Hostname: page-component-7479d7b7d-wxhwt Total loading time: 0 Render date: 2024-07-13T00:12:39.692Z Has data issue: false hasContentIssue false

The Exceptional Preservation of Plant Fossils: A Review of Taphonomic Pathways and Biases in the Fossil Record

Published online by Cambridge University Press:  21 July 2017

Emma R. Locatelli*
Affiliation:
Department of Geology and Geophysics, Yale University, PO Box 208109, New Haven, CT 06520 USA
Get access

Abstract

The exceptional preservation of plant fossils falls into two categories: whole plant preservation and anatomical detail. Whole plant preservation is controlled primarily by transport and event preservation (e.g., ash falls), whereas anatomical preservation can occur through one of several taphonomic pathways: compression-impression, silicification, coal-ball formation, pyritization, and charcoalification. This review focuses on these taphonomic pathways, highlighting important factors and controls on the exceptional preservation of plants. Special emphasis is given to data garnered from experimental and actualistic approaches.

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

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
Alvin, K. 1974. Leaf anatomy of Weichselia based on fusainized material. Palaeontology, 17:587598.Google Scholar
Anderson, T. F., Brownlee, M. E., and Phillips, T. L. 1980. A stable isotope study on the origin of permineralized peat zones in the Herrin Coal. The Journal of Geology, 88:713722.Google Scholar
Andrews, H. N. 1951. American coal-ball floras. The Botanical Review, 17:431469.Google Scholar
Baghai, N. L., and Jorstad, R. B. 1995. Paleontology, paleoclimatology and paleoecology of the late middle Miocene Musselshell Creek flora, Clearwater County Idaho. A preliminary study of a new fossil flora. PALAIOS, 10:424436.Google Scholar
Ballhaus, C., Gee, C. T., Bockrath, C., Greef, K., Mansfeldt, T., and Rhede, D. 2012. The silicification of trees in volcanic ash—an experimental study. Geochimica et Cosmochimica Acta, 84:6274.Google Scholar
Basinger, J. F., and Rothwell, G. 1977. Anatomically preserved plants from the middle Eocene (Allenby Formation) of British Columbia. Canadian Journal of Botany, 55:19841990.Google Scholar
Beck, C. B. 1955. A technique for obtaining polished surfaces of sections of pyritized plant fossils. Bulletin of the Torrey Botanical Club, 82:286291.Google Scholar
Beck, C. B. 1960a. Connection between Archaeopteris and Callixylon . Science, 131:15241525.Google Scholar
Beck, C. B. 1960b. The identity of Archaeopteris and Callixylon . Brittonia, 12:351368.Google Scholar
Beerling, D. J., and Chaloner, W. G. 1993. The impact of atmospheric CO2 and temperature changes on stomatal density: observation from Quercus robur lammas leaves. Annals of Botany, 71:231235.Google Scholar
Beerling, D., and Royer, D. 2002. Fossil plants as indicators of the Phanerozoic global carbon cycle. Annual Review of Earth and Planetary Sciences, 30:527556.Google Scholar
Berner, R. A. 1984. Sedimentary pyrite formation: an update. Geochimica et Cosmochimica Acta, 48:605615.Google Scholar
Blanchette, R. A., Nilsson, T., Daniel, G., and Abad, A. 1990. Biological degradation of wood, p. 141174. In Rowell, R. M. and Borbour, R. J. (eds.), Archaeological Wood: Properties, Chemistry and Preservation: Advances in Chemistry. American Chemical Society, Washington, D. C.Google Scholar
Bomfleur, B., McLoughlin, S., and Vajda, V. 2014. Fossilized nuclei and chromosomes reveal 180 million years of genomic stasis in royal ferns. Science, 343:13761377.Google Scholar
Briggs, D. E. G. 1999. Molecular taphonomy of animal and plant cuticles: selective preservation and diagenesis. Philosophical Transactions of the Royal Society of London B-Biological Sciences, 354:717.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, 31:275.Google Scholar
Briggs, D. E. G., Evershed, R. P., and Lockheart, M. J. 2000. The biomolecular paleontology of continental fossils, p. 169193. In Erwin, D. H. and Wing, S. L. (eds.), Deep Time: Paleobiology's Perspective. Paleobiology, 26 (supplement to no. 4).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
