Hostname: page-component-8448b6f56d-wq2xx Total loading time: 0 Render date: 2024-04-23T23:09:07.592Z Has data issue: false hasContentIssue false

Isotaphonomy in concept and practice: an exploration of vertebrate microfossil bonebeds in the Upper Cretaceous (Campanian) Judith River Formation, north-central Montana

Published online by Cambridge University Press:  21 February 2017

Raymond R. Rogers
Affiliation:
Geology Department, Macalester College, 1600 Grand Avenue, Saint Paul, Minnesota 55105, U.S.A. E-mail: rogers@macalester.edu.
Matthew T. Carrano
Affiliation:
Department of Paleobiology, National Museum of Natural History, Smithsonian Institution, Post Office Box 37012, Washington, D.C. 20013, U.S.A. E-mail: carranom@si.edu
Kristina A. Curry Rogers
Affiliation:
Geology Department and Biology Department, Macalester College, 1600 Grand Avenue, Saint Paul, Minnesota 55105, U.S.A. E-mail: rogersk@macalester.edu
Magaly Perez
Affiliation:
Geology Department, Macalester College, 1600 Grand Avenue, Saint Paul, Minnesota 55105, U.S.A. E-mail: rogers@macalester.edu.
Anik K. Regan
Affiliation:
Geology Department, Macalester College, 1600 Grand Avenue, Saint Paul, Minnesota 55105, U.S.A. E-mail: rogers@macalester.edu.

Abstract

Vertebrate microfossil bonebeds (VMBs)—localized concentrations of small resilient vertebrate hard parts—are commonly studied to recover otherwise rarely found small-bodied taxa, and to document relative taxonomic abundance and species richness in ancient vertebrate communities. Analyses of taphonomic comparability among VMBs have often found significant differences in size and shape distributions, and thus considered them to be non-isotaphonomic. Such outcomes of “strict” statistical tests of isotaphonomy suggest discouraging limits on the potential for broad, comparative paleoecological reconstruction using VMBs. Yet it is not surprising that sensitive statistical tests highlight variations among VMB sites, especially given the general lack of clarity with regard to the definition of “strict” isotaphonomic comparability. We rigorously sampled and compared six VMB localities representing two distinct paleoenvironments (channel and pond/lake) of the Upper Cretaceous Judith River Formation to evaluate biases related to sampling strategies and depositional context. Few defining distinctions in bioclast size and shape are evident in surface collections, and most site-to-site comparisons of sieved collections are indistinguishable (p≤0.003). These results provide a strong case for taphonomic equivalence among the majority of Judith River VMBs, and bode well for future studies of paleoecology, particularly in relation to investigations of faunal membership and community structure in Late Cretaceous wetland ecosystems. The taphonomic comparability of pond/lake and channel-hosted VMBs in the Judith River Formation is also consistent with a formative model that contends that channel-hosted VMBs were reworked from pre-existing pond/lake assemblages, and thus share taphonomic history.

