Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-25T21:01:04.946Z Has data issue: false hasContentIssue false

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

Published online by Cambridge University Press:  08 April 2016

Ronald E. Martin
Affiliation:
Department of Geology, University of Delaware, Newark, Delaware 19716
John F. Wehmiller
Affiliation:
Department of Geology, University of Delaware, Newark, Delaware 19716
M. Scott Harris
Affiliation:
Department of Geology, University of Delaware, Newark, Delaware 19716
W. David Liddell
Affiliation:
Department of Geology, Utah State University, Logan, Utah 84322

Abstract

We compare the preservation (taphonomic grade) and age of Chione (bivalve) and foraminifera from modern siliciclastic tidal flat sediments of Bahia la Choya, Sonora, Mexico (northern Gulf of California). Disarticulated shells of Chione collected from the sediment-water interface of Choya Bay exhibit a substantial range in taphonomic grade and age, several hundred years to ∼80–125 ka based on Accelerator Mass Spectrometer 14C dates and D-Alloisoleucine/L-Isoleucine values. There is not, however, a one-to-one correspondence between age and taphonomic alteration of Chione: old (or young) valves may be highly altered or they may be relatively pristine. In contrast to Chione, most foraminiferal tests at Choya Bay are quite pristine, which suggests a quite young age, but tests are surprisingly old (up to ∼2,000 calendar years based on Accelerator Mass Spectrometer 14C dates).

We suggest that following seasonal pulses in reproduction, some foraminiferal tests are rapidly incorporated into a subsurface shell layer by “Conveyor Belt” deposit feeders and preserved there, while the rest of the reproductive pulse rapidly dissolves. Ultimately, some of these buried tests, along with Chione, are transported back to the surface by biological activity and storms. The much greater range of taphonomic grades and ages among Chione shells suggests that they, unlike foraminifera, are sufficiently large and preservable (low surface/volume ratio and chemical reactivity) to undergo many cycles of degradation, burial, and exhumation before complete destruction. The age of foraminiferal tests indicates that time-averaging of microfossil assemblages at Choya Bay is much more insidious than would be expected considering the relatively pristine state of the tests alone.

Based on our studies, the lower limit of temporal resolution of shallow shelf microfossil assemblages appears to be ∼1000 years. We caution, however, that each depositional setting (taphofacies) should be evaluated on a case-by-case basis before gross generalizations are made. Indeed, the discrepancy between age and taphonomic grade of fossil assemblages at Choya Bay suggests that neither hardpart size or taphonomic grade are infallible indicators of test preservability or likely temporal resolution of the host assemblage, and that the dynamics of hardpart input and loss must also be evaluated.