Buurman, P. 1972. Mineralization of fossil wood. Scripta Geologica, 12:143.Google Scholar
Canfield, D. E. 1989. Reactive iron in marine sediments. Geochimica et Cosmochimica Acta, 53:619632.Google Scholar
Canfield, D. E., Raiswell, R., and Bottrell, S. H. 1989. Sulfate reduction and oxic respiration in marine sediments: implications for organic carbon preservation in euxinic environments. Deep Sea Research Part A. Oceanographic Research Papers, 36:121138.Google Scholar
Canfield, D. E., Raiswell, R., and Bottrell, S. H. 1992. The reactivity of sedimentary iron minerals toward sulfide. American Journal of Science, 292:659683.Google Scholar
Cecil, C. B., Stanton, R. W., Neuzil, S. G., Dulong, F. T., Ruppert, L. F., and Pierce, B. S. 1985. Paleoclimate controls on late Paleozoic sedimentation and peat formation in the central Appalachian Basin (USA). International Journal of Coal Geology, 5:195230.Google Scholar
Chaloner, W. G., and Collinson, M. E. 1975. Application of SEM to a sigillarian impression fossil. Review of Paleobotany and Palynology, 20:85101.Google Scholar
Chaloner, W. G., Hill, A., and Rogerson, E. C. W. 1978. Early Devonian plant fossils from a southern England borehole. Palaeontology, 21:693707.Google Scholar
Chaney, D. S., Mamay, S. H., DiMichele, W., and Kerp, H. 2009. Auritifolia gen. nov., probable seed plant foliage with comioid affinities from the early Permian of Texas, U.S.A. International Journal of Plant Sciences, 170:247266.Google Scholar
Channing, A., and Edwards, D. 2004. Experimental taphonomy: silicification of plants in Yellowstone hot-spring environments. Transactions of the Royal Society of Edinburgh: Earth Sciences, 94:503521.Google Scholar
Channing, A., and Edwards, D. 2009. Silicification of higher plants in geothermally influenced wetlands: Yellowstone as a Lower Devonian Rhynie analog. PALAIOS, 24:505521.Google Scholar
Cleal, C. J., and Shute, C. H. 2007. The effect of drying on epidermal cell parameters preserved on plant cuticles. Acta Palaeobotanica Krakow, 47:315326.Google Scholar
Collinson, M. E. 2011. Molecular taphonomy of plant organic skeletons, p. 223247. In Allison, P. A. and Bottjer, D. J. (eds.), Taphonomy: Process and Bias through Time. Topics in Geobiology 32, Springer, New York.Google Scholar
Conway Morris, S. 1986. The community structure of the Middle Cambrian phyllopod bed (Burgess Shale). Palaeontology, 29:423467.Google Scholar
Crepet, W. L., Nixon, K. C., and Gandolfo, M. A. 2004. Fossil evidence and phylogeny: the age of major angiosperm clades based on mesofossil and macrofossil evidence from Cretaceous deposits. American Journal of Botany, 91:16661682.Google Scholar
Darroch, S. A. F., Laflamme, M., Schiffbauer, J. D., and Briggs, D. E. G. 2012. Experimental formation of a microbial death mask. PALAIOS, 27:293303.Google Scholar
DeMaris, P. J. 2000. Formation and distribution of coal balls in the Herrin Coal (Pennsylvanian), Franklin County, Illinois Basin, USA. Journal of the Geological Society, 157:221228.Google Scholar
DeMaris, P. J., Bauer, R., Cahill, R., and Damberger, H. 1983. Geologic investigation of roof and floor strata: longwall demonstration, Old Ben Mine No. 24. Prediction of coal balls in the Herrin Coal. Final technical report: Part 2. [Mineralized peat balls]. Illinois State Geological Survey, Urbana, Illinois.Google Scholar
Dettmann, M. E., Clifford, H. T., and Peters, M. 2009. Lovellea wintonensis gen. et sp. nov.—Early Cretaceous (late Albian), anatomically preserved, angiospermous flowers and fruits from the Winton Formation, western Queensland, Australia. Cretaceous Research, 30:339355.Google Scholar
DiMichele, W. A., and Phillips, T. L. 1994. Paleobotanical and paleoecological constraints on models of peat formation in the late Carboniferous of Euramerica. Palaeogeography, Palaeoclimatology, Palaeoecology, 106:3990.Google Scholar
Doria, G., Royer, D. L., Wolfe, P. A., Fox, A., Westgate, J. A., and Beerling, D. J. 2011. Declining atmospheric CO2 during the late Middle Eocene climate transition. American Journal of Science, 311:6375.Google Scholar
Drum, R. W. 1968. Silicification of Betula woody tissue in vitro. Science, 161:175176.Google Scholar
