Hostname: page-component-848d4c4894-4rdrl Total loading time: 0 Render date: 2024-07-03T01:49:49.426Z Has data issue: false hasContentIssue false

Pyritization of Soft Tissues in the Fossil Record: An Overview

Published online by Cambridge University Press:  21 July 2017

Úna C. Farrell*
Affiliation:
University of Kansas, Biodiversity Institute, 1475 Jayhawk Boulevard, Lawrence, KS 66045 USA
Get access

Abstract

Authigenic pyrite preserves non-biomineralized tissues in the fossil record under exceptional circumstances. Diagenetic models and taphonomic experiments demonstrate that active, localized sulfate reduction in iron-rich pore waters results in a strong concentration gradient, which confines pyrite precipitation to decaying organic matter. The locus and timing of pyrite precipitation is also influenced by the original composition of the organic matter. In recent decades, new sites with three-dimensional pyritized soft tissues have been discovered, although the Hunsrück Slate (Devonian) and Beecher's Trilobite Bed (Ordovician), known since the late 1800s, remain the primary examples in terms of diversity, abundance, and quality of preservation. Sedimentological and geochemical analyses at these sites have shown that rapid burial in fine-grained, reworked sediments sets up the high iron, low organic carbon conditions necessary for soft-tissue pyritization. Soft-tissue pyritization may also occur in association with other taphonomic modes, in particular with Burgess Shale-type preservation and carbonaceous preservation in lakes, although many of these specimens are now weathered. Continued comparison among sites and between specimens with variable degrees of preservation could help clarify the limits to soft-tissue pyritization and its distribution in ancient sediments.

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

Alessandrello, A., Bracchi, G., and Riou, B. 2004. Polychaete, sipunculan and enteropneust worms from the Lower Callovian (Middle Jurassic) of La Voulte-sur-Rhône (Ardèche, France). Memorie della Società Italiana di Scienze Naturali e del Museo Civico di Storia Naturale di Milano 32, Società Italiana di Scienze Naturali e Museo Civico di Storia Naturale di Milano, Milan.Google Scholar
Allen, R. E. 2002. Role of diffusion-precipitation reactions in authigenic pyritization. Chemical Geology, 182:461472.CrossRefGoogle Scholar
Aller, R. C. 1980. Quantifying solute distributions in the bioturbated zone of marine sediments by defining an average microenvironment. Geochimica et Cosmochimica Acta, 44:19551965.Google Scholar
Aller, R. C. 1998. Mobile deltaic and continental shelf muds as suboxic, fluidized bed reactors. Marine Chemistry, 61:143155.CrossRefGoogle Scholar
Aller, R. C., Mackin, J. E., and Cox, R. T. Jr. 1986. Diagenesis of Fe and S in Amazon inner shelf muds: apparent dominance of Fe reduction and implications for the genesis of ironstones. Continental Shelf Research, 6:263289.CrossRefGoogle Scholar
Aller, R. C., Madrid, V., Chistoserdov, A., Aller, J. Y., and Heilbrum, C. 2010. Unsteady diagenetic processes and sulfur biogeochemistry in tropical deltaic muds: implications for oceanic isotope cycles and the sedimentary record. Geochimica et Cosmochimica Acta, 74:46714692.Google Scholar
Allison, P. A. 1988. Konservat-Lagerstätten: cause and classification. Paleobiology, 14:331344.CrossRefGoogle Scholar
Allison, P. A., and Briggs, D. E. G. 1991. Taphonomy of non-mineralized tissues, p. 2670. In Allison, P. A. and Briggs, D. E. G. (eds.), Taphonomy: Releasing the Data Locked in the Fossil Record. Plenum Press, New York.Google Scholar
Anderson, E. P., Schiffbauer, J. D., and Xiao, S. 2011. Taphonomic study of Ediacaran organic-walled fossils confirms the importance of clay minerals and pyrite in Burgess Shale-type preservation. Geology, 39:643646.Google Scholar
Ausich, W. I., Bartels, C., and Kammer, T. W. 2013. Tube foot preservation in the Devonian crinoid Codiacrinus from the Lower Devonian Hunsrück Slate, Germany. Lethaia, 46:416420.CrossRefGoogle Scholar
Bartels, C., and Blind, W. 1995. Röntgenuntersuchung pyritisch vererzter Fossilien aus dem Hunsrückschiefer (Unter-Devon, Rheinisches Schiefergebirge). Metalla, 2:79100.Google Scholar
Bartels, C., and Brassel, G. 1990. Fossilien im Hunsrückschiefer: Dokumente des Meereslebens im Devon. Museum Idar-Oberstein, 232 p.Google Scholar
Bartels, C., Briggs, D. E. G., and Brassel, G. 1998. Fossils of the Hunsrück Slate, Marine Life in the Devonian. Cambridge University Press, Cambridge.Google Scholar
Bartels, C., and Poschmann, M. 2002. Linguloid brachiopods with preserved pedicles: Occurrence and taphonomy (Hunsrück Slate, Lower Emsian, Kaub Formation, Rhenish Massif, SW Germany). Metalla (Bochum), 9:123130.Google Scholar
Beecher, C. E. 1896. The morphology of Triarthrus . American Journal of Science, 1:251257.CrossRefGoogle Scholar