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

Aberhan, M., and Fürsich, F. T. 1987. Paleoecology and paleoenvironments of the Pleistocene deposits of Bahia la Choya (Gulf of California, Sonora, Mexico). In Fürsich, F. T., and Flessa, K. W., eds. Ecology, taphonomy, and paleoecology of Recent and Pleistocene molluscan faunas of Bahia la Choya, northern Gulf of California. Zitteliana 18:135163.Google Scholar
Aller, R. C. 1982. Carbonate dissolution in nearshore terrigenous muds: the role of physical and biological reworking. Journal of Geology 90:7995.CrossRefGoogle Scholar
Atkinson, K. 1969. The association of living foraminifera with algae from the littoral zone, South Cardigan Bay, Wales. Journal of Natural History 3:517542.CrossRefGoogle Scholar
Ben-Yaakov, S. 1973. pH buffering of pore water of recent anoxic marine sediments. Limnology and Oceanography 18:8694.Google Scholar
Berner, R. A., Scott, M. R., and Thomlinson, C. 1970. Carbonate alkalinity in the pore waters of anoxic marine sediments. Limnology and Oceanography 15:544549.Google Scholar
Berger, W. H., and Heath, G. R. 1968. Vertical mixing in pelagic sediments. Journal of Marine Research 26:134143.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
Brasier, M. D., and Green, O. R. 1993. Winners and losers: stable isotopes and microhabitats of living Archaiadae and Eocene Nummulites (larger foraminifera). Marine Micropaleontology 20:267276.Google Scholar
Bremer, M. L., and Lohmann, G. P. 1982. Evidence for primary control of the distribution of certain Atlantic Ocean benthic foraminifera by degree of carbonate saturation. Deep-Sea Research 29:987998.CrossRefGoogle Scholar
Brett, C. E., and Baird, G. C. 1986. Comparative taphonomy: a key to paleoenvironmental interpretation based on fossil preservation. Palaios 3:207227.Google Scholar
Buchardt, B., and Hansen, H. J. 1977. Oxygen isotope fractionation and algal symbiosis in benthic foraminifera from the Gulf of Elat, Israel. Bulletin of the Geological Society of Denmark 26:185194.Google Scholar
Canfield, D. E., and Raiswell, D. R. 1991. Carbonate precipitation and dissolution: its relevance to fossil preservation. pp. 411453In Allison, P. A. and Briggs, D. E. G., eds. Taphonomy: releasing the data locked in the fossil record. Plenum, New York.Google Scholar
Collins, M. J. 1985. Post mortality strength loss in shells of the Recent articulate brachiopod Terebratulina retusa (L.) from the west coast of Scotland. Biostratigraphie du Paleozoique 4:209218.Google Scholar
Corliss, B. H., and Honjo, S. 1981. Dissolution of deep-sea benthonic foraminifera. Micropaleontology 27:356378.Google Scholar
Cutler, A. H. 1987. Surface textures of shells as taphonomic indicators. In Flessa, K. W., ed. Paleoecology and taphonomy of Recent to Pleistocene intertidal deposits, Gulf of California. Paleontological Society Special Publication No. 2:164176.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
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
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.CrossRefGoogle Scholar
Driscoll, E. G. 1975. Sediment-animal-water interaction, Buzzards Bay, Massachusetts. Journal of Marine Research 33:275302.Google Scholar
Dubois, L. G., and Prell, W. L. 1988. Effects of carbonate dissolution on the radiocarbon age structure of sediment mixed layers. Deep-Sea Research 35:18751885.Google Scholar
Feige, A., and Fürsich, F. F. T. 1991. Taphonomy of the Recent molluscs of Bahia la Choya (Gulf of California, Sonora, Mexico). In Fürsich, F. T. and Flessa, K. W., eds. Ecology, taphonomy, and paleoecology of Recent and Pleistocene molluscan faunas of Bahia la Choya, northern Gulf of California. Zitteliana 18:89133.Google Scholar
Flessa, K. W. 1990. The “facts” of mass extinctions. In Sharpton, V. L. and Ward, P. D., eds. Global catastrophes in earth history: an interdiosciplinary conference on impacts, volcanism, and mass mortality. Geological Society of America Special Paper 247:17. Boulder, Colo.Google Scholar
Flessa, K. W. 1993. Time-averaging and temporal resolution in Recent marine shelly faunas. pp. 933in Kidwell and Behrensmeyer 1993b.Google Scholar