Dunn, K. A., McLean, R. J. C., Upchurch, G. R., and Folk, R. L. 1997. Enhancement of leaf fossilization potential by bacterial biofilms. Geology, 25:11191122.Google Scholar
Elrick, S. D., and Nelson, W. 2010. Facies relationships of the middle Pennsylvanian Springfield Coal and Dykersberg Shale: constraints on sedimentation, development of coal splits and climate change during transgression. Geological Society of America Abstracts with Programs, 42(2):51.Google Scholar
Farrell, U. C. 2014. Pyritization of soft tissues in the fossil record: an overview, p. 3557. 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
Ferguson, D. K. 2005. Plant taphonomy: ruminations on the past, the present, and the future. PALAIOS, 20:418428.Google Scholar
Fioretto, A., Di Nardo, C., Papa, S., and Fuggi, A. 2005. Lignin and cellulose degradation and nitrogen dynamics during decomposition of three leaf litter species in a Mediterranean ecosystem. Soil Biology and Biochemistry, 37:10831091.Google Scholar
Friis, E. M., Crane, P. R., and Pedersen, K. R. 2011. Early Flowers and Angiosperm Evolution. Cambridge University Press.Google Scholar
Friis, E. M., and Skarby, A. 1981. Structurally preserved angiosperm flowers from the Upper Cretaceous of southern Sweden. Nature, 291:484486.Google Scholar
Gastaldo, R. 1988. A conspectus of phytotaphonomy, p. 1428. In DiMichele, W. A. and Wing, S. L. (eds.), Methods and Applications of Plant Paleoecology: Notes for a Short Course. Paleontological Society Special Publication 3, University of Tennessee, Knoxville.Google Scholar
Gee, C. T., and Gastaldo, R. A. 2005. Sticks and mud, fruits and nuts, leaves and climate: plant taphonomy comes of age. PALAIOS, 20:415417.Google Scholar
Gensel, P. G. 2008. The earliest land plants. Annual Review of Ecology, Evolution, and Systematics, 39:459477.Google Scholar
Giblin, A. E., and Howarth, R. W. 1984. Porewater evidence for a dynamic sedimentary iron cycle in salt marshes. Limnology and Oceanography, 29:4763.Google Scholar
Giles, M., Indrelid, S., and James, D. 1998. Compaction—the great unknown in basin modelling. Geological Society of London Special Publications, 141:1543.Google Scholar
Götze, J., Möckel, R., Langhof, N., Hengst, M., and Klinger, M. 2008. Silicification of wood in the laboratory: Ceramics Silikáty, 52:268277.Google Scholar
Gray, J., Chaloner, W. G., and Westoll, T. S. 1985. The microfossil record of early land plants: advances in understanding of early terrestrialization, 1970–1984 [and Discussion]. Philosophical Transactions of the Royal Society of London B-Biological Sciences, 309:167195.Google Scholar
Greenwood, D. R. 1991. The taphonomy of plant macrofossils, p. 141169. In Donovan, S. K. (ed.), The Processes of Fossilization. Columbia University Press, New York.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
Gupta, N. S., Collinson, M. E., Briggs, D. E. G., Evershed, R. P., and Pancost, R. D. 2006. Reinvestigation of the occurrence of cutan in plants: implications for the leaf fossil record. Paleobiology, 32:432449.Google Scholar
Gupta, N. S., Michels, R., Briggs, D. E. G., Collinson, M. E., Evershed, R. P., and Pancost, R. D. 2007a. Experimental evidence for the formation of geomacromolecules from plant leaf lipids. Organic Geochemistry, 38:2836.Google Scholar
Gupta, N. S., Yang, H., and Briggs, D. E. G. 2007b. Molecular taphonomy of Metasequoia . Bulletin of the Peabody Museum of Natural History, 48:329338.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
Hesse, R. 1989. Silica diagenesis: origin of inorganic and replacement cherts. Earth-Science Reviews, 26:253284.Google Scholar
Huang, Y., Shuman, B., Wang, Y., and Webb, T. 2004. Hydrogen isotope ratios of individual lipids in lake sediments as novel tracers of climatic and environmental change: a surface sediment test. Journal of Paleolimnology, 31:363375.Google Scholar
Jefferson, T. H. 1987. The preservation of conifer wood: examples from the Lower Cretaceous of Antarctica. Palaeontology, 30:233249.Google Scholar
Jones, T. P., and Chaloner, W. G. 1991. Fossil charcoal, its recognition and palaeoatmospheric significance. Palaeogeography, Palaeoclimatology, Palaeoecology, 97:3950.Google Scholar