Type
Articles
Copyright
Copyright © 2017 The Paleontological Society. All rights reserved 

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

Aguirre, M. L., Richiano, S., Farinati, E., and Fucks, E.. 2011. Taphonomic comparison between two bivalves (Mactra and Brachidontes) from Late Quaternary deposits in northern Argentina: which intrinsic and extrinsic factors prevail under different palaeoenvironmental conditions? Quaternary International 233:113129.Google Scholar
Archibald, J. D. 1982. A study of Mammalia across the Cretaceous–Tertiary boundary in Garfield County, Montana. University of California Publications in Geological Sciences 122:1286.Google Scholar
Badgley, C. 1986. Taphonomy of mammalian fossil remains from Siwalik rocks of Pakistan. Paleobiology 12:119142.Google Scholar
Badgley, C., Downs, W., and Flynn, L. J.. 1998. Taphonomy of small-mammal assemblages from the middle Miocene Chinji Formation, Siwalik Group, Pakistan. Pp. 145166 in Y. Tomida, L. J. Flynn, and L. L. Jacobs, eds. Advances in vertebrate paleontology and geochronology. National Science Museum Monographs, Tokyo.Google Scholar
Behrensmeyer, A. K. 1988a. The pull of the recent analogue. Palaios 3:373.Google Scholar
Behrensmeyer, A. K. 1988b. Vertebrate preservation in fluvial channels. Palaeogeography, Palaeoclimatology, Palaeoecology 63:183199.CrossRefGoogle Scholar
Behrensmeyer, A. K. 1991. Terrestrial vertebrate accumulations. Pp. 291335. in P. Allison, and D. E. G. Briggs, eds. Taphonomy: releasing the data locked in the fossil record. Plenum, New York.Google Scholar
Behrensmeyer, A. K., and Barry, J.. 2005. Biostratigraphic surveys in the Siwaliks of Pakistan. A method for standardized surface sampling of the vertebrate fossil record. Palaeontologia Electronica 8(1):124.Google Scholar
Behrensmeyer, A. K., and Hook, R. W.. 1992. Paleoenvironmental contexts and taphonomic modes. Pp. 15136 in A. K. Behrensmeyer, J. D. Damuth, W. A. DiMichele, R. Potts, H.-D. Sues, and S. L. Wing, eds. Evolutionary paleoecology of terrestrial plants and animals. University of Chicago Press, Chicago.Google 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.Google Scholar
Behrensmeyer, A. K., Kidwell, S. M., and Gastaldo, R. A.. 2000. Taphonomy and paleobiology. Paleobiology 26:103144.Google Scholar
Blob, R. W., and Badgley, C.. 2007. Numerical methods for bonebed analysis. Pp. 333–396 in Rogers et al. 2007.CrossRefGoogle Scholar
Blob, R. W., and Fiorillo, A. R.. 1996. The significance of vertebrate microfossil size and shape distributions for faunal abundance reconstructions: a Late Cretaceous example. Paleobiology 22:422435.Google Scholar
Brinkman, D. B. 1990. Paleoecology of the Judith River Formation (Campanian) of Dinosaur Provincial Park, Alberta, Canada: evidence from vertebrate microfossil localities. Palaeogeography, Palaeoclimatology, Palaeoecology 78:3754.Google Scholar
Brinkman, D. B., Russell, A. P., Eberth, D. A., and Peng, J.. 2004. Vertebrate palaeocommunities of the lower Judith River Group (Campanian) of southeastern Alberta, Canada, as interpreted from vertebrate microfossil assemblages. Palaeogeography, Palaeoclimatology, Palaeoecology 213:295313.Google Scholar
Brinkman, D. B., Eberth, D. A., and Currie, P. J.. 2007. From bonebeds to paleobiology: applications of bonebed data. Pp. 221–264 in Rogers et al. 2007.Google Scholar
Bronstein, J. L. 2009. The evolution of facilitation and mutualism. Journal of Ecology 97:1601170.Google Scholar
Bryant, L. J. 1989. Non-dinosaurian lower vertebrates across the Cretaceous–Tertiary boundary in northeastern Montana. University of California Publications in Geological Sciences 134:1107.Google Scholar
Carrano, M. T., and Velez-Juarbe, J.. 2006. Paleoecology of the Quarry 9 vertebrate assemblage from Como Bluff, Wyoming (Morrison Formation, Late Jurassic). Palaeogeography, Palaeoclimatology, Palaeoecology 237:147159.Google Scholar
