Hostname: page-component-848d4c4894-5nwft Total loading time: 0 Render date: 2024-04-30T14:38:24.567Z Has data issue: false hasContentIssue false

Concretions as Agents of Soft-Tissue Preservation: A Review

Published online by Cambridge University Press:  21 July 2017

Victoria E. McCoy*
Affiliation:
Yale University, Department of Geology and Geophysics, PO Box 208109, New Haven, CT 06520 USA
Get access

Abstract

Carbonate concretions may preserve exceptional soft-tissue fossils. Organismal decay in sediment can produce HCO3 faster than it diffuses away, creating a local microenvironment of high alkalinity around the decaying organism that promotes carbonate precipitation. Infilling of sediment pore-space around the decaying organism decreases permeability and inhibits decay, thus increasing preservation potential within the concretion and promoting soft-tissue preservation. Different patterns of concretion growth may promote different mechanisms of soft-tissue preservation. Other factors such as shifting salinity and clay mineral chemistry, which increase preservation potential in the depositional environment, also promote soft tissue preservation within concretions. The rate of concretion nucleation and growth affects preservation; the faster concretions nucleate and grow, the better the preservation. Concretionary preservation is biased both by concretion-promoting environments and organism size effects on concretion formation.

Type
Research Article
Copyright
Copyright © 2014 by 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

Allison, P. A. 1988. The role of anoxia in the decay and mineralization of proteinaceous macro-fossils. Paleobiology, 14:139154.CrossRefGoogle Scholar
Allison, P. A., and Briggs, D. E. G. 1993. Exceptional fossil record: distribution of soft-tissue preservation through the Phanerozoic. Geology, 21:527530.Google Scholar
Allison, P. A., and Pye, K. 1994. Early diagenetic mineralization and fossil preservation in modern carbonate concretions. PALAIOS, 9:561575.Google Scholar
Archer, A. W., Kuecher, G. J., and Kvale, E. P. 1995. The role of tidal-velocity asymmetries in the deposition of silty tidal rhythmites (Carboniferous, eastern Interior Coal Basin, U.S.A.). Journal of Sedimentary Research, 65:408416.Google Scholar
Astin, T. R., and Scotchman, I. C. 1988. The diagenetic history of some septarian concretions from the Kimmeridge Clay, England. Sedimentology, 35:349368.Google Scholar
Babcock, L. E., and Ciampaglio, C. N. 2007. Frondose fossil from the Conasauga Formation (Cambrian: Drumian Stage) of Georgia, USA. Memoirs of the Association of Australasian Palaeontologists, 34:555562.Google Scholar
Baird, G. C. 1981. Submarine erosion on a gentle paleoslope: a study of two discontinuities in the New York Devonian. Lethaia, 14:105122.Google Scholar
Baird, G. C. 1997a. Francis Creek diagenetic events, p. 3034. In Shabica, C. W. and Hay, A. A. (eds.), Richardson's Guide to the Fossil Fauna of Mazon Creek. Northeastern Illinois University Press, Chicago.Google Scholar
Baird, G. C. 1997b. Geologic setting of the Mazon Creek area fossil deposit, p. 1620. In Shabica, C. W. and Hay, A. A. (eds.), Richardson's Guide to the Fossil Fauna of Mazon Creek. Northeastern Illinois University Press, Chicago.Google Scholar
Baird, G. C., Shabica, C. W., Anderson, J. L., and Richardson, E. S. Jr. 1985a. Biota of a Pennsylvanian muddy coast: habitats within the Mazonian delta complex, Northeast Illinois. Journal of Paleontology, 59:253281.Google Scholar