Jones, T. P., Scott, A. C., and Cope, M. 1991. Reflectance measurements and the temperature of formation of modern charcoals and implications for studies of fusain. Bulletin de la Société Géologique de France, 162:193200.Google Scholar
Kaplan, I., Emery, K., and Rittenbebg, S. 1963. The distribution and isotopic abundance of sulphur in recent marine sediments off southern California. Geochimica et Cosmochimica Acta, 27:297331.Google Scholar
Keegstra, K. 2010. Plant cell walls. Plant Physiology, 154:483486.Google Scholar
Kelleher, B. P., Simpson, M. J., and Simpson, A. J. 2006. Assessing the fate and transformation of plant residues in the terrestrial environment using HR-MAS NMR spectroscopy. Geochimica et Cosmochimica Acta, 70:40804094.Google Scholar
Kenrick, P., and Crane, P. R. 1991. Water-conducting cells in early fossil land plants: implications for the early evolution of tracheophytes. Botanical Gazette, 152:335356.Google Scholar
Kenrick, P., and Edwards, D. 1988. The anatomy of Lower Devonian Gosslingia breconensis Heard based on pyritized axes, with some comments on the permineralization process. Botanical Journal of the Linnean Society, 97:95123.Google Scholar
Kidston, R., and Lang, W. 1917. XXIV.—On Old Red Sandstone plants showing structure, from the Rhynie Chert Bed, Aberdeenshire Part I. Rhynia Gwynne-Vaughani, Kidston and Lang. Transactions of the Royal Society of Edinburgh, 51:761784.Google Scholar
Kidston, R., and Lang, W. 1920a. XXVI.—On Old Red Sandstone plants showing structure, from the Rhynie Chert Bed, Aberdeenshire. Part III. Asteroxylon mackiei, Kidston and Lang. Transactions of the Royal Society of Edinburgh, 52:643680.Google Scholar
Kidston, R., and Lang, W. 1920b, XXIV—On Old Red Sandstone plants showing structure, from the Rhynie Chert Bed, Aberdeenshire. Part II. Additional Notes on Rhynia Gwynne-Vaughani, Kidston and Lang; with Descriptions of Rhynia major, n. sp., and Hornea lignieri, n. g., n. sp: Transactions of the Royal Society of Edinburgh, 52:603627.Google Scholar
Kidston, R., and Lang, W. 1921a. XXXII.—On Old Red Sandstone plants showing structure, from the Rhynie Chert Bed, Aberdeenshire. Part IV Restorations of the vascular cryptogams, and discussion of their bearing on the general morphology of the Pteridophyta and the origin of the organisation of land-plants. Transactions of the Royal Society of Edinburgh, 52:831854.Google Scholar
Kidston, R., and Lang, W. 1921b. XXXIII.—On Old Red Sandstone plants showing structure, from the Rhynie Chert Bed, Aberdeenshire. Part V The Thallophyta occurring in the peat-bed; the succession of the plants throughout a vertical section of the bed, and the conditions of accumulation and preservation of the deposit. Transactions of the Royal Society of Edinburgh, 52:855902.Google Scholar
Knoll, A. H. 1985. Exceptional preservation of photosynthetic organisms in silicified carbonates and silicified peats. Philosophical Transactions of the Royal Society of London B-Biological Sciences, 311:111122.Google Scholar
Kuder, T., and Kruge, M. A. 1998. Preservation of biomolecules in sub-fossil plants from raised peat bogs—a potential paleoenvironmental proxy. Organic Geochemistry, 29:13551368.Google Scholar
Leo, R. F., and Barghoorn, E. S. 1976. Silicification of Wood. Harvard University Botanical Museum Leaflets, 25:147.Google Scholar
Locatelli, E. R. 2013. The exceptional preservation of leaves in iron-rich sediments from Oceania. Geological Society of America Abstracts with Programs, 45(7):455.Google Scholar
Logan, G. A., Smiley, C. J., and Eglinton, G. 1995. Preservation of fossil leaf waxes in association with their source tissues, Clarkia, northern Idaho, USA. Geochimica et Cosmochimica Acta, 59:751763.Google Scholar
Lomax, J. 1902. On the occurrence of the nodular concretions (coal balls) in the Lower Coal Measures. Annals of Botany, 4:603604.Google Scholar
Lupia, R. 1995. Paleobotanical data from fossil charcoal: an actualistic study of seed plant reproductive structures. PALAIOS, 10:465465.Google Scholar
Magan, N. 1997. Fungi in extreme environments, p. 99114. In Wicklow, D. T. and Söderström, B. (eds.), The Mycota IV Environmental and Microbial Relationships. Springer-Verlag, Berlin.Google Scholar
Mamay, S. H., and Yochelson, E