Bergmann, A., and Rust, J. 2014. Morphology, paleobiology and phylogeny of Oryctocaris balssi gen. nov. (Arthropoda), a phyllocarid from the Lower Devonian Hunsrück Slate (Germany). Journal of Systematic Palaeontology, 12:427444.Google Scholar
Bergström, J., and Brassel, G. 1984. Legs in the trilobite Rhenops from the Lower Devonian Hunsrück Slate. Lethaia, 17:6772.Google Scholar
Bergström, J., Briggs, D. E. G., Dahl, E., Rolfe, W. D. I., and Stürmer, W. 1987. Nahecaris stuertzi, a phyllocarid crustacean from the Lower Devonian Hunsrück Slate. Paläontologische Zeitschrift, 61(3–4):273298.Google Scholar
Berner, R. A. 1964. Iron sulfides formed from aqueous solution at low temperatures and atmospheric pressure. The Journal of Geology, 72:293306.Google Scholar
Berner, R. A. 1969. Migration of iron and sulfur within anaerobic sediments during early diagenesis. American Journal of Science, 267:1942.CrossRefGoogle Scholar
Berner, R. A. 1970. Sedimentary pyrite formation. American Journal of Science, 268:123.Google Scholar
Berner, R. A. 1978. Sulfate reduction and the rate of deposition of marine sediments. Earth and Planetary Science Letters, 37:492498.CrossRefGoogle Scholar
Berner, R. A. 1980. Early Diagenesis: A Theoretical Approach. Princeton University Press, Princeton, New Jersey.Google Scholar
Berner, R. A. 1984. Sedimentary pyrite formation: an update. Geochimica et Cosmochimica Acta, 48:605615.Google Scholar
Berner, R. A., and Raiswell, R. 1984. C/S method for distinguishing freshwater from marine sedimentary rocks. Geology, 12:365368.Google Scholar
Bertics, V. J., and Ziebis, W. 2010. Bioturbation and the role of microniches for sulfate reduction in coastal marine sediments. Environmental Microbiology, 12:30223034.CrossRefGoogle ScholarPubMed
Bickert, T. 2006. Influence of geochemical processes on stable isotope distribution in marine sediments p. 339369. In Schulz, H. D. and Zabel, M. (eds.), Marine Geochemistry. Springer, Berlin.CrossRefGoogle Scholar
Bloch, J., and Krouse, H. R. 1992. Sulfide diagenesis and sedimentation in the Albian Harmon member, Western Canada. Journal of Sedimentary Petrology, 62:235249.Google Scholar
Botting, J. P. 2004. An exceptional Caradoc sponge fauna from the Llanfawr Quarries, central Wales and phylogenetic implications. Journal of Systematic Palaeontology, 2:3163.Google Scholar
Botting, J. P., and Muir, L. A. 2013. Spicule structure and affinities of the Late Ordovician hexactinellidlike sponge Cyathophycus loydelli from the Llanfawr Mudstones Lagerstätte, Wales. Lethaia, 46:454469.Google Scholar
Botting, J. P., Muir, L. A., Sutton, M. D., and Barnie, T. 2011. Welsh gold: A new exceptionally preserved pyritized Ordovician biota. Geology, 39:879882.CrossRefGoogle Scholar
Boyer, D. L., Owens, J. D., Lyons, T. W., and Droser, M. L. 2011. Joining forces: Combined biological and geochemical proxies reveal a complex but refined high-resolution palaeooxygen history in Devonian epeiric seas. Palaeogeography, Palaeoclimatology, Palaeoecology, 306:134146.CrossRefGoogle Scholar
Brett, C. E. 1999. Middle Devonian Arkona Shale of Ontario, Canada, and Silica Shale of Ohio, USA, p. 129134. In Hess, H., Ausich, W. I., Brett, C. E., and Simms, M. J. (eds.), Fossil Crinoids. Cambridge University Press.Google Scholar
Brett, C. E., Algeo, T. J., and McLaughlin, P. I. 2008. Use of event beds and sedimentary cycles in high-resolution stratigraphic correlation of lithologically repetitive successions p. 315350. In Harries, P. J. (ed.), High-Resolution Approaches in Stratigraphic Paleontology. Springer.Google Scholar
Brett, C. E., and Baird, G. C. 1986. Comparative taphonomy: a key to paleoenvironmental interpretation based on fossil preservation. PALAIOS, 1:207227.Google Scholar
Briggs, D. E. G. 1995. Experimental taphonomy. PALAIOS, 10:539550.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 Bartels, C. 2001. New arthropods from the Lower Devonian Hunsrück Slate (Lower Emsian, Rhenish Massif, western Germany). Palaeontology, 44:275303.Google Scholar
Briggs, D. E. G., and Bartels, C. 2010. Annelids from the Lower Devonian Hunsrück Slate (lower Emsian, Rhenish Massif, Germany). Palaeontology, 53:215232.Google Scholar
Briggs, D. E. G., Bottrell, S. H., and Raiswell, R. 1991. Pyritization of soft-bodied fossils: Beecher's Trilobite Bed, Upper Ordovician, New York State. Geology, 19:12211224.Google Scholar
Briggs, D. E. G., and Edgecombe, G. D. 1993. Beecher's Trilobite Bed. Geology Today, 9:97102.CrossRefGoogle Scholar
Briggs, D. E. G., and Kear, A. J. 1993a. Decay and preservation of polychaetes: taphonomic thresholds in soft-bodied organisms. Paleobiology, 19:107135.CrossRefGoogle Scholar
Briggs, D. E. G., and Kear, A. J. 1993b. Fossilization of soft tissue in the laboratory. Science, 259:14391442.CrossRefGoogle ScholarPubMed