Baird, G. C., Sroka, S. D., Shabica, C. W., Beard, T. L., Scott, A. C., and Broadhurst, F. M. 1985b. Mazon Creek-type fossil assemblages in the U.S. Midcontinent Pennsylvanian: their recurrent character and palaeoenvironmental significance [and discussion]. Philosophical Transactions of the Royal Society of London B-Biological Sciences, 311:8799.Google Scholar
Baird, G. C., Sroka, S. D., Shabica, C. W., and Kuecher, G. J. 1986. Taphonomy of middle Pennsylvanian Mazon Creek area fossil localities, Northeast Illinois: significance of exceptional fossil preservation in syngenetic concretions. PALAIOS, 1:271285.Google Scholar
Berner, R. A. 1968. Calcium carbonate concretions formed by the decomposition of organic matter. Science, 159:195197.CrossRefGoogle ScholarPubMed
Berner, R. A. 1980. Early Diagenesis: A Theoretical Approach. Princeton University Press, Princeton, N.J. Google Scholar
Berner, R. A. 1981. Authigenic mineral formation resulting from organic matter decomposition in modern sediments. Fortschritte der Mineralogie, 59:117135.Google Scholar
Bottjer, D. J., Etter, W., Hagadorn, J. W., and Tang, C. M. 2002. Exceptional Fossil Preservation: A Unique View on the Evolution of Marine Life. Columbia University Press, New York.Google Scholar
Brett, C. E., Allison, P. A., and Hendy, A. J. W. 2011. Comparative taphonomy and sedimentology of small-scale mixed carbonate/siliciclastic cycles: Synopsis of Phanerozoic examples, p. 108198. In Allison, P. A. and Bottjer, D. J. (eds.), Taphonomy: Process and Bias through Time. Springer, Dordrecht.Google Scholar
Brett, C. E., Zambito, J. J., Hunda, B. R., and Schindler, E. 2012. Mid-Paleozoic trilobite Lagerstätten: models of diagenetically enhanced obrution deposits. PALAIOS, 27:326345.Google Scholar
Briggs, D. E. G. 2003. The role of decay and mineralization in the preservation of soft-bodied fossils. Annual Review of Earth and Planetary Sciences, 31:275301.Google Scholar
Briggs, D. E. G., and Gall, J. C. 1990. The continuum in soft-bodied biotas from transitional environments: a quantitative comparison of Triassic and Carboniferous Konservat-Lagerstätten. Paleobiology, 16:204218.Google Scholar
Briggs, D. E. G., and Kear, A. J. 1993. Decay and preservation of polychaetes: taphonomic thresholds in soft-bodied organisms. Paleobiology, 19:107135.CrossRefGoogle Scholar
Briggs, D. E. G., Kear, A. J., Martill, D. M., and Wilby, P. R. 1993. Phosphatization of soft-tissue in experiments and fossils. Journal of the Geological Society, 150:10351038.CrossRefGoogle Scholar
Briggs, D. E. G., Siveter, D. J., and Siveter, D. J. 1996. Soft-bodied fossils from a Silurian volcaniclastic deposit. Nature, 382:248250.Google Scholar
Briggs, D. E. G., Siveter, D. J., Siveter, D. J., and Sutton, M. D. 2008. Virtual fossils from 425 million-year-old volcanic ash. American Scientist, 96:474481.CrossRefGoogle Scholar
Butterfield, N. J. 1990. Organic preservation of non-mineralizing organisms and the taphonomy of the Burgess Shale. Paleobiology, 16:272286.Google Scholar
Butterfield, N. J. 1995. Secular distribution of Burgess-Shale-type preservation. Lethaia, 28:113.Google Scholar
Canfield, D. C., and Raiswell, R. 1991. Carbonate precipitation and dissolution, p. 411453. In Allison, P. A. A. and Briggs, D. E. G. (eds.), Taphonomy: Releasing the Data Locked in the Fossil Record. Plenum Press, New York.Google Scholar
Charbonnier, S. 2009. Le Lagerstätte de La Voulte: Un Environnement Bathyal au Jurassique. Mémoires du Muséum National d'Histoire Naturelle 199. Publications Scientifiques du Muséum, Paris. [In French] Google Scholar
Charbonnier, S., Vannier, J., Galtier, J., Perrier, V., Chabard, D., and Sotty, D. 2008. Diversity and paleoenvironment of the flora from the nodules of the Montceau-Les-Mines biota (Late Carboniferous, France). PALAIOS, 23:210222.Google Scholar