Carrano, M. T., Oreska, M. P. J., and Lockwood, R.. 2016. Vertebrate paleontology of the Cloverly Formation (Lower Cretaceous), II: paleoecology. Journal of Vertebrate Paleontology 26:112.Google Scholar
Chaney, D. S., and DiMichele, W. A.. 2003. Paleobotany of the classic redbeds (Clear Fork Group-Early Permian) of north central Texas. Pp. 357–356. in T. E. Wong, ed. Proceedings of the Fifteenth International Congress on Carboniferous and Permian Stratigraphy. Royal Netherlands Academy of Arts and Sciences, Utrecht, Netherlands.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
DeMar, D. G. Jr., and Breithaupt, B. H.. 2006. The nonmammalian vertebrate microfossil assemblages of the Mesaverde Formation (Upper Cretaceous, Campanian) of the Wind River and Bighorn Basins, Wyoming. Bulletin of the New Mexico Museum of Natural History and Science 35:3353.Google Scholar
Dodson, P. 1971. Sedimentology and taphonomy of the Oldman Formation (Campanian), Dinosaur Provincial Park, Alberta (Canada). Palaeogeography, Palaeoclimatology, Palaeoecology 10:2174.Google Scholar
Dodson, P. 1987. Microfaunal studies of dinosaur paleoecology, Judith River Formation of southern Alberta. Pp. 7075 in P. J. Currie, and E. Koster, eds. Fourth symposium on Mesozoic terrestrial ecosystems. Royal Tyrrell Museum Occasional Papers 3, Drumheller, Alberta.Google Scholar
Dunn, O. J. 1961. Multiple comparisons among means. Journal of the American Statistical Association 56:5264.Google Scholar
Eberth, D. A. 1990. Stratigraphy and sedimentology of vertebrate microfossil sites in the uppermost Judith River Formation (Campanian), Dinosaur Provincial Park, Alberta, Canada. Palaeogeography, Palaeoclimatology, Palaeoecology 78:136.Google Scholar
Eberth, D. A., and Hamblin, A. P.. 1993. Tectonic, stratigraphic, and sedimentologic significance of a regional discontinuity in the upper Judith River Group (Belly River wedge) of southern Alberta, Saskatchewan, and northern Montana. Canadian Journal of Earth Sciences 30:174200.Google Scholar
Eberth, D. A., Rogers, R. R., and Fiorillo, A. R.. 2007. A practical approach to the study of bonebeds. Pp. 265–331 in Rogers et al. 2007.Google Scholar
Efremov, I. A. 1940. Taphonomy: a new branch of paleontology. Pan-American Geologist 74:8193.Google Scholar
Estes, R. 1964. Fossil vertebrates from the Late Cretaceous Lance Formation, eastern Wyoming. University of California Publications in Geological Sciences 49:1180.Google Scholar
Estes, R. 1969. A new fossil discoglossid frog from Montana and Wyoming. Breviora 328:17.Google Scholar
Estes, R. 1976. Middle Paleocene lower vertebrates from the Tongue River Formation, southeastern Montana. Journal of Paleontology 50:500520.Google Scholar
Estes, R., and Berberian, P.. 1970. Paleoecology of a Late Cretaceous vertebrate community from Montana. Breviora 343:135.Google Scholar
Fagerstrom, J. A. 1964. Fossil communities in paleoecology: their recognition and significance. Geological Society of America Bulletin 75:11971216.Google Scholar
Fiorillo, A. R. 1989. The vertebrate fauna from the Judith River Formation (Late Cretaceous) of Wheatland and Golden Valley Counties, Montana. Mosasaur 4:127142.Google Scholar
Folk, R. L. 1966. Petrology of sedimentary rocks. Hemphill, Austin, Tex.Google Scholar
Foster, J. R., and Heckert, A. B.. 2011. Ichthyoliths and other microvertebrate remains from the Morrison Formation (Upper Jurassic) of northeastern Wyoming: a screen-washed sample indicates a significant aquatic component to the fauna. Palaeogeography, Palaeoclimatology, Palaeoecology 305:264279.Google Scholar
Frostick, L., and Reid, I.. 1983. Taphonomic significance of sub-aerial transport of vertebrate fossils on steep semi-arid slopes. Lethaia 16:157164.Google Scholar
Gastaldo, R. A., Adendorff, R., Bamford, M., LaBandeira, C. C., Neveling, J., and Sims, H.. 2005. Taphonomic trends of macrofloral assemblages across the Permian–Triassic boundary, Karoo Basin, South Africa. Palaios 20:479497.Google Scholar
