Hostname: page-component-76fb5796d-22dnz Total loading time: 0 Render date: 2024-04-26T11:14:25.131Z Has data issue: false hasContentIssue false

Origins of microfossil bonebeds: insights from the Upper Cretaceous Judith River Formation of north-central Montana

Published online by Cambridge University Press:  08 April 2016

Raymond R. Rogers
Affiliation:
Geology Department, Macalester College, St. Paul, Minnesota 55105. E-mail: rogers@macalester.edu
Mara E. Brady
Affiliation:
Geology Department, Macalester College, St. Paul, Minnesota 55105. E-mail: rogers@macalester.edu

Abstract

Microfossil bonebeds are multi-individual accumulations of disarticulated and dissociated vertebrate hardparts dominated by elements in the millimeter to centimeter size range (≤75% of bioclasts ≤5 cm maximum dimension). Modes of accumulation are often difficult to decipher from reports in the literature, although predatory (scatological) and fluvial/hydraulic origins are typically proposed. We studied the sedimentology and taphonomy of 27 microfossil bonebeds in the Campanian Judith River Formation of Montana in order to reconstruct formative histories. Sixteen of the bonebeds examined are hosted by fine-grained facies that accumulated in low-energy aquatic settings (pond/lake microfossil bonebeds). Eleven of the bonebeds are embedded in sandstones that accumulated in ancient fluvial settings (channel-hosted microfossil bonebeds). In lieu of invoking separate pathways to accumulation based on facies distinctions, we present a model that links the accumulation of bioclasts in the two facies. We propose that vertebrate material initially accumulates to fossiliferous levels in ponds/lakes and is later reworked and redeposited as channel-hosted assemblages. This interpretation is grounded in reasonable expectations of lacustrine and fluvial depositional systems and supported by taphonomic data. Moreover, it is consistent with faunal data that indicate that channel-hosted assemblages and pond/lake assemblages are similar with regard to presence/absence and rank-order abundance of taxa.

This revised model of bonebed formation has significant implications for studies of vertebrate paleoecology that hinge on analyses of faunal data recovered from vertebrate microfossil assemblages. Pond/lake microfossil bonebeds in the Judith River record are preserved in situ at the scale of the local paleoenvironment, with no indication of postmortem transport into or out of the life habitat. Moreover, they are time-averaged samples of their source communities, which increases the likelihood of capturing both ecologically abundant species and more rare or transient members of the paleocommunity. These attributes make pond/lake microfossil bonebeds excellent targets for paleoecological studies that seek to reconstruct overall community membership and structure. In contrast, channel-hosted microfossil bonebeds in the Judith River record are out of place from a paleoenvironmental perspective because they are reworked from preexisting pond/lake assemblages and redeposited in younger channel facies. However, despite a history of exhumation and redeposition, channel-hosted microfossil bonebeds are preserved in relatively close spatial proximity to original source beds. This taphonomic reconstruction is counter to the commonly held view that microfossil bonebeds are biased samples that have experienced long-distance transport and significant hydrodynamic sorting.