Flessa, K. W., and Brown, T. J. 1983. Selective solution of macroinvertebrate calcareous hard parts: a laboratory study. Lethaia 16:193205.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.CrossRefGoogle 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., eds. 1991. Ecology, taphonomy, and paleoecology of Recent and Pleistocene molluscan faunas of Bahia la Choya, northern Gulf of California. Zitteliana 18:1180.Google Scholar
Holland, S. M. 1995. The stratigraphic distribution of fossils. Paleobiology 21:92109.Google Scholar
Kidwell, S. M. 1986. Models for fossil concentrations: paleobiologic implications. Paleobiology 12:624.Google Scholar
Kidwell, S. M. 1989. Stratigraphic condensation of marine transgressive records: origin of major shell deposits in the Miocene of Maryland. Journal of Geology 97:124.Google Scholar
Kidwell, S. M. 1991. The stratigraphy of shell concentrations. pp. 211290In Allison, P. A. and Briggs, D. E. G., eds. Taphonomy: releasing the data locked in the fossil record. Plenum, New York.Google Scholar
Kidwell, S. M. 1993a. Patterns of time-averaging in the shallow marine fossil record. pp. 275300in Kidwell and Behrensmeyer 1993b.Google Scholar
Kidwell, S. M. 1993b. Taphonomic expressions of sedimentary hiatuses: field observations on bioclastic concentrations and sequence anatomy in low, moderate, and high subsidence settings. Geologische Rundschau 82:189202.Google Scholar
Kidwell, S. M., and Behrensmeyer, A. K. 1993a. Summary: estimates of time-averaging. pp. 301302in Kidwell and Behrensmeyer 1993b.Google Scholar
Kidwell, S. M., and Behrensmeyer, A. K., eds. 1993b. Taphonomic approaches to time resolution in fossil assemblages: Paleontological Society Short Courses in Paleontology No. 6. University of Tennessee, Knoxville.CrossRefGoogle Scholar
Kotler, E., Martin, R. E., and Liddell, W. D. 1991. Abrasion-resistance of modern reef-dwelling foraminifera from Discovery Bay, Jamaica—implications for test preservation. pp. 125138in Bain, R., ed. Proceedings of the Fifth Symposium on the Geology of the Bahamas.Google Scholar
Kotler, 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
Loubere, P. 1989. Bioturbation and sedimentation rate control of benthic microfossil taxon abundances in surface sediments: a theoretical approach to the analysis of species microhabitats. Marine Micropaleontology 14:317325.CrossRefGoogle Scholar
Loubere, P., and Gary, A. 1990. Taphonomic process and species microhabitats in the living to fossil assemblage transition of deeper water benthic foraminifera. Palaios 5:375381.Google Scholar
Loubere, P., Gary, A., and Lagoe, M. 1993. Generation of the benthic foraminiferal assemblage: theory and preliminary data. Marine Micropaleontology 20:165181.CrossRefGoogle Scholar
MacLeod, N., and Keller, G. 1994. Comparative biogeographic analysis of planktic foraminiferal survivorship across the Cretaceous/Tertiary (K/T) boundary. Paleobiology 20:143177.Google Scholar
Maluf, L. Y. 1983. Physical Oceanography. pp. 2645In Case, T. J. and Cody, M. L., eds. Island biogeography in the Sea of Cortez. University of California Press, Berkeley.Google Scholar
Marshall, C. R. 1990. Confidence intervals on stratigraphic ranges. Paleobiology 16:110.Google Scholar
Martin, R. E. 1993. Time and taphonomy: actualistic evidence for time-averaging of benthic foraminiferal assemblages. pp. 3456in Kidwell and Behrensmeyer 1993b.Google Scholar
Martin, R. E. and Liddell, W. D. 1991. Taphonomy of foraminifera in modern carbonate environments: implications for the formation of foraminiferal assemblages. pp. 170194in Donovan, S. K., ed. Fossilization: the processes of taphonomy. Belhaven, London.Google Scholar
Martin, R. E., Harris, M. S., and Liddell, W. D. 1996. Taphonomy and time-averaging of foraminiferal assemblages in Holocene tidal flat sediments, Bahia la Choya, Sonora, Mexico (northern Gulf of California). Marine Micropaleontology (in press).CrossRefGoogle Scholar
McCave, I. N. 1988. Biological pumping upwards of the coarse fraction of deep-sea sediments. Journal of Sedimentary Petrology 58:148158.Google Scholar
Meldahl, K. H. 1987. Sedimentologic and taphonomic implications of biogenic stratification. Palaios 2:350358.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
Murray-Wallace, C. V., and Belperio, A. P. 1994. Identification of remanié fossils using amino acid racemisation. Alcheringa 18:219227.Google Scholar
Myers, E. H. 1942. A quantitative study of the productivity of the foraminifera in the sea. Proceedings of the American Philosophical Society 85:325342.Google Scholar