Ciampaglio, C. N., Babcock, L. E., Wellman, C. L., York, A. R., and Brunswick, H. K. 2006. Phylogenetic affinities and taphonomy of Brooksella from the Cambrian of Georgia and Alabama, USA. Palaeoworld, 15:256265.CrossRefGoogle Scholar
Cloutier, R. 2010. The fossil record of fish ontogenies: insights into developmental patterns and processes. Seminars in Cell & Developmental Biology, 21:400413.Google Scholar
Coleman, M. L. 1993. Microbial processes: Controls on the shape and composition of carbonate concretions. Marine Geology, 113:127140.Google Scholar
Coleman, M. L., and Raiswell, R. 1995. Source of carbonate and origin of zonation in pyritiferous carbonate concretions: evaluation of a dynamic model. American Journal of Science, 295:282308.Google Scholar
Coleman, M. L., Raiswell, R., Brown, A., Curtis, C. D., Aplin, A. C., Ortoleva, P. J., Gruszczynski, M., Lyons, T., Lovely, D. R., and Eglinton, G. 1993. Microbial mineralization of organic matter: mechanisms of self-organization and inferred rates of precipitation of diagenetic minerals [and discussion]. Philosophical Transactions of the Royal Society of London Series A-Physical and Engineering Sciences, 344:6987.Google Scholar
Coniglio, M., and Cameron, J. S. 1990. Early diagenesis in a potential oil shale; evidence from calcite concretions in the Upper Devonian Kettle Point Formation, southwestern Ontario. Bulletin of Canadian Petroleum Geology, 38:6477.Google Scholar
Curtis, C. D., Coleman, M. L., and Love, L. G. 1986. Pore water evolution during sediment burial from isotopic and mineral chemistry of calcite, dolomite and siderite concretions. Geochimica et Cosmochimica Acta, 50:23212334.Google Scholar
Dick, V. B., and Brett, C. E. 1986. Petrology, taphonomy, and sedimentary environments of pyritic fossil beds from the Hamilton Group (Middle Devonian) of western New York, p. 102127. In Brett, C. E. (ed.), Dynamic Stratigraphy and Depositional Environments of the Hamilton Group (Middle Devonian) in New York State, Part 1. New York State Museum Bulletin 457. University of the State of New York, the State Education Dept., State Science Service, Cultural Education Center, Albany, New York.Google Scholar
Dornbos, S. Q. 2011. Phosphatization through the Phanerozoic, p. 435456. In Allison, P. A. and Bottjer, D. J. (eds.), Taphonomy: Process and Bias through Time. Springer, Dordrecht.Google Scholar
Fara, E., Saraiva, A. Á. F., De Almeida Campos, D., Moreira, J. K. R., De Carvalho Siebra, D., and Kellner, A. W. A. 2005. Controlled excavations in the Romualdo Member of the Santana Formation (Early Cretaceous, Araripe Basin, northeastern Brazil): stratigraphic, palaeoenvironmental and palaeoecological implications. Palaeogeography, Palaeoclimatology, Palaeoecology, 218:145160.Google Scholar
Felitsyn, S., and Morad, S. 2002. REE patterns in latest Neoproterozoic—early Cambrian phosphate concretions and associated organic matter. Chemical Geology, 187:257265.Google Scholar
Fernandes, A. S. 2012. A geobiological investigation of the Mazon Creek concretions of northeastern Illinois, mechanisms of formation and diagenesis. Unpublished Master's Thesis, The University of Western Ontario, 198 p.Google Scholar
Filipiak, P., and Krawczynski, W. 1996. Westphalian xiphosurans (Chelicerata) from the Upper Silesia Coal Basin of Sosnowiec, Poland. Acta Palaeontologica Polonica, 41:413425.Google Scholar
Fischer, J.-C. 2003. Remarkable invertebrates from the Lower Callovian of La Voulte-sur-Rhone (Ardeche, France). Annales de Paleontologie, 89:223252. [In French] CrossRefGoogle Scholar