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

Abbott, S. T. 1998. Transgressive systems tracts and onlap shellbeds from mid-Pleistocene sequences, Wanganui Basin, New Zealand. Journal of Sedimentary Research 68:253268.CrossRefGoogle Scholar
Allulee, J. L., and Holland, S. M. 2005. The sequence stratigraphic and environmental context of primitive vertebrates: Harding sandstone, Upper Ordovician, Colorado, USA. Palaios 20:518533.Google Scholar
Andrews, P. 1990. Owls, caves and fossils. Natural History Museum Publications, London.Google Scholar
Andrews, P., and Nesbit-Evans, E. M. 1983. Small mammal bone accumulations produced by mammalian carnivores. Paleobiology 9:289307.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
Argast, S., Farlow, J. O., and Gabet, R. M. 1987. Transport-induced abrasion of fossil reptilian teeth—implications for the existence of Tertiary dinosaurs in the Hell Creek Formation, Montana. Geology 15:927930.Google Scholar
Asian, A., and Behrensmeyer, A. K. 1996. Taphonomy and time resolution of bone assemblages in a contemporary fluvial system: the East Fork River, Wyoming. Palaios 11:411421.Google Scholar
Asian, A., and Blum, M. D. 1999. Contrasting styles of Holocene avulsion, Texas Gulf Coastal Plain, USA. Pp. 193209 in Smith, N. D. and Rogers, J., eds. Fluvial sedimentology. VI. Blackwell Science, Oxford.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 Tomida, Y., Flynn, L. J., and Jacobs, L. L., eds. Advances in vertebrate paleontology and geochronology. National Science Museum Monographs, Tokyo.Google Scholar
Behrensmeyer, A. K. 1975. The taphonomy and paleoecology of Plio-Pleistocene vertebrate assemblages east of Lake Rudolf, Kenya. Bulletin of the Museum of Comparative Zoology 146:473578.Google Scholar
Behrensmeyer, A. K. 1978. Taphonomic and ecologic information from bone weathering. Paleobiology 4:150162.CrossRefGoogle Scholar
Behrensmeyer, A. K. 1982. Time resolution in fluvial vertebrate assemblages. Paleobiology 8:211227.Google Scholar
Behrensmeyer, A. K. 1987. Miocene fluvial facies and vertebrate taphonomy in northern Pakistan. Society of Economic Paleontologists and Mineralogists Special Publication 39:169176.Google Scholar
Behrensmeyer, A. K. 1988. Vertebrate preservation in fluvial channels. Palaeogeography, Palaeoclimatology, Palaeoecology 63:183189.Google Scholar
Behrensmeyer, A. K., and Chapman, R. E. 1993. Models and simulations of time-averaging in terrestrial vertebrate accumulations. In Kidwell, S. M. and Behrensmeyer, A. K., eds. Taphonomic approaches to time resolution in fossil assemblages. Short Courses in Paleontology 6:125149. Paleontological Society, Knoxville, Tenn. Google Scholar
Blob, R. W. 1997. Relative hydrodynamic dispersal potentials of soft-shelled turtle elements: implications for interpreting skeletal sorting in assemblages of non-mammalian terrestrial vertebrates. Palaios 12:151164.Google Scholar
Blob, R. W., and Badgley, C. 2007. Numerical methods for bonebed analysis. Pp. 333396 in Rogers, et al. 2007.Google 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
Blob, R. W., Carrano, M. T., Rogers, R. R., Forster, C. A., and Espinoza, N. R. 1997. New taxonomic and taphonomic data from the herpetofauna of the Judith River Formation (Campanian), Montana. Journal of Vertebrate Paleontology 17(Suppl. to No. 3):3233.Google Scholar
Blob, R. W., Carrano, M. T., Rogers, R. R., Forster, C. A., and Espinoza, N. R. 2001. A new fossil frog from the Upper Cretaceous Judith River Formation of Montana. Journal of Vertebrate Paleontology 21:190194.Google Scholar
Bown, T. M., and Kraus, M. J. 1981. Vertebrate fossil-bearing paleosol units (Willwood Formation, Lower Eocene, Northwest Wyoming, U.S.A.): implications for taphonomy, biostratigraphy, and assemblage analysis. Palaeogeography, Palaeoclimatology, Palaeoecology 34:3156.Google Scholar