Gill, J. R., and Cobban, W. A.. 1973. Stratigraphy and geologic history of the Montana Group and equivalent rocks, Montana, Wyoming, and North and South Dakota. U.S. Geological Survey Professional Paper 776:137.Google Scholar
Goodwin, M. B., and Deino, A. L.. 1989. The first radiometric ages from the Judith River Formation (Late Cretaceous), Hill County, Montana. Canadian Journal of Earth Sciences 26:13841391.Google Scholar
Hamblin, A. P., and Abrahamson, B. W.. 1996. Stratigraphic architecture of “Basal Belly River” cycles, Foremost Formation, Belly River Group, subsurface of southern Alberta and southwestern Saskatchewan. Bulletin of Canadian Petroleum Geology 44:654673.Google Scholar
Hayden, F. V. 1857. Notes explanatory of a map and section illustrating the geologic structure of the country bordering the Missouri River from the mouth of the Platte River to Fort Benton. Proceedings of the Academy of Natural Sciences of Philadelphia 9:109116.Google Scholar
Heckert, A. B. 2004. Late Triassic microvertebrates from the lower Chinle Group (Otischalkian–Adamanian: Carnian), southwestern U.S.A. New Mexico Museum of Natural History and Science Bulletin 27:1170.Google Scholar
Heckert, A. B., Mitchell, J. S., Schneider, V. P., and Olsen, P. E.. 2012. Diverse new microvertebrate assemblage from the Upper Triassic Cumnock Formation, Sanford Subbasin, North Carolina, USA. Journal of Paleontology 86:368390.Google Scholar
Hunt, A. P. 1991. Integrated vertebrate, invertebrate and plant taphonomy of the Fossil Forest Area (Fruitland and Kirtland formations—Late Cretaceous), San Juan County, New Mexico, USA. Palaeogeography, Palaeoclimatology, Palaeoecology 88:85107.Google Scholar
Hunter, A., and Donovan, S.. 2005. Field sampling bias, museum collections, and completeness of the fossil record. Lethaia 38:305314.Google Scholar
Jamniczky, H. A., Brinkman, D. B., and Russell, A. P.. 2003. Vertebrate microsite sampling: how much is enough? Journal of Vertebrate Paleontology 23:725734.Google Scholar
Jerzykiewicz, T., and Norris, D. K.. 1994. Stratigraphy, structure and syntectonic sedimentation of the Campanian “Belly River” clastic wedge in the southern Canadian Cordillera. Cretaceous Research 15:367399.Google Scholar
Johnson, R. G. 1960. Models and methods for analysis of the mode of formation of fossil assemblages. Geological Society of America Bulletin 71:10751086.Google Scholar
Kauffman, E. G. 1977. Geological and biological overview: Western Interior Cretaceous basin. Mountain Geology 14:7599.Google Scholar
Khajuria, C. K., and Prasad, G. V. R.. 1998. Taphonomy of a Late Cretaceous mammal-bearing microvertebrate assemblage from the Deccan inter-trappean beds of Naskal, peninsular India. Palaeogeography, Palaeoclimatology, Palaeoecology 137:153172.Google Scholar
Korth, W. A. 1979. Taphonomy of microvertebrate fossil assemblages. Annals of the Carnegie Museum 48:235285.Google Scholar
Kowalewski, M., and Hoffmeister, A. P.. 2003. Sieves and fossils: effects of mesh size on paleontological patterns. Palaios 18:460469.Google Scholar
Lawrence, D. R. 1968. Taphonomy and information losses in fossil communities. Geological Society of America Bulletin 79:13151330.Google Scholar
Leidy, J. 1856. Notices of the remains of extinct reptiles and fishes discovered by Dr. F. V. Hayden in the badlands of the Judith River, Nebraska Territory. Proceedings of the Academy of Natural Sciences of Philadelphia 8:7273.Google Scholar
Leidy, J. 1860. Extinct Vertebrata from the Judith River and Great Lignite formations of Nebraska. Transactions of the American Philosophical Society 11:139154.Google Scholar
Lillegraven, J. A., and Eberle, J. J.. 1999. Vertebrate faunal changes through Lancian and Puercan time in southern Wyoming. Journal of Paleontology 73:691710.Google Scholar