Fisher, Q. J., Raiswell, R., and Marshall, J. D. 1998. Siderite concretions from nonmarine shales (Westphalian A) of the Pennines, England; controls on their growth and composition. Journal of Sedimentary Research, 68:10341045.Google Scholar
Foree, E. G., and McCarty, P. L. 1970. Anaerobic decomposition of algae. Environmental Science & Technology, 4:842849.CrossRefGoogle Scholar
Foster, M. 1979. Soft-bodied coelenterates in the Pennsylvanian of Illinois, p. 191267. In Nitecki, M. H. (ed.), Mazon Creek Fossils. Academic Press, New York.Google Scholar
Gaines, R. R., Briggs, D. E. G., Orr, P. J., and Van Roy, P. 2012a. Preservation of giant anomalocaridids in silica-chlorite concretions from the early Ordovician of Morocco. PALAIOS, 27:317325.Google Scholar
Gaines, R. R., Hammerlund, E. U., Hou, X., Qi, C., Gabbott, S. E., Zhao, Y., Peng, J., and Canfield, D. E. 2012b. Mechanism for Burgess Shale-type preservation. Proceedings of the National Academy of Sciences, 109:51805184.Google Scholar
Gaines, R. R., Kennedy, M. J., and Droser, M. L. 2005. A new hypothesis for organic preservation of Burgess Shale taxa in the middle Cambrian Wheeler Formation, House Range, Utah. Palaeogeography, Palaeoclimatology, Palaeoecology, 220:193205.Google Scholar
Garwood, R., Dunlop, J. A., and Sutton, M. D. 2009. High-fidelity x-ray micro-tomography reconstruction of siderite-hosted Carboniferous arachnids. Biology Letters, 5:841844.Google Scholar
Howarth, M. K. 1962. The Jet Rock series and the Alum Shale series of the Yorkshire Coast. Proceedings of the Yorkshire Geological and Polytechnic Society, 33:381422.Google Scholar
Kraus, O. 2005. On the structure and biology of Arthropleura species (Atelocerata, Diplopoda; Upper Carboniferous/Lower Permian). Verhandlungen des Naturwissenschaftlichen Vereins zu Hamburg, 41:523.Google Scholar
Landman, N. H., and Klofak, S. M. 2012. Anatomy of a concretion: Life, death, and burial in the Western Interior Seaway. PALAIOS, 27:671692.Google Scholar
Legg, D. A., Garwood, R. J., Dunlop, J. A., and Sutton, M. D. 2012. A taxonomic revision of orthosternous scorpions from the English Coal-measures aided by X-ray micro-tomography (XMT). Palaeontologia Electronica, 15:14A.Google Scholar
Leonowicz, P. 2010. Origin of siderites from the Lower Jurassic Ciechocinek Formation from SW Poland. Geological Quarterly, 51:6778.Google Scholar
Liu, X. 2000. Heterogeneous nucleation or homogeneous nucleation? The Journal of Chemical Physics, 112:99499955.Google Scholar
Long, J. A., and Trinajstic, K. 2010. The Late Devonian Gogo Formation Lagerstätte of Western Australia: exceptional early vertebrate preservation and diversity. Annual Review of Earth and Planetary Sciences, 38:255279.Google Scholar
Machemer, S. D., and Hutcheon, I. D. 1988. Geochemistry of early carbonate cements in the Cardium Formation, central Alberta. Journal of Sedimentary Research, 58:136147.Google Scholar
Maeda, H., Tanaka, G., Shimobayashi, N., Ohno, T., and Matsuoka, H. 2011. Cambrian Orsten Lagerstätte from the Alum Shale Formation: fecal pellets as a probably source of phosphorus preservation. PALAIOS, 26:225231.Google Scholar
Maples, C. G. 1986. Enhanced paleoecological and paleoenvironmental interpretations result from analysis of early diagenetic concretions in Pennsylvanian shales. PALAIOS, 1:512516.Google Scholar
Markov, I. V. 1995. Crystal growth for beginners: fundamentals of nucleation, crystal growth, and epitaxy. World Scientific Singapore.Google Scholar
Martill, D. M. 1988. Preservation of fish in the Cretaceous Santana Formation of Brazil. Palaeontology, 31:1.Google Scholar
Martill, D. M. 2007. The age of the Cretaceous Santana Formation fossil Konservat Lagerstätte of north-east Brazil: a historical review and an appraisal of the biochronostratigraphic utility of its palaeobiota. Cretaceous Research, 28:895920.Google Scholar