Brett, C. E. 1995. Sequence stratigraphy, biostratigraphy, and taphonomy in shallow marine environments. Palaios 10:597616.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 microfossil assemblages. Palaeogeography, Palaeoclimatology, Palaeoecology 213:295313.CrossRefGoogle Scholar
Brinkman, D. B., Eberth, D. A., and Currie, P. J. 2007. From bonebeds to paleobiology: applications of bonebed data. Pp. 221263 in Rogers, et al. 2007.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.Google Scholar
Buscalioni, A. D., Fregenal, M. A., Bravo, A., Poyato-Ariza, F. J., Sanchíz, B., Báez, A. M., Cambra Moo, O., Martín Closas, C., Evans, S. E., and Marugán Lobón, J. 2008. The vertebrate assemblage of Buenache de la Sierra (Upper Barremian of Serrania de Cuenca, Spain) with insights into its taphonomy and palaeoecology. Cretaceous Research 29:687710.Google Scholar
Canavan, R., Rogers, R., Koenig, A., Brady, M., Harwood, C. 2008. A geochemical approach to deciphering the origins of microfossil bonebeds in the Upper Cretaceous Judith River Formation, Montana. Journal of Vertebrate Paleontology 28(Suppl. to No. 3):60A.Google Scholar
Carrano, M. T., and Velez-Juarbe, J. 2006. Paleoecology of the Quarry 9 vertebrate assemblage from Como Bluff, Wyoming (Morrison Formation, Late Cretaceous). Palaeogeography, Palaeoclimatology, Palaeoecology 237:147159.CrossRefGoogle Scholar
Carrano, M. T., Blob, R. W., Flynn, J. J., Rogers, R. R., and Forster, C. A. 1997. The mammalian fauna of the Judith River Formation type area (Campanian, central Montana) revisited. Journal of Vertebrate Paleontology 17(Suppl. to No. 3):36.Google Scholar
Case, G. R. 1978. A new selachian fauna from the Judith River Formation (Campanian) of Montana. Palaeontographica Abteilung A 160:176205.Google Scholar
Citterio, A., and Piégay, H. 2009. Overbank sedimentation rates in former channel lakes: characterization and control factors. Sedimentology 56:461482.Google Scholar
Clemens, W. A., and Goodwin, M. B. 1985. Vertebrate paleontology of the Judith River Formation, Montana. National Geographic Society Research Reports 21:7178.Google Scholar
Conkin, J. E., Conkin, B. M., and Dasari, M. R. 1999. Sequential disconformities in the Devonian succession of southern Indiana and northwestern Kentucky. American Association of Petroleum Geologists Bulletin 83:1367.Google Scholar
Currie, P. J., and Koppelhus, E. B. 2005. Dinosaur Provincial Park: a spectacular ancient ecosystem revealed. Indiana University Press, Bloomington.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
Denys, C., and Mahboubi, M. 1992. Alterations structurales et chimiques des éléments squelettiques de pelotes de régurgitation d'un rapace diurne. Bulletin du Muséum National d'Histoire Naturelle, Paris, 4e série 14:229249.Google Scholar
Denys, C., Kowalski, K., and Dauphin, Y. 1992. Mechanical and chemical alterations of skeletal tissues in a recent Saharian accumulation of faeces from Vulpes rueppelli (Carnivora, Mammalia). Acta Zoologica Cracov 35:265283.Google Scholar
Dodson, P. 1971. Sedimentology and taphonomy of Oldman Formation (Campanian), Dinosaur Provincial Park, Alberta (Canada). Palaeogeography, Palaeoclimatology, Palaeoecology 10:2174.Google Scholar
Dodson, P. 1973. The significance of small bones in paleoecological interpretations. Rocky Mountain Geology 12:1519.Google Scholar
Dodson, P. 1987. Microfaunal studies of dinosaur paleoecology, Judith River Formation of southern Alberta. In Currie, P. J. and Koster, E., eds. Fourth symposium on Mesozoic terrestrial ecosytems, short papers. Royal Tyrrell Museum Occasional Paper 3:7075. Drumheller, Alberta.Google Scholar
Dodson, P., and Wexlar, D. 1979. Taphonomic investigation of owl pellets. Paleobiology 5:275284.Google Scholar
Dwyer, C., Rogers, R., Fricke, H., Wirth, K., and Thole, J. 2004. A comparative study of REE signatures and authigenic cements in dinosaur teeth and gar scales from the Upper Cretaceous Two Medicine and Judith River Formations of Montana. Geological Society of America Abstracts with Programs 37(5):81.Google Scholar