Briggs, D. E. G., and Kear, A. J. 1993c. Phosphatization of soft-tissue in experiments and fossils. Journal of the Geological Society, 150:10351038.Google Scholar
Briggs, D. E. G., and Kear, A. J. 1994. Decay and mineralization of shrimps. PALAIOS, 9:431456.Google Scholar
Briggs, D. E. G., Kear, A., Martill, D., and Wilby, P. 1993. Phosphatization of soft-tissue in experiments and fossils. Journal of the Geological Society, 150:10351038.Google Scholar
Briggs, D. E. G., Raiswell, R., Bottrell, S. H., Hatfield, D., and Bartels, C. 1996. Controls on the pyritization of exceptionally preserved fossils: an analysis of the Lower Devonian Hunsrück Slate of Germany. American Journal of Science, 296:633663.Google Scholar
Brock, F., Parkes, R. J., and Briggs, D. E. G. 2006. Experimental pyrite formation associated with decay of plant material. PALAIOS, 21:499506.Google Scholar
Broili, F. 1928. Crustaceenfunde aus dem rheinischen Unterdevon. Sitzungsberichte der Bayerischen Akademie der Wissenschaften Mathematischnaturwissenschaftliche Abteilung, München: 197201.Google Scholar
Brüchert, V. 2004. Physiological and ecological aspects of sulfur isotope fractionation during bacterial sulfate reduction, p. 116. In Amend, J. P., Edwards, K. J., and Lyons, T. W. (eds.), Sulfur Biogeochemistry: Past and Present. Geological Society of America Special Paper 379, Geological Society of America, Boulder, Colorado.Google Scholar
Butler, I. B., and Rickard, D. 2000. Framboidal pyrite formation via the oxidation of iron (II) monosulfide by hydrogen sulphide. Geochimica et Cosmochimica Acta, 64:26652672.CrossRefGoogle Scholar
Butterfield, N. J. 1994. Burgess Shale-type fossils from a Lower Cambrian shallow-shelf sequence in northwestern Canada. Nature, 369:477479.Google Scholar
Butterfield, N. J. 1995. Secular distribution of Burgess Shale-type preservation. Lethaia, 28:113.Google Scholar
Butterfield, N. J. 2003. Exceptional fossil preservation and the Cambrian Explosion. Integrative and Comparative Biology, 43:166177.Google Scholar
Cai, Y., Hua, H., Xiao, S., Schiffbauer, J. D., and Li, P. 2010. Biostratinomy of the Late Ediacaran pyritized Gaojiashan Lagerstätte from southern Shaanxi, South China: importance of event deposits. PALAIOS, 25:487506.Google Scholar
Cai, Y., Schiffbauer, J. D., Hua, H., and Xiao, S. 2011. Morphology and paleoecology of the late Ediacaran tubular fossil Conotubus hemiannulatus from the Gaojiashan Lagerstätte of southern Shaanxi Province, South China. Precambrian Research, 191:4657.Google Scholar
Cai, Y., Schiffbauer, J. D., Hua, H., and Xiao, S. 2012. Preservational modes in the Ediacaran Gaojiashan Lagerstätte: Pyritization, aluminosilicification, and carbonaceous compression. Palaeogeography, Palaeoclimatology, Palaeoecology, 326:109117.Google Scholar
Canfield, D. E. 1989. Reactive iron in marine sediments. Geochimica et Cosmochimica Acta, 53:619632.Google Scholar
Canfield, D. E. 1994. Factors influencing organic carbon preservation in marine sediments. Chemical Geology, 114:315329.Google Scholar
Canfield, D. E. 2001. Biogeochemistry of sulfur isotopes. Reviews in Mineralogy and Geochemistry, 43:607636.Google Scholar
Canfield, D. E., Jørgensen, B. B., Fossing, H., Glud, R., Gundersen, J., Ramsing, N. B., Thamdrup, B., Hansen, J. W., Nielsen, L. P., and Hall, P. O. 1993. Pathways of organic carbon oxidation in three continental margin sediments. Marine Geology, 113:2740.Google Scholar
Canfield, D. E., and Raiswell, R. 1991. Pyrite formation and fossil preservation, p. 337387. In Allison, P. A. and Briggs, D. E. G. (eds.), Taphonomy: Releasing the Data Locked in the Fossil Record. Plenum Press, New York.Google Scholar
Canfield, D. E., Raiswell, R., and Bottrell, S. 1992. The reactivity of sedimentary iron minerals towards sulfide. American Journal of Science, 292:659683.Google Scholar
Canfield, D. E., Thamdrup, B., and Fleischer, S. 1998. Isotope fractionation and sulfur metabolism by pure and enrichment cultures of elemental sulfur-disproportionating bacteria. Limnology and Oceanography, 43:253264.Google Scholar
Charbonnier, S. 2009. Le Lagerstätte de La Voulte: un environnement bathyal au Jurassique. Mémoires du Muséum National d'histoire Naturelle, 199:1272.Google Scholar
Charbonnier, S., Audo, D., Caze, B., and Biot, V. 2014. The La Voulte-sur-Rhône Lagerstätte (Middle Jurassic, France). Comptes Rendus Palevol, 13:369381.Google Scholar
Charbonnier, S., Vannier, J., Hantzpergue, P., and Gaillard, C. 2010. Ecological significance of the arthropod fauna from the Jurassic (Callovian) La Voulte Lagerstätte. Acta Palaeontologica Polonica, 55:111132.Google Scholar
Charbonnier, S., Vannier, J., Gaillard, C., Bourseau, J.-P., and Hantzpergue, P. 2007a. The La Voulte Lagerstätte (Callovian): evidence for a deep water setting from sponge and crinoid communities. Palaeogeography, Palaeoclimatology, Palaeoecology, 250:216236.Google Scholar