Martill, D. M., and Wilby, P. R. 1994. Lithified prokaryotes associated with fossil soft tissues from the Santana Formation (Cretaceous) of Brazil. Darmstadter Beitrage zur Natugeschichte, 4:7177.Google Scholar
McCoy, V. E., Young, R., and Briggs, D. E. G. In review. Factors controlling exceptional preservation within concretions. PALAIOS.Google Scholar
McGowan, A. J., and Smith, A. B. 2009. Are global Phanerozoic marine diversity curves truly global? A study of the relationship between regional rock records and global Phanerozoic marine diversity. Paleobiology, 34:80103.Google Scholar
Mozley, P. S. 1989. Complex compositional zonation in concretionary siderite: implications for geochemical studies. Journal of Sedimentary Research, 59:815818.Google Scholar
Mozley, P. S., and Burns, S. J. 1993. Oxygen and carbon isotopic composition of marine carbonate concretions; an overview. Journal of Sedimentary Research, 63:7383.Google Scholar
Müller, K. J. 1985. Exceptional preservation in calcareous nodules. Philosophical Transactions of the Royal Society of London B-Biological Sciences, 311:6773.Google Scholar
Orr, P. J., Briggs, D. E. G., Siveter, D. J., and Siveter, D. J. 2000. Three-dimensional preservation of a non-biomineralized arthropod in concretions in Silurian volcaniclastic rocks from Herefordshire, England. Journal of the Geological Society, 157:173186.Google Scholar
Park, L. E., and Downing, K. F. 2001. Paleoecology of an exceptionally preserved arthropod fauna from lake deposits of the Miocene Barstow Formation, Southern California, U.S.A. PALAIOS, 16:175184.2.0.CO;2>CrossRefGoogle Scholar
Pearson, M. 1974. Siderite concretions from the Westphalian of Yorkshire: a chemical investigation of the carbonate phase. Mineralogical Magazine, 39:696699.Google Scholar
Pease, C. M. 1992. On the declining extinction and origination rates of fossil taxa. Paleobiology, 18:8992.Google Scholar
Pirrie, D., and Marshall, J. D. 1991. Field relationships and stable isotope geochemistry of concretions from James Ross Island, Antarctica. Sedimentary Geology, 71:137150.Google Scholar
Pye, K. 1984. SEM analysis of siderite cements in intertidal marsh sediments, Norfolk, England. Marine Geology, 56:112.Google Scholar
Pye, K., Dickson, J. A. D., Schiavon, N., Coleman, M. L., and Cox, M. 1990. Formation of siderite-Mg-calcite-iron sulphide concretions in intertidal marsh and sandflat sediments, north Norfolk, England. Sedimentology, 37:325343.Google Scholar
Raiswell, R. 1971. The growth of Cambrian and Liassic concretions. Sedimentology, 17:147171.Google Scholar
Raiswell, R. 1976. The microbiological formation of carbonate concretions in the Upper Lias of NE England. Chemical Geology, 18:227244.Google Scholar
Raiswell, R. 1987. Non-steady state microbiological diagenesis and the origin of concretions and nodular limestones. Geological Society of London, Special Publications 36:4154.Google Scholar
Raiswell, R. 1988. Evidence for surface reaction-controlled growth of carbonate concretions in shales. Sedimentology, 35:571575.Google Scholar
Raiswell, R., Bottrell, S. H., Dean, S. P., Marshall, J. D., Carr, A., and Hatfield, D. 2002. Isotopic constraints on growth conditions of multiphase calcite–pyrite–barite concretions in Carboniferous mudstones. Sedimentology, 49:237254.Google Scholar
Raiswell, R., and Fisher, Q. J. 2000. Mudrockhosted carbonate concretions: a review of growth mechanisms and their influence on chemical and isotopic composition. Journal of the Geological Society, 157:239251.Google Scholar