Dyke, G. J., and Malakhov, D. V. 2004. Abundance and taphonomy of dinosaur teeth and other vertebrate remains from the Botobynskaya Formation, north-east Aral Sea region, Republic of Kazakhstan. Cretaceous Research 25:669674.CrossRefGoogle 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. 2005. The geology. Pp. 5487 in Currie, and Koppelhus, 2005.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., Shannon, M., and Noland, B. G. 2007. A bonebeds database: classification, biases and patterns of occurrence. Pp. 103219 in Rogers, et al. 2007.Google Scholar
Eckblad, J. W., Peterson, N. L., Ostlie, K., and Tempte, A. 1977. The morphometry, benthos, and sedimentation rates of a floodplain lake in Pool 9 of the upper Mississippi River. American Midland Naturalist 97:433443.CrossRefGoogle 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. 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
Estes, R., Spinar, Z. V., and Nevo, E. 1978. Early Cretaceous pipid tadpoles from Israel (Amphibia: Anura). Herpetologica 34:374393 Google Scholar
Fiorillo, A. R. 1988. Taphonomy of Hazard Homestead Quarry (Ogallala Group), Hitchcock County, Nebraska. University of Wyoming Contributions to Geology 26:5797.Google Scholar
Fiorillo, A. R. 1989. The vertebrate fauna from the Judith River Formation (Late Cretaceous) of Wheatland and Golden Valley Counties, Montana. The Mosasaur 4:127142.Google Scholar
Fiorillo, A. R., Padian, K., and Musikasinthorn, C. 2000. Taphonomy and depositional setting of the Placerias quarry (Chinle Formation: Late Triassic, Arizona). Palaios 15:373386.Google Scholar
Fisher, D. C. 1981a. Crocodilian scatology, microvertebrate concentrations, and enamel-less teeth. Paleobiology 7:262275.Google Scholar
Fisher, D. C. 1981b. Mode of preservation of the Shotgun Local Fauna (Paleocene, Wyoming) and its implications for the taphonomy of a microvertebrate concentration. Contributions from the Museum of Paleontology, University of Michigan 25(12):247257.Google Scholar
Folk, R. L. 1980. Petrology of sedimentary rocks. Hemphill, Austin.Google Scholar
Foreman, B. Z., Rogers, R. R., Deino, A. L., Wirth, K. R., and Thole, J. T. 2008. Geochemical characterization of bentonite beds in the Two Medicine Formation (Campanian, Montana), including a new 40Ar/39Ar age. Cretaceous Research 29:373385.CrossRefGoogle Scholar
Fricke, H. C. 2007. Stable isotope geochemistry of bonebed fossils: reconstructing paleoenvironments, paleoecology, and paleobiology. Pp. 437490 in Rogers, et al. 2007.Google Scholar
Fricke, H. C., Rogers, R. R., Backlund, R., Dwyer, C. N., Echt, S. 2008. Preservation of stable isotope signals in dinosaur remains, and environmental gradients of the Late Cretaceous of Montana and Alberta. Palaeogeography, Palaeoclimatology, Palaeoecology 266:1327.Google Scholar
Frison, G. C., and Todd, L. C. 1986. The Colby Mammoth Site: taphonomy and archaeology of a Clovis kill in northern Wyoming. University of New Mexico Press, Albuquerque.Google Scholar
Froese, A. D., and Burghardt, G. M. 1975. A dense natural population of the common snapping turtle (Chelydra s. serpentina). Herpetologica 31:204208.Google Scholar
Gibbons, J. W., Winne, C. T., Scott, D. E., Willison, J. D., Glaudas, X., Andrews, K. M., Todd, B. D., Fedewa, L. A., Wilkison, L., Tsaliagos, R. N., Harper, S. J., Greene, J. L., Tuberville, T. D., Metts, B. S., Dorcas, M. E., Nestor, J. P., Young, C. A., Akre, T., Reed, R. N., Buhlmann, K. A., Norman, J., Croshaw, D. A., Hagen, C., and Rothermel, B. B. 2006. Remarkable amphibian biomass and abundance in an isolated wetland: implications for wetland conservation. Conservation Biology 20:14571465.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.Google Scholar