Myers, E. H. 1943. Life activities of foraminifera in relation to marine ecology. Proceedings of the American Philosophical Society 86:439458.Google Scholar
Pospichal, J. J., Wise, S. W., Asaro, F., and Hamilton, N. 1990. The effects of bioturbation across a biostratigraphically complete high southern latitude Cretaceous/Tertiary boundary. In Sharpton, V. L. and Ward, P. D., eds. Global catastrophes in earth history: an interdiosciplinary conference on impacts, volcanism, and mass mortality. Geological Society of America Special Paper 247:497507. Boulder, Colo.Google Scholar
Powell, E. N., Cummins, H., Stanton, R. J., and Staff, G. 1984. Estimation of the size of molluscan larval settlement using the death assemblage. Estuarine and Coastal Shelf Science 18:367384.Google Scholar
Powell, E. N., Staff, G. M., Davies, D. J., and Russell Callender, W. 1989. Macrobenthic death assemblages in modern marine environments: formation, interpretation, and application. CRC Critical Reviews in Aquatic Sciences 1:555589.Google Scholar
Pride, C. J., Dean, W. E., and Thunell, R. C. 1994. Sedimentation in the Gulf of California: fluxes and accumulation rates of biogenic sediments and trace elements. Geological Society of America Abstracts with Programs 26:A23.Google Scholar
Reiss, Z., and Hottinger, L. 1984. The Gulf of Aqaba: ecological micropaleontology. Springer, Berlin.Google Scholar
Rhoads, D. C., and Stanley, D. J. 1965. Biogenic graded bedding. Journal of Sedimentary Petrology 35:956963.Google Scholar
Robinson, M. K. 1973. Atlas of monthly mean sea surface and subsurface temperatures in the Gulf of California. San Diego Society of Natural History Memoir 5.Google Scholar
Roden, G. I., 1964. Oceanographic aspects of Gulf of California. In van Andel, Tj., and Sjor, G. G., eds. Marine geology of the Gulf of California. American Association of Petroleum Geologists Memoir 3:3058. Tulsa, Okla.Google Scholar
Signor, P. W., and Lipps, J. H. 1982. Sampling bias, gradual extinction patterns and catastrophes in the fossil record. In Silver, L. T. and Schulz, P. H., eds. Geological implications of impacts of large asteroids and comets on the Earth. Geological Society of America Special Paper 190:291296. Boulder, Co.Google Scholar
Staff, G. M., Stanton, R. J., 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
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
Stuiver, M., and Braziunas, T. F. 1993. Modeling atmospheric 14C influences and 14C ages of marine samples to 10,000 b.c. Radiocarbon 35:137189.Google Scholar
Stuiver, M., and Polach, H. A. 1977. Discussion: reporting of 14C data. Radiocarbon 19:355363.Google Scholar
Stuiver, M., and Reimer, P. J. 1993. Extended 14C data base and revised CALIB 3.0 14C age calibration program. Radiocarbon 35:215230.Google Scholar
Stuiver, M., Pearson, G. W., and Braziunas, T. F. 1986. Radiocarbon age calibration of marine samples back to 9,000 Cal Yr b.c. Radiocarbon 28:9801021.Google Scholar
Sumpter, L. T. 1987. Grain size and provenance of Bahia la Choya sediments. In Flessa, K. W., ed. Paleoecology and taphonomy of Recent to Pleistocene intertidal deposits, Gulf of California. Paleontological Society Special Publication No. 2:4451.Google Scholar
van Straaten, L. M. J. U. 1952. Biogenic textures and the formation of shell beds in the Dutch Wadden Sea. Koninklijke Nederlandse Akademie van Wetenschappen, Proceedings. Series B, Physical Sciences 55:500516.Google Scholar
Walter, L. M., and Burton, E. A. 1990. Dissolution of recent platform carbonate sediments in marine pore fluids. American Journal of Science 290:601643.Google Scholar
Wehmiller, J. F. 1984a. Interlaboratory comparison of amino acid enantiomeric ratios in fossil Pleistocene mollusks. Quaternary Research 22:109120.Google Scholar
Wehmiller, J. F. 1984b. Relative and absolute dating of Quaternary mollusks with amino acid racemization: evaluation, applications, questions. pp. 171193in Mahaney, W. C., ed. Quaternary dating methods. Elsevier, Amsterdam.Google Scholar
Wehmiller, J. F. 1990. Amino acid racemization: applications in chemical taxonomy and chronostratigraphy of Quaternary fossils. pp. 583608in Carter, J. G., ed. Skeletal biomineralization: Patterns, processes, and evolutionary trends, Vol. 1. Van Nostrand Reinhold, New York.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