Charbonnier, S., Vannier, J., and Riou, B. 2007b. New sea spiders from the Jurassic La Voulte-sur-Rhône Lagerstätte. Proceedings of the Royal Society of London B-Biological Sciences, 274:25552561.Google Scholar
Chester, R. 2000. Marine Geochemistry. Blackwell Science, Oxford.Google Scholar
Cisne, J. L. 1973. Beecher's Trilobite Bed revisited: ecology of an Ordovician deepwater fauna. Postilla, 160:125.Google Scholar
Coleman, M. L., and Raiswell, R. 1993. Microbial mineralization of organic matter: mechanisms of self-organization and inferred rates of precipitation of diagenetic minerals. Philosophical Transactions of the Royal Society of London-A, 344:6987.Google Scholar
Davison, W. 1988. Interactions of iron, carbon and sulphur in marine and lacustrine sediments. Geological Society of London Special Publications, 40:131137.Google Scholar
Detmers, J., Brüchert, V., Habicht, K. S., and Kuever, J. 2001. Diversity of sulfur isotopes fractionations by sulfate-reducing prokaryotes. Applied and Environmental Microbiology, 67:888894.Google Scholar
Farrell, Ú. C., and Briggs, D. E. G. 2007. A pyritized polychaete from the Devonian of Ontario. Proceedings of the Royal Society of London B-Biology, 274:499504.Google Scholar
Farrell, Ú. C., Briggs, D. E. G. and Gaines, R. R. 2011. Paleoecology of the Olenid trilobite Triarthrus: new evidence from Beecher's Trilobite Bed and other sites of pyritization. PALAIOS, 26:730742.CrossRefGoogle Scholar
Farrell, Ú. C., Briggs, D. E. G., Hammarlund, E. U., Sperling, E. A., and Gaines, R. R. 2013. Paleoredox and pyritization of soft-bodied fossils in the Ordovician Frankfort Shale of New York. American Journal of Science, 313:452489.Google Scholar
Farrell, Ú. C., Martin, M. J., Hagadorn, J. W., Whiteley, T., and Briggs, D. E. G. 2009. Beyond Beecher's Trilobite Bed: widespread pyritization of soft tissues in the Late Ordovician Taconic foreland basin. Geology, 37:907910.Google Scholar
Fenchel, T., and Finlay, B. 2008. Oxygen and the spatial structure of microbial communities. Biological Reviews, 83:553569.CrossRefGoogle ScholarPubMed
Forchielli, A., Steiner, M., Kasbohm, J., Hu, S., and Keupp, H. 2014. Taphonomic traits of clay-hosted early Cambrian Burgess Shale-type fossil Lagerstätten in South China. Palaeogeography, Palaeoclimatology, Palaeoecology, 398:5985.Google Scholar
Froelich, P. N., Klinkhammer, G. P., Bender, M. L., Luedtke, N. A., Heath, G. R., Cullen, D., and Dauphin, P. 1979. Early oxidation of organic matter in pelagic sediments of the eastern equatorial Atlantic: suboxic diagenesis. Geochimica et Cosmochimica Acta, 43:10751090.CrossRefGoogle Scholar
Gabbott, S. E., Xian-Guang, H., Norry, M. J., and Siveter, D. J. 2004. Preservation of Early Cambrian animals of the Chengjiang biota. Geology, 32:901904.CrossRefGoogle Scholar
Gaines, R. R. 2014. Burgess Shale-type preservation and its distribution in space and time, p. 123146. In Laflamme, M., Schiffbauer, J. D., and Darroch, S. A. F. (eds.), Reading and Writing of the Fossil Record: Preservational Pathways to Exceptional Fossilization. The Paleontological Society Papers 20. Yale Press, New Haven, CT.Google Scholar
Gaines, R. R., Briggs, D. E. G., and Zhao, Y. 2008. Cambrian Burgess Shale-type deposits share a common mode of fossilization. Geology, 36:755758.Google Scholar
Gaines, R. R., Hammarlund, E. U., Hou, X., Qi, C., Gabbott, S. E., Zhao, Y., Peng, J., and Canfield, D. E. 2012. Mechanism for Burgess Shale-type preservation. Proceedings of the National Academy of Sciences of the United States of America, 109:51805184.Google Scholar
Gaines, R. R., Peters, S., Hammarlund, E., Briggs, D. E., Qi, C., Hou, X., Gabbott, S. E., and Canfield, D. E. The Early Phanerozoic “Taphonomic Window.” Geological Society of America Abstracts with Programs, 45(7):306.Google Scholar
García-Bellido, D. C., and Collins, D. H. 2006. A new study of Marrella splendens (Arthropoda, Marrellomorpha) from the Middle Cambrian Burgess Shale, British Columbia, Canada. Canadian Journal of Earth Science, 43:721742.Google Scholar
Gill, B. C., Lyons, T. W., and Saltzman, M. R. 2007. Parallel, high-resolution carbon and sulfur isotope records of the evolving Paleozoic marine sulfur reservoir. Palaeogeography, Palaeoclimatology, Palaeoecology, 256:156173.Google Scholar
Glass, A. 2006. Pyritized tube feet in a protasterid ophiuroid from the Upper Ordovician of Kentucky, U.S.A. Acta Palaeontologica Polonica, 51:171184.Google Scholar
Glass, A., and Blake, D. B. 2004. Preservation of tube feet in an ophiuroid (Echinodermata) from the Lower Devonian Hunsrück Slate of Germany and a redescription of Bundenbachia beneckei and Palaeophiomyxa grandis . Paläontologische Zeitschrift, 78:7395.Google Scholar
Goldhaber, M., and Kaplan, I. 1974. The sulfur cycle, p. 569655. In Goldberg, E. D. (ed.), The Sea, Volume 5: Marine Chemistry. Wiley-Interscience, New York.Google Scholar