Gillespie, J. L., Nelson, C. S., and Nodder, S. D. 1998. Post-glacial sea-level control and sequence stratigraphy of carbonate-terrigenous sediments, Wanganui Shelf, New Zealand. Sedimentary Geology 122:245266.CrossRefGoogle 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.CrossRefGoogle Scholar
Google. 2007. Google Earth. http://earth.google.com. Google Scholar
Hanson, C. B. 1980. Fluvial taphonomic processes: models and experiments. Pp. 156181 in Behrensmeyer, A. K. and Hill, A. P., eds. Fossils in the making. University of Chicago Press, Chicago.Google Scholar
Harwood, C. L., Rogers, R. R., Koenig, A. E., Fricke, H. C., and Thole, J. T. 2005. A comparative study of authigenic mineralization and rare earth element geochemistry of vertebrate microfossil assemblages in the Campanian Judith River Formation of Montana. Geological Society of America Abstracts with Programs 37:301.Google Scholar
Haynes, G. 1988. Mass deaths and serial predation: comparative taphonomic studies of modern large mammal death sites. Journal of Archaeological Science 15:219235.Google Scholar
Henrici, A. C., and Fiorillo, A. R. 1993. Catastrophic death assemblage of Chelomophrynus bayi (Anura, Rhinophrynidae) from the Middle Eocene Wagon Bed Formation of central Wyoming. Journal of Vertebrate Paleontology 67:10161026.Google Scholar
Hoffman, R. 1988. The contribution of raptorial birds to patterning in small mammal assemblages. Paleobiology 14:8190.Google Scholar
Holland, W. J., and Burk, C. J. 1982. Relative ages of western Massachusetts oxbow lakes. Northeastern Geology 4:2332.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.CrossRefGoogle Scholar
Irmis, R., Pyenson, N., and Lipps, J. 2007. Formation of marine bonebeds: insights from the Middle Miocene Sharktooth Hill bonebed of California. Journal of Vertebrate Paleontology 27:94A.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
Jerolmack, D. J., and Paola, C. 2007. Complexity in a cellular model of river avulsion. Geomorphology 91:259270.Google Scholar
Kauffman, E. G. 1977. Geological and biological overview: Western Interior Cretaceous basin. Mountain Geologist 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
Kidwell, S. M. 1993. Influence of subsidence on the anatomy of marine siliciclastic sequences and on the distribution of shell and bone beds. Journal of the Geological Society, London 150:165167.CrossRefGoogle Scholar
Kidwell, S. M., and Flessa, K. 1996. The quality of the fossil record: populations, species, and communities. Annual Review of Earth and Planetary Sciences 24:433464.Google Scholar
Kidwell, S. M., and Holland, S. M. 1991. Field description of coarse bioclastic fabrics. Palaios 6:426434.CrossRefGoogle Scholar
Kidwell, S. M., Fürsich, F. T., and Aigner, T. 1986. Conceptual framework for the analysis and classification of fossil concentrations. Palaios 1:228238.Google Scholar
Koenig, A. E., Rogers, R. R., and Trueman, C. N. 2009. Visualizing fossilization using laser ablation-inductively coupled plasma-mass spectrometry maps of trace elements in Late Cretaceous bones. Geology 37:511514.Google Scholar
Kondo, Y., Abbott, S. T., Kitamura, A., Kamp, P. J. J., Naish, T. R., Kamataki, T., and Saul, G. S. 1998. The relationship between shellbed type and sequence architecture: examples from Japan and New Zealand. Sedimentary Geology 122:109128.Google Scholar
Korth, W. A. 1979. Taphonomy of microvertebrate fossil assemblages. Annals of the Carnegie Museum 48:235285.CrossRefGoogle Scholar
Koster, E. 1987. Vertebrate taphonomy applied to the analysis of ancient fluvial systems. Society of Economic Paleontologists and Mineralogists Special Publication 39:159168.Google Scholar
Kusmer, K. D. 1990. Taphonomy of owl pellet deposition. Journal of Paleontology 64:629637.Google Scholar