Lillegraven, J. A., and McKenna, M. C.. 1986. Fossil mammals from the “Mesaverde” Formation (Late Cretaceous, Judithian) of the Bighorn and Wind River basins, Wyoming: with definitions of Late Cretaceous North American land-mammal ages. American Museum Novitates 2840:168.Google Scholar
Lyman, R. L. 2012. The influence of screen mesh size, and size and shape of rodent teeth on recovery. Journal of Archaeological Science 39:18541861.Google Scholar
Maas, M. C. 1985. Taphonomy of a late Eocene microvertebrate locality, Wind River Basin, Wyoming (USA). Palaeogeography, Palaeoclimatology, Palaeoecology 52:123142.Google Scholar
McKenna, M. C. 1962. Collecting small fossils by washing and screening. Curator 5:221235.Google Scholar
Meek, F. B., and Hayden, F. V.. 1856. Descriptions of a new species of Acephala and Gastropoda from the Tertiary formations of Nebraska Territory, with some general remarks on the geology of the country about the sources of the Missouri River. Proceedings of the Academy of Natural Sciences of Philadelphia 8:111126.Google Scholar
Mellett, J. 1974. Scatological origins of microvertebrate fossil accumulations. Science 185:349350.Google Scholar
Moore, J. R. 2012. Do terrestrial vertebrate fossil assemblages show consistent taphonomic patterns? Palaios 27:220234.Google Scholar
Moore, J. R., and Norman, D. B.. 2009. Quantitatively evaluating the sources of taphonomic biasing of skeletal element abundances in fossil assemblages. Palaios 24:591602.Google Scholar
Moss, A. J., and Walker, P. H.. 1978. Particle transport by continental water flows in relation to erosion, deposition, soils, and human activities. Sedimentary Geology 20:81139.Google Scholar
Nell, L. A., Frederick, P. C., Mazzotti, F. J., Vliet, K. A., and Brandt, L. S.. 2016. Presence of breeding birds improves body condition for a crocodilian nest protector. PLoS ONE 11(3), e0149572. doi: 10.1371/journal.pone.0149572.Google Scholar
Ogg, J. G., and Hinnov, L. A.. 2012. Cretaceous. Pp. 793853 in F. M. Gradstein, J. G. Ogg, M. D. Schmitz, and G. M. Ogg, eds. The geologic time scale, 2012. Elsevier, Amsterdam.Google Scholar
Ottens, K. J., Dietl, G. P., Kelley, P. H., and Stanford, S. D.. 2012. A comparison of analyses of drilling predation on fossil bivalves: bulk- vs. taxon-specific sampling and the role of collector experience. Palaeogeography, Palaeoclimatology, Palaeoecology 319:8492.Google Scholar
Payne, S. 1972. Partial recovery and sample bias: the results of some sieving experiments. Pp. 4964 in E. S. Higgs, ed. Papers in economic prehistory. Cambridge University Press, Cambridge.Google Scholar
Peng, J., Russell, A. P., and Brinkman, D. B.. 2001. Vertebrate microsite assemblages (exclusive of mammals) from the Foremost and Oldman formations of the Judith River Group (Campanian) of southeastern Alberta: an illustrated guide. Provincial Museum of Alberta Natural History Occasional Paper 25:154.Google Scholar
Peterson, J. E., Scherer, R. P., and Huffman, K. M.. 2011. Methods of microvertebrate sampling and their influences on taphonomic interpretations. Palaios 26:8188.Google Scholar
Rasband, W.S. 1997–2015. ImageJ. U. S. National Institutes of Health, Bethesda, Md. http://imagej.nih.gov/ij/.Google Scholar
Raup, D. M. 1976. Species diversity in the Phanerozoic: an interpretation. Paleobiology 2:289297.Google Scholar
Rogers, R. R. 1994. Nature and origin of through-going discontinuities in nonmarine foreland basin deposits, Upper Cretaceous, Montana: implications for sequence analysis. Geology 22:11191122.Google Scholar
Rogers, R. R. 1998. Sequence analysis of the upper Cretaceous Two Medicine and Judith River Formations, Montana: nonmarine response to the Claggett and Bearpaw marine cycles. Journal of Sedimentary Research 68:615631.CrossRefGoogle Scholar
Rogers, R. R., and Brady, M. E.. 2010. Origins of microfossil bonebeds: insights from the upper Cretaceous Judith River Formation of north-central Montana. Paleobiology 36:80112.Google Scholar