Goldhaber, M. B. 2003. Sulfur-rich sediments, p. 257288. In Holland, H. D. and Turekian, K. K. (eds.), Treatise on Geochemistry, Volume 7. Elsevier.Google Scholar
Goldhaber, M. B., Aller, R. C., Cochran, J. K., Rosenfeld, J. K., Martens, C. S., and Berner, R. A. 1977. Sulfate reduction, diffusion, and bioturbation in Long Island Sound sediments: report of the FOAM group. American Journal of Science, 277:193237.Google Scholar
Grimes, S. T., Brock, F., Rickard, D., Davies, K. L., Edwards, D., Briggs, D. E. G., and Parkes, R. J. 2001. Understanding fossilization: experimental pyritization of plants. Geology, 29:123126.Google Scholar
Grimes, S. T., Davies, K. L., Butler, I. B., Brock, F., Edwards, D., Rickard, D., Briggs, D. E. G., and Parkes, R. J. 2002. Fossil plants from the Eocene London Clay: the use of pyrite textures to determine the mechanism of pyritization. Journal of the Geological Society, 159:493501.Google Scholar
Habicht, K. S., and Canfield, D. E. 1997. Sulfur isotope fractionation during bacterial sulfate reduction in organic-rich sediments. Geochimica et Cosmochimica Acta, 61:53515361.Google Scholar
Habicht, K. S., and Canfield, D. E. 2001. Isotope fractionation by sulfate-reducing natural populations and the isotopic composition of sulfide in marine sediments. Geology, 29:555558.Google Scholar
Haese, R. R. 2006. The biogeochemistry of iron, p. 241270. In Schulz, H. D. and Zabel, M. (eds.), Marine Geochemistry. Springer, Berlin.Google Scholar
Han, J., Shu, D., Zhang, Z., Liu, J., Zhang, X., and Yao, Y. 2006. Preliminary notes on soft-bodied fossil concentrations from the Early Cambrian Chengjiang deposits. Chinese Science Bulletin, 51:24822492.Google Scholar
Heimhofer, R., and Martill, D. 2007. The sedimentology and depositional environment of the Crato Formation, p. 4462. In Martill, D., Bechly, G., and Loveridge, R. F. (eds.), The Crato Fossil Beds of Brazil: Window Into An Ancient World. Cambridge University Press, Cambridge.Google Scholar
Heimhofer, U., Ariztegui, D., Lenniger, M., Hesselbo, S. P., Martill, D. M., and Rios-Netto, A. M. 2010. Deciphering the depositional environment of the laminated Crato fossil beds (Early Cretaceous, Araripe Basin, Northeastern Brazil). Sedimentology, 57:677694.Google Scholar
Helfferich, F. G., and Katchalsky, A. 1970. A simple model of interdiffusion with precipitation. The Journal of Physical Chemistry, 74:308314.Google Scholar
Hof, C. H., and Briggs, D. E. G. 1997. Decay and mineralization of mantis shrimps (Stomatopoda; Crustacea); a key to their fossil record. PALAIOS, 12:420438.Google Scholar
Horita, J., Zimmermann, H., and Holland, H. D. 2002. Chemical evolution of seawater during the Phanerozoic: Implications from the record of marine evaporites. Geochimica et Cosmochimica Acta, 66:37333756.Google Scholar
Hudson, J. 1982. Pyrite in ammonite-bearing shales from the Jurassic of England and Germany. Sedimentology, 29:639667.Google Scholar
Hunger, S., and Benning, L. G. 2007. Greigite: a true intermediate on the polysulfide pathway to pyrite. Geochemical Transactions, 8:1. doi:10.1186/1467-4866-8-1 Google Scholar
Hurtgen, M. T., Arthur, M. A., and Halverson, G. P. 2005. Neoproterozoic sulfur isotopes, the evolution of microbial sulfur species, and the burial efficiency of sulfide as sedimentary pyrite. Geology, 33:4144.Google Scholar
Jørgensen, B. 1977. Bacterial sulfate reduction within reduced microniches of oxidized marine sediments. Marine Biology, 41:717.Google Scholar
Kampschulte, A., and Strauss, H. 2004. The sulfur isotopic evolution of Phanerozoic seawater based on the analysis of structurally substituted sulfate in carbonates. Chemical Geology, 204:255286.CrossRefGoogle Scholar
Kühl, G., Bartels, C., Briggs, D. E. G., and Rust, J. 2012. Visions of a Vanished World. Yale University Press, New Haven.Google Scholar
Kühl, G., Poschmann, M., and Rust, J. 2013. A ten-legged sea spider (Arthropoda: Pycnogonida) from the Lower Devonian Hunsrück Slate (Germany). Geological Magazine, 150:556564.Google Scholar
Last, W. M., and Ginn, F. M. 2005. Saline systems of the Great Plains of western Canada: an overview of the limnogeology and paleolimnology. Saline Systems, 1:10. doi:10.1186/1746-1448-1-10 Google Scholar
Locatelli, E. R. 2014. The exceptional preservation of plant fossils: a review of taphonomic pathways and biases in the fossil record, p. 237257. In Laflamme, M., Schiffbauer, J. D., and Darroch, S. A. F. (eds.), Reading and Writing of the Fossil Record: Preservational Pathways to Exceptional Fossilization. The Paleontological Society Papers 20. Yale Press, New Haven, CT.Google Scholar
Love, L., and Amstutz, G. 1966. Review of microscopic pyrite from the Devonian Chattanooga Shale and Rammelsberg Banderz. Fortschritte der Mineralogie, 43:273309.Google Scholar
Luther, G. W. III. 1991. Pyrite synthesis via polysulfide compounds. Geochimica et Cosmochimica Acta, 55:28392849.Google Scholar