Raiswell, R., and Fisher, Q. J. 2004. Rates of carbonate cementation associated with sulphate reduction in DSDP/ODP sediments: implications for the formation of concretions. Chemical Geology, 211:7185.Google Scholar
Raiswell, R., and White, N. 1978. Spatial aspects of concretionary growth in the Upper Lias of northeast England. Sedimentary Geology, 20:291300.Google Scholar
Richardson, E. S. 1980. Life at Mazon Creek, p. 217224. In Langenheim, R. J. and Mann, C. J. (eds.), Middle and Late Pennsylvanian Strata on the Margin of the Illinois Basin, Vermilion County, Illinois, Vermillion and Parke Counties, Indiana. 10th Annual Field Conference, Great Lakes Section, Society of Economic Paleontologists and Mineralogists. University of Illinois, Urbana.Google Scholar
Rietz, D. N., and Haynes, R. J. 2003. Effects of irrigation-induced salinity and sodicity on soil microbial activity. Soil Biology and Biochemistry, 35:845854.Google Scholar
Rogers, R. R., Arcucci, A. B., Abdala, F., Sereno, P. C., Forster, C. A., and May, C. L. 2001. Paleoenvironment and taphonomy of the Chanares Formation tetrapod assemblage (Middle Triassic), Northwestern Argentina: spectacular preservation in volcanogenic concretions. PALAIOS, 16:461481.2.0.CO;2>CrossRefGoogle Scholar
Sagemann, J., Bale, S. J., Briggs, D. E. G., and Parkes, R. J. 1999. Controls on the formation of authigenic minerals in association with decaying organic matter: an experimental approach. Geochimica et Cosmochimica Acta, 63:10831095.Google Scholar
Schram, F. R., and Nitecki, M. H. 1975. Hydra from the Illinois Pennsylvanian. Journal of Paleontology, 49:549551.Google Scholar
Schultze, H.-P. 1989. Three-dimensional muscle preservation in Jurassic fishes of Chile. Andean Geology, 16:183215.Google Scholar
Singhal, P. K., Gaur, S., and Talegaonkar, L. 1992. Relative contribution of different decay processes to the decomposition of Eichhornia crassipes (Mart.) solms. Aquatic Botany, 42:265272.Google Scholar
Sroka, S. D., and Baird, G. C. 1997. Freeze-thawing technique for opening nodules, p. 281281. In Shabica, C. W. and Hay, A. A. (eds.), Richardson's Guide to the Fossil Fauna of Mazon Creek. Northeastern Illinois University, Chicago.Google Scholar
Thompson, I. 1979. Errant polychaetes (Annelida) from the Pennsylvanian Essex fauna of northern Illinois. Palaeontographica Abteilung A, 163:169199.Google Scholar
Waagé, K. M., MacClintock, C., and Hickey, L. J. 2000. Post-glacial fossils from Long Island Sound off West Haven, Connecticut. Postilla, 225:126.Google Scholar
Walossek, D., and Müller, K. J. 1990. Upper Cambrian stem-lineage crustaceans and their bearing upon the monophyletic origin of Crustacea and the position of Agnostus . Lethaia, 23:409427.Google Scholar
Wenz, S., Brito, P. M., and Martill, D. M. 1993. The fish fauna of the Santana Formation concretions, p. 76107. In Martill, D. M. (ed.), Fossils of the Santana and Crato Formations, Brazil (Field Guides to Fossils 5). The Palaeontological Association, London.Google Scholar
Wilby, P. R., Briggs, D. E. G., and Riou, B. 1996. Mineralization of soft-bodied invertebrates in a Jurassic metalliferous deposit. Geology, 24:847850.Google Scholar
Wilson, D. D., and Brett, C. E. 2013. Concretions as sources of exceptional preservation, and decay as a source of concretions: examples from the Middle Devonian of New York. PALAIOS, 28:305316.Google Scholar
Xiao, S., Schiffbauer, J. D., McFadden, K. A., and Hunter, J. 2010. Petrographic and SIMS pyrite sulfur isotope analyses of Ediacaran chert nodules: implications for microbial processes in pyrite rim formation, silicification, and exceptional fossil preservation. Earth and Planetary Science Letters, 297:481495.CrossRefGoogle Scholar