Laudet, F., and Selva, N. 2005. Ravens as small mammal bone accumulators: first taphonomic study on mammal remains as raven pellets. Palaeogeography, Palaeoclimatology, Palaeoecology 226:272286.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, new series 11:139154 Google Scholar
Lofgren, D. L., Hotton, C. L., and Runkel, A. C. 1990. Reworking of Cretaceous dinosaurs into Paleocene channel deposits, upper Hell Creek Formation, Montana. Geology 18:874877.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
Macquaker, J. H. S. 1994. Paleoenvironmental significance of bone-beds in organic-rich mudstone successions—an example from the Upper Triassic of South-West Britain. Zoological Journal of the Linnean Society 112:285308.Google Scholar
Mayhew, D. F. 1977. Avian predators as accumulators of fossil mammal material. Boreas 6:2531.CrossRefGoogle Scholar
McGrew, P. O. 1963. Environmental significance of sharks in the Shotgun Fauna, Paleocene of Wyoming. University of Wyoming Contributions to Geology 2:3941.Google Scholar
McKenna, M. C. 1960. Fossil mammalia from the early Eocene Wasatchian Four Mile fauna Eocene of northwestern Colorado. University of California Publications in Geological Sciences 37:1130.Google Scholar
McKenna, M. C. 1962. Collecting small fossils by washing and screening. Curator 5:221235.Google Scholar
Mellett, J. 1974. Scatalogical origins of microvertebrate fossil accumulations. Science 185:349350.CrossRefGoogle Scholar
Montellano, M. 1991. Mammalian fauna of the Judith River Formation (Late Cretaceous, Judithian) north central Montana. University of California Publications in Geological Sciences 136.Google Scholar
Neuman, A. G., and Brinkman, D. B. 2005. Fishes of the fluvial beds. Pp. 167185 in Currie, and Koppelhus, 2005.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
Prieto-Marquez, A. 2005. New information on the cranium of Brachylophosaurus canadensis (Dinosauria, Hadrosauridae); with a revision of its phylogenetic position. Journal of Vertebrate Paleontology 25:144156.Google Scholar
Pyenson, N. D., Irmis, R. B., Lipps, J. H., Barnes, L. G., Mitchell, E. D. Jr., and McLeod, S. A. 2009. Origin of a widespread marine bonebed deposited during the middle Miocene Climatic Optimum. Geology 37:519522.Google Scholar
Ralrick, P. 2006. Big trouble at Little Fish Lake: taphonomy of a diverse vertebrate mass mortality assemblage in Alberta, Canada. Journal of Vertebrate Paleontology 26:113A.Google Scholar
Räsänen, M. E., Salo, J. S., and Jungner, H. 1991. Holocene floodplain lake sediments in the Amazon: 14C dating and palaeocological use. Quaternary Science Reviews 10:363372.Google Scholar
Rogers, R. R. 1993. Systematic patterns of time averaging in the terrestrial vertebrate record: a Cretaceous case study. In Kidwell, S. M. and Behrensmeyer, A. K., eds. Taphonomic approaches to time resolution in fossil assemblages. Short Courses in Paleontology 6:228249. Paleontological Society, Knoxville, Tenn. Google Scholar
Rogers, R. R. 1995. Sequence stratigraphy and vertebrate taphonomy of the Upper Cretaceous Two Medicine and Judith River Formations, Montana. . University of Chicago, Chicago.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.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 Swisher, C. C. 1996. The Claggett and Bearpaw transgressions revisited; new 40Ar/39Ar data and a review of possible drivers. Geological Society of America Abstracts with Programs 28(6):62.Google Scholar
Rogers, R. R., Krause, D. W., and Curry Rogers, K. 2003. Cannibalism in the Madagascar dinosaur Majungatholus atopus . Nature 422:515518.Google Scholar
Rogers, R. R., Fricke, H. C., Koenig, A. E., Dwyer, C. N., Harwood, C. L., and Williams, J. 2005. A comparative study of diagenesis in fossil bones and teeth; a case study from the Upper Cretaceous Two Medicine and Judith River Formations. Journal of Vertebrate Paleontology Abstracts of Papers 25:106.Google Scholar
Rogers, R. R., Eberth, D. A., and Fiorillo, A. R. 2007. Bonebeds: genesis, analysis, and paleobiological significance. University of Chicago Press, Chicago.CrossRefGoogle Scholar