Lyons, T. W. 1997. Sulfur isotopic trends and pathways of iron sulfide formation in upper Holocene sediments of anoxic Black Sea. Geochimica et Cosmochimica Acta, 61:33673382.Google Scholar
Lyons, T. W., Anbar, A. D., Severmann, S., Scott, C., and Gill, B. C. 2009. Tracking euxinia in the ancient ocean: a multiproxy perspective and Proterozoic case study. Annual Review of Earth and Planetary Sciences, 37:507534.Google Scholar
Lyons, T. W., and Severmann, S. 2006. A critical look at iron paleoredox proxies: New insights from modern euxinic marine basins. Geochimica et Cosmochimica Acta, 70:56985722.Google Scholar
Marnette, E. C., van Breemen, N., Hordijk, K. A., and Cappenberg, T. E. 1993. Pyrite formation in two freshwater systems in the Netherlands. Geochimica et Cosmochimica Acta, 57:41654177.Google Scholar
Martill, D. M., Bechly, G., and Loveridge, R. F. 2007. The Crato Fossil Beds of Brazil: Window Into an Ancient World. Cambridge University Press, Cambridge, UK.Google Scholar
McDermott Buck, M. 2004. Beecher's Trilobite Bed in the 21st Century: MicroCT analysis of an Ordovician Lagerstätte. Unpublished B.A. Thesis, Amherst College, 129 p.Google Scholar
McKay, J. L., and Longstaffe, F. J. 2003. Sulphur isotope geochemistry of pyrite from the Upper Cretaceous Marshybank Formation, Western Interior Basin. Sedimentary Geology, 157:175195.Google Scholar
Menon, F., and Martill, D. 2007. Taphonomy and preservation of Crato Formation arthropods, p. 7996. In Martill, D., Bechly, G., and Loveridge, R. F. (eds.), The Crato Fossil Beds of Brazil: Window Into An Ancient World. Cambridge University Press, Cambridge, UK.Google Scholar
Moore, R. A., and Lieberman, B. S. 2009. Preservation of early and Middle Cambrian soft-bodied arthropods from the Pioche Shale, Nevada, USA. Palaeogeography, Palaeoclimatology, Palaeoecology, 277:5762.Google Scholar
Morse, J. W., and Wang, Q. 1997. Pyrite formation under conditions approximating those in anoxic sediments: II. Influence of precursor iron minerals and organic matter. Marine Chemistry, 57:187193.Google Scholar
Mucci, A., and Edenborn, H. M. 1992. Influence of an organic-poor landslide deposit on the early diagenesis of iron and manganese in a coastal marine sediment. Geochimica et Cosmochimica Acta, 56:39093921.CrossRefGoogle Scholar
Ohfuji, H., and Rickard, D. 2005. Experimental syntheses of framboids—a review. Earth-Science Reviews, 71:147170.Google Scholar
Orr, P. J., Briggs, D. E. G., and Kearns, S. L. 1998. Cambrian Burgess Shale animals replicated in clay minerals. Science, 281:11731175.Google Scholar
Picard, M. D. 1971. Classification of fine-grained sedimentary rocks. Journal of Sedimentary Research, 41:179195.Google Scholar
Poulton, S. W., and Canfield, D. E. 2005. Development of a sequential extraction procedure for iron: implications for iron partitioning in continentally derived particulates. Chemical Geology, 214:209221.Google Scholar
Poulton, S. W., Fralick, P. W., and Canfield, D. E. 2004. The transition to a sulfidic ocean ∼1.84 billion years ago. Nature, 431:173177.Google Scholar
Poulton, S. W., Fralick, P. W., and Canfield, D. E. 2010. Spatial variability in oceanic redox structure 1.8 billion years ago. Nature Geoscience, 3:486490.Google Scholar
Poulton, S. W., Krom, M. D., and Raiswell, R. 2004. A revised scheme for the reactivity of iron (oxyhydr)oxide minerals towards dissolved sulfide. Geochimica et Cosmochimica Acta, 68:37033715.Google Scholar
Poulton, S. W., and Raiswell, R. 2002. The low-temperature geochemical cycle of iron: from continental fluxes to marine sediment deposition. American Journal of Science, 302:774805.Google Scholar
Raiswell, R. 1993. Kinetic controls on depth variations in localised pyrite formation. Chemical Geology, 107:467469.Google Scholar
Raiswell, R. 1997. A geochemical framework for the application of stable sulphur isotopes to fossil pyritization. Journal of the Geological Society, 154:343356.Google Scholar
Raiswell, R., and Berner, R. A. 1987. Organic carbon losses during burial and thermal maturation of normal marine shales. Geology, 15:853856.Google Scholar
Raiswell, R., Buckley, F., Berner, R. A., and Anderson, T. F. 1988. Degree of pyritization of iron as a paleoenvironmental indicator of bottom-water oxygenation. Journal of Sedimentary Petrology, 58:812819.Google Scholar
Raiswell, R., and Canfield, D. E. 1998. Sources of iron for pyrite formation in marine sediments. American Journal of Science, 298:219245.Google Scholar
Raiswell, R., Canfield, D. E., and Berner, R. A. 1994. A comparison of iron extraction methods for the determination of degree of pyritisation and the recognition of iron-limited pyrite formation. Chemical Geology, 111:101110.Google Scholar
Raiswell, R., Newton, R., Bottrell, S. H., Coburn, P. M., Briggs, D. E. G., Bond, D. P., and Poulton, S. W. 2008. Turbidite depositional influences on the diagenesis of Beecher's Trilobite Bed and the Hunsrück Slate; sites of soft tissue pyritization. American Journal of Science, 308:105129.Google Scholar
