaction between spionid polychaetes and bouchardiid brachiopods: Paleoecological, taphonomic, and evolutionary implications. Acta Palaeontologica Polonica 53:657668.Google Scholar
Sadler, P. M. 1981. Sediment accumulation and the completeness of stratigraphic sections. Journal of Geology 89:569584.CrossRefGoogle Scholar
Sheskin, D. J. 2004. Handbook of parametric and nonparametric statistical procedures, 3d ed. Chapman and Hall, London.Google Scholar
Simões, M. G., Kowalewski, M., Mello, L. H. C., Rodland, D. L., and Carroll, M. 2004. Recent brachiopods from the Southern Brazilian Shelf: palaeontological and biogeographical implications. Palaeontology 47:515533.Google Scholar
Simóes, M. G., Rodrigues, S. C., Leme, J. M., and Bissaro, M. C. Junior. 2005. The settling pattern of brachiopod shells: stratigraphic and taphonomic implications to shell bed formation and paleoecology. Revista Brasileira de Geosciências 35:383391.Google Scholar
Simões, M. G., Rodrigues, S.C., and Kowalewski, M. 2009. Bouchardia rosea, a vanishing brachiopod species of the Brazilian platform: taphonomy, historical ecology, and conservation paleobiology. Historical Biology 21:123137.CrossRefGoogle Scholar
Simões, M. G., Rodrigues, S. C., Neves, J. P., and Kowalewski, M. 2007. Experimental and field observations on the preservational bias of brachiopod valves in present-day assemblages of Bouchardia rosea from the Brazilian shelf: paleoecological implications. Geological Society of America Abstracts with Programs 39(2):96.Google Scholar
Soares-Gomes, A., and Pires-Vanin, A. M. S. 2003. Padrões de abundância, riqueza e diversidade de moluscos bivalves na plataforma continental ao largo de Ubatuba, São Paulo, Brasil: uma comparação metológica. Revista Brasileira de Zoologia, 20:717725.Google Scholar
Stead, R. A., Urrutia, G., Garrido, O., Clasing, E., Lardies, M. A., and Arratia, L. P. 2002. The significance of contrasting feeding strategies on the reproductive cycle in two coexisting tellinacean bivalves. Journal of the Marine Biological Association of the United Kingdom 82:443–53.Google Scholar
Stempien, J. A. 2005. Brachyuran taphonomy in a modern tidal-flat environment: preservation potential and anatomical bias. Palaios 20:400410.CrossRefGoogle Scholar
Stuiver, M., Reimer, P. J., and Reimer, R. W. 2005. CALIB 5.0.2, radiocarbon calibration program, http://calib.qub.ac.uk/calib/ Google Scholar
Teece, M. A., Tuross, N., Kress, W. M., Peterson, P. M., Russell, G., and Fogel, M. L. 2000. Preservation of amino acids in museum herbarium samples. Pp. 8387 in Goodfriend, et al. 2000.Google Scholar
Teixeira, C., and Tundisi, J. G. 1981. The effects of nitrogen and phosphorous enrichment on phytoplankton in the region of Ubatuba. Solatum do Instituto Oceanografico São Paulo 30:7786.Google Scholar
Tomašových, A., and Rothfus, T. A. 2005. Differential taphonomy of modern brachiopods (San Juan Islands, Washington State): effect of intrinsic factors on damage and community-level abundance. Lethaia 38:271292.Google Scholar
Tomašových, A., Fürsich, F. T., and Olszewski, T. D. 2006a. Modeling shelliness and alteration in shell beds: variation in hardpart input and burial rates leads to opposing predictions. Paleobiology 32:278298.Google Scholar
Tomašových, A., Fürsich, F. T., and Wilmsen, M. 2006b. Preservation of autochthonous shell beds by positive feedback between increased hardpart-input rates and increased sedimentation rates. Journal of Geology 114:287312.CrossRefGoogle Scholar
Tommasi, L. R. 1970. Sobre o braquiópode Bouchardia rosea (Mawe, 1823). Boletim do Instituto Oceanográfico 19:3-3.CrossRefGoogle Scholar
Torello, F. F. 2004. Tafonomia experimental do fossil-vivo Bouchardia rosea (Brachiopoda, Terebratellidae) e suas aplicações em Paleontologia. Ph.D. dissertation. University of São Paulo, Institute of Geosciences, São Paulo.Google Scholar
Torello, F. F., and Simões, M. G. 2003. Comparative bivalve and brachiopod shell resistance to abrasion tested in a taphonomic tumbling barrel. Third Latin American Congress of Sedimentology, Belém 2003. Abstracts, pp. 223224.Google Scholar
Waite, E. R., and Collins, M. J. 2000. The interpretation of aspartic acid racemization of dentine proteins. Pp. 182194 in Goodfriend, et al. 2000.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(7):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. 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., and Miller, G. H. 2000. Aminostratigraphic dating methods in Quaternary geology. In Noller, J. S., Sowers, J. M., and Lettis, W. R., eds. Quaternary geochronology, methods and applications. American Geophysical Union Reference Shelf 4:187222.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. H57H155 in Williams, A. et al. Brachiopoda 1. Part H of Moore, R. C. ed. Treatise on invertebrate paleontology. Geological Society of America, Boulder, Colo., and University of Kansas, Lawrence.Google Scholar
Williams, A., James, M. A., Emig, C. C., MacKay, S., and Rhodes, M. C. 1997. Anatomy. Pp. 7188 in 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 (Paleoscene No. 9):131148.Google Scholar
Wing, S. L., and DiMichele, W. A. 1995. Conflict between local and global changes in plant diversity through geological time. Palaios 10:551564.Google Scholar
Wright, P., Cherns, L., and Hodges, P. 2003. Missing molluscs: field testing taphonomic loss in the Mesozoic through early large-scale aragonite dissolution. Geology 31:211214.Google Scholar
Yanes, Y., Kowalewski, M., Ortiz, J. E., Castillo, C., Torres, T., and Nuez, J. 2007. Scale and structure of time-averaging (age mixing) in terrestrial gastropod assemblages from Quaternary eolian deposits of the eastern Canary Islands. Palaeogeography, Palaeoclimatology, Palaeoecology 251:283299.Google Scholar
Supplementary material: File

Krause et al. supplementary material

Appendices

Download Krause et al. supplementary material(File)
File 330.8 KB