Ryan, M., Russell, A. P., and Eberth, D. A. 2001. The taphonomy of a Centrosaurus (Ornithischia: Certopsidae) bone bed from Dinosaur Park Formation (Upper Campanian), Alberta, Canada, with comments on cranial ontogeny. Palaios 16:482506.Google Scholar
Sahni, A. 1972. The vertebrate fauna of the Judith River Formation, Montana. American Museum of Natural History Museum Bulletin 147:321412.Google Scholar
Sankey, J. T. 2001. Late Campanian southern dinosaurs, Aguja Formation, Big Bend, Texas. Journal of Paleontology 75:208215.2.0.CO;2>CrossRefGoogle Scholar
Sankey, J. T., and Baszio, S. 2008. Vertebrate microfossil assemblages: their role in paleoecology and paleobiogeography. Indiana University Press, Bloomington.Google Scholar
Schiebout, J. A., White, P. D., and Boardman, G. S. 2008. Taphonomic issues relating to concentrations of pedogenic nodules and vertebrates in the Paleocene and Miocene Gulf Coastal Plain: Examples from Texas and Louisiana, USA. Pp. 1730 in Sankey, and Baszio, 2008.Google Scholar
Schmitt, D. N., and Juell, K. E. 1994. Toward the identification of coyote scatalogical faunal accumulations in archaeological contexts. Journal of Archaeological Science 21:249262.Google Scholar
Shipman, P. 1981. Applications of scanning electron-microscopy to taphonomic problems. Annals of the New York Academy of Sciences 376:357385.Google Scholar
Slingerland, R., and Smith, N. D. 2004. River avulsions and their deposits. Annual Review of Earth and Planetary Sciences 32:257285.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
Sykes, J. H. 1977. British Rhaetian bone-beds. Mercian Geologist 5:3948.Google Scholar
Terry, R. C. 2004. Owl pellet taphonomy: a preliminary study of the post-regurgitation history of pellets in a temperate forest. Palaios 19:497506.2.0.CO;2>CrossRefGoogle Scholar
Thomas, R. G., Smith, D. G., Wood, J. M., Visser, J., Calverly Range, E. A., and Koster, E. 1987. Inclined heterolithic stratification: terminology, description, interpretation, and significance. Sedimentary Geology 53:123179.Google Scholar
Trapani, J. 1998. Hydrodynamic sorting of avian skeletal remains. Journal of Archaeological Science 25:477487.Google Scholar
Trueman, C. 2007. Trace element geochemistry of bonebeds. Pp. 397435 in Rogers, et al. 2007.Google Scholar
Tulu, Y., and Rogers, R. 2004. Late Cretaceous chondrichthyans from the Woodhawk Bonebed, Judith River Formation (Campanian), Fergus County, Montana. Journal of Vertebrate Paleontology 24(Suppl. to No. 3):123.Google Scholar
Turner, A. H., Brett, C. E., McLaughlin, P. I., Over, D. J., and Storrs, G. W. 2001. Middle–Upper Devonian (Givetian–Famennian) bone/conodont beds from central Kentucky; reworking and event condensation in the distal Acadian foreland basin. Geological Society of America, North-Central Section, Abstracts with Programs 33(4):8.Google Scholar
Tweet, J. S., Chin, K., Bramen, D. R., and Murphy, N. L. 2008. Probable gut contents within a specimen of Brachylophosaurus canadensis (Dinosauria, Hadrosauridae) from the Upper Cretaceous Judith River Formation of Montana. Palaios 23:624635.Google Scholar
van der Valk, A. G. 2006. The biology of freshwater wetlands. Oxford University Press, New York.Google Scholar
Vasileiadou, K., Hooker, J. J., and Collinson, M. E. 2009. Paleocommunity reconstruction and accumulation of micromammalian remains (late Eocene, southern England). Palaios 24:553567.Google Scholar
Voorhies, M. R. 1969. Taphonomy and population dynamics of an early Pliocene vertebrate fauna, Knox County, Nebraska. University of Wyoming Contributions to Geology Special Paper 1:169.Google Scholar
Wetzel, R. G. 2001. Limnology: lake and river ecosystems, 3d ed. Academic Press, San Diego.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
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