Raiswell, R., Whaler, K., Dean, S., Coleman, M. L., and Briggs, D. E. G. 1993. A simple three-dimensional model of diffusion-with-precipitation applied to localised pyrite formation in framboids, fossils and detrital iron minerals. Marine Geology, 113:89100.Google Scholar
Rickard, D. 1997. Kinetics of pyrite formation by the H2S oxidation of iron (II) monosulfide in aqueous solutions between 25 and 125°C: The rate equation. Geochimica et Cosmochimica Acta, 61:115134.Google Scholar
Rickard, D., and Luther, G. W. III. 1997. Kinetics of pyrite formation by the H2S oxidation of iron (II) monosulfide in aqueous solutions between 25 and 125° C: The mechanism. Geochimica et Cosmochimica Acta, 61:135147.Google Scholar
Rickard, D., and Luther, G. W. III. 2007. Chemistry of iron sulfides. Chemical Reviews, 107:514562.Google Scholar
Rickard, D., Schoonen, M. A., and Luther, G. III. 1995. Chemistry of iron sulfides in sedimentary environments, p. 168193. In Vairavamurthy, M. A. and Schoonen, M. (eds.), Geochemical Transformations of Sedimentary Sulfur. ACS Symposium Series 612. American Chemical Society Publications, Washington, D. C.Google Scholar
Rickard, D. T. 1975. Kinetics and mechanism of pyrite formation at low temperatures. American Journal of Science, 275:636652.Google Scholar
Ries, J. B., Fike, D. A., Pratt, L. M., Lyons, T. W., and Grotzinger, J. P. 2009. Superheavy pyrite (δ34Spyr > δ34Scas) in the terminal Proterozoic Nama Group, southern Namibia: a consequence of low seawater sulfate at the dawn of animal life. Geology, 37:743746.CrossRefGoogle Scholar
Römer, F. 1900. Neue asteriden und crinoiden aus Devonischem Dachschiefer von Bundenbach bei Birkenfeld. Palaeontographica, 9:143152.Google Scholar
Sageman, B. B., and Lyons, T. W. 2004. Geochemistry of fine-grained sediments and sedimentary rocks, p. 115158. In MacKenzie, F. T., Holland, H. D., and Turekian, K. K. (eds.), Treatise on Geochemistry Volume 7. Elsevier, Amsterdam.Google 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
Sansom, R. S. 2014. Experimental decay of soft tissues, p. 217236. In Laflamme, M., Schiffbauer, J. D., and Darroch, S. A. F. (eds.), Reading and Writing of the Fossil Record: Preservational Pathways to Exceptional Fossilization. The Paleontological Society Papers 20. Yale Press, New Haven, CT.Google Scholar
Schieber, J. 2002. The role of an organic slime matrix in the formation of pyritized burrow trails and pyrite concretions. PALAIOS, 17:104109.Google Scholar
Schoonen, M., and Barnes, H. 1991a. Reactions forming pyrite and marcasite from solution: I. Nucleation of FeS2 below 100° C. Geochimica et Cosmochimica Acta, 55:14951504.Google Scholar
Schoonen, M., and Barnes, H. 1991b. Reactions forming pyrite and marcasite from solution: II. Via FeS precursors below 100°C. Geochimica et Cosmochimica Acta, 55:15051514.Google Scholar
Schoonen, M. A. 2004. Mechanisms of sedimentary pyrite formation, p. 117134. In Amend, J. P., Edwards, K. J., and Lyons, T. W. (eds.), Sulfur Biogeochemistry—Past and Present. Geological Society of America Special Publication 379, Geological Society of America, Boulder, Colorado.Google Scholar
Schulz, H. D. 2006. Conceptual models and computer models, p. 513547. In Schulz, H. and Zabel, M. (eds.), Marine Geochemistry. Springer, Berlin.Google Scholar
Schwimmer, D. R., and Montante, W. M. 2007. Exceptional fossil preservation in the Conasauga Formation, Cambrian, northwestern Georgia, USA. PALAIOS, 22:360372.Google Scholar
Seilacher, A. 1970. Begriff und bedeutung der Fossil-Lagerstätten. Neues Jahrbuch für Geologie und Paläontologie Abhandlungen, p. 339.Google Scholar
Seilacher, A., Reif, W.-E., and Westphal, F. 1985. Sedimentological, ecological and temporal patterns of fossil Lagerstätten. Philosophical Transactions of the Royal Society of London B-Biological Sciences, 311:523.Google Scholar
Siveter, D. J., Tanaka, G., Farrell, Ú. C., Martin, M. J., Siveter, D. J., and Briggs, D. E. G. 2014. Exceptionally preserved 450-million-year-old Ordovician ostracods with brood care. Current Biology, 24:801806.Google Scholar
Skinner, E. S. 2005. Taphonomy and depositional circumstances of exceptionally preserved fossils from the Kinzers Formation (Cambrian), southeastern Pennsylvania. Palaeogeography, Palaeoclimatology, Palaeoecology, 220:167192.Google Scholar
Sochaczewski, L., Stockdale, A., Davison, W., Tych, W., and Zhang, H. 2008. A three-dimensional reactive transport model for sediments, incorporating microniches. Environmental Chemistry, 5:218225.Google Scholar
Stanley, G. D. 1983. The first fossil ctenophore from the Lower Devonian of West Germany. Nature, 303:518520.Google Scholar
Stanley, G. D., and Stürmer, W. 1987. A new fossil ctenophore discovered by X-rays. Nature, 328:6163.Google Scholar