Rogers, R. R., and Kidwell, S. M.. 2000. Associations of vertebrate skeletal concentrations and discontinuity surfaces in terrestrial and shallow marine records: a test in the Cretaceous of Montana. Journal of Geology 108:131154.Google Scholar
Rogers, R. R., and Kidwell, S. M.. 2007. A conceptual framework for the genesis and analysis of vertebrate skeletal concentrations. Pp. 1–63 in Rogers et al. 2007.Google Scholar
Rogers, R. R., Eberth, D. A., and Fiorillo, A. R.. 2007. Bonebeds: genesis, analysis, and paleobiological significance. University of Chicago Press, Chicago.Google Scholar
Rogers, R. R., Fricke, H. C., Addona, V., Canavan, R. R., Dwyer, C. N., Harwood, C. L., Koenig, A. E., Murray, R., Thole, J. T., and Williams, J.. 2010. Using laser ablation–inductively coupled plasma–mass spectrometry (LA-ICP-MS) to explore geochemical taphonomy of vertebrate fossils in the Upper Cretaceous Two Medicine and Judith River formations of Montana. Palaios 25:183195.Google Scholar
Rogers, R. R., Kidwell, S. M., Deino, A. L., Mitchell, J. P., Nelson, K., and Thole, J. T.. 2016. Age, correlation, and lithostratigraphic revision of the Upper Cretaceous (Campanian) Judith River Formation in its type area (north-central Montana), with a comparison of low- and high-accommodation alluvial records. Journal of Geology 124:99135.Google Scholar
Sahni, A. 1972. The vertebrate fauna of the Judith River Formation, Montana. Bulletin of the American Museum of Natural History 147:321412.Google Scholar
Sankey, J. T. 2001. Late Campanian southern dinosaurs, Aguja Formation, Big Bend, Texas. Journal of Paleontology 75:208215.Google Scholar
Sessa, J. A., Patzkowsky, M. E., and Bralower, T. J.. 2009. The impact of lithification on the diversity, size distribution, and recovery dynamics of marine invertebrate assemblages. Geology 37:115118.Google Scholar
Simpson, G. G. 1926. Mesozoic Mammalia, V; Dromatherium and Microconodon . American Journal of Science 68:87108.Google Scholar
Sloan, R. E. 1969. Cretaceous and Paleocene terrestrial communities of western North America. Proceedings of the North American Paleontological Convention E:427–453.Google Scholar
Sloan, R. E., and Van Valen, L.. 1965. Cretaceous mammals from Montana. Science 148:220227.Google Scholar
Smith, A. B. 2001. Large-scale heterogeneity of the fossil record: implications for Phanerozoic biodiversity studies. Philosophical Transactions of the Royal Society of London B 356:351367.Google Scholar
Smith, G. R., Stearley, R. F., and Badgley, C. E.. 1988. Taphonomic bias in fish diversity from Cenozoic floodplain environments. Palaeogeography, Palaeoclimatology, Palaeoecology 63:263273.Google Scholar
Srivastava, R., and Kumar, K.. 1996. Taphonomy and palaeoenvironment of the middle Eocene rodent localities of northwestern Himalaya, India. Palaeogeography, Palaeoclimatology, Palaeoecology 122:185211.Google Scholar
Stanton, T. W., and Hatcher, J. B.. 1905. Geology and paleontology of the Judith River beds. U.S. Geological Survey Bulletin 257:1128.Google Scholar
Wilson, L. E. 2008. Comparative taphonomy and paleoecological reconstruction of two microvertebrate accumulations from the Late Cretaceous Hell Creek Formation (Maastrichtian), eastern Montana. Palaios 23:289297.Google Scholar
Wolff, R. G. 1973. Hydrodynamic sorting and ecology of a Pleistocene mammalian assemblage from California (U.S.A.). Palaeogeography, Palaeoclimatology, Palaeoecology 13:91101.Google Scholar
Wolff, R. G. 1975. Sampling and sample size in ecological analyses of fossil mammals. Paleobiology 1:195204.Google Scholar
Wood, J. M., Thomas, R. G., and Visser, J.. 1988. Fluvial processes and vertebrate taphonomy: The Upper Cretaceous Judith River Formation, south central Dinosaur Provincial Park, Alberta, Canada. Palaeogeography, Palaeoclimatology, Palaeoecology 66:127143.Google Scholar