Stockdale, A., Davison, W., and Zhang, H. 2010. Formation of iron sulfide at faecal pellets and other microniches within suboxic surface sediment. Geochimica et Cosmochimica Acta, 74:26652676.Google Scholar
Stürmer, W. 1970. Soft parts of cephalopods and trilobites: some surprising results of X-ray examinations of Devonian slates. Science, 170:13001302.Google Scholar
Stürmer, W., and Bergström, J. 1973. New discoveries on trilobites by X-rays. Paläontologische Zeitschrift, 47:104141.Google Scholar
Suits, N. S., and Wilkin, R. T. 1998. Pyrite formation in the water column and sediments of a meromictic lake. Geology, 26:10991102.Google Scholar
Sutcliffe, O. E. 1997. An ophiuroid trackway from the Lower Devonian Hunsrück Slate, Germany. Lethaia, 30:3339.Google Scholar
Sutcliffe, O. E., Briggs, D. E. G., and Bartels, C. 1999. Ichnological evidence for the environmental setting of the Fossil-Lagerstätten in the Devonian Hunsrück Slate, Germany. Geology, 27:275278.2.3.CO;2>CrossRefGoogle Scholar
Sutcliffe, O. E., Tibbs, S., and Briggs, D. E. G. 2002. Sedimentology and environmental interpretation of the fine-grained turbidites in the Kaub Formation of the Hunsrück Slate: analysis of a section excavated for Project Nahecaris. Metalla (Bochum), 9:89104.Google Scholar
Tanaka, G., Hou, X., Ma, X., Edgecombe, G. D., and Strausfeld, N. J. 2013. Chelicerate neural ground pattern in a Cambrian great appendage arthropod. Nature, 502:364367.Google Scholar
Van Cappellen, P., and Wang, Y. 1996. Cycling of iron and manganese in surface sediments; a general theory for the coupled transport and reaction of carbon, oxygen, nitrogen, sulfur, iron, and manganese. American Journal of Science, 296:197243.Google Scholar
Van Der Weijden, C. H. 1992. Early diagenesis and marine pore water, p. 13134. In Wolf, K. H. and Chilingarian, G. V. (eds.), Diagenesis, III. Developments in Sedimentology 47, Elsevier.Google Scholar
Van Roy, P., Orr, P. J., Botting, J. P., Muir, L. A., Vinther, J., Lefebvre, B., El Hariri, K., and Briggs, D. E. G. 2010. Ordovician faunas of Burgess Shale type. Nature, 465:215218.Google Scholar
Vinther, J., Van Roy, P., and Briggs, D. E. 2008. Machaeridians are Palaeozoic armoured annelids. Nature, 451:185188.Google Scholar
Wang, B., Zhao, F., Zhang, H., Fang, Y., and Zheng, D. 2012. Widespread pyritization of insects in the Early Cretaceous Jehol Biota. PALAIOS, 27:707711.Google Scholar
Wang, Q., and Morse, J. W. 1996. Pyrite formation under conditions approximating those in anoxic sediments I. Pathway and morphology. Marine Chemistry, 52:99121.Google Scholar
Weary, D. J., Ryder, R. T., and Nyahay, R. 2000. Thermal Maturity Patterns (CAI and %R0) in the Ordovician and Devonian Rocks of the Appalachian Basin in New York State. United States Geological Survey Open-file report 00-494.Google Scholar
Westrich, J. T., and Berner, R. A. 1984. The role of sedimentary organic matter in bacterial sulfate reduction: The G model tested. Limnology and Oceanography, 29:236249.Google Scholar
Whittington, H. B., and Almond, J. E. 1987. Appendages and habits of the Upper Ordovician trilobite Triarthrus eatoni . Philosophical Transactions of the Royal Society of London B-Biological Sciences, 317:146.Google Scholar
Wilby, P. R., and Briggs, D. E. G. 1997. Taxonomic trends in the resolution of detail preserved in fossil phosphatized soft tissues. Geobios, 30:493502.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
Wilkin, R., and Barnes, H. 1996. Pyrite formation by reactions of iron monosulfides with dissolved inorganic and organic sulfur species. Geochimica et Cosmochimica Acta, 60:41674179.CrossRefGoogle Scholar
Wilkin, R. T., Barnes, H. L., and Brantley, S. L. 1996. The size distribution of framboidal pyrite in modern sediments: An indicator of redox conditions. Geochimica et Cosmochimica Acta, 60:38973912.Google Scholar
Williams, M., Siveter, D. J., Ashworth, A. C., Wilby, P. R., Horne, D. J., Lewis, A. R., and Marchant, D. R. 2008. Exceptionally preserved lacustrine ostracods from the Middle Miocene of Antarctica: implications for high-latitiude paleoenvioronment at 77° South. Proceedings of the Royal Society of London B-Biological Sciences, 275:24492454.Google Scholar
Xiao, S., Yuan, X., Steiner, M., and Knoll, A. H. 2002. Macroscopic carbonaceous compressions in a terminal Proterozoic shale: a systematic reassessment of the Miaohe biota, South China. Journal of Paleontology, 76:347376.Google Scholar
Yao, W., and Millero, F. J. 1996. Oxidation of hydrogen sulfide by hydrous Fe (III) oxides in seawater. Marine Chemistry, 52:116.Google Scholar
Zhu, M., Babcock, L. E., and Steiner, M. 2005. Fossilization modes in the Chengjiang Lagerstätte (Cambrian of China): testing the roles of organic preservation and diagenetic alteration in exceptional preservation. Palaeogeography, Palaeoclimatology, Palaeoecology, 220:3146.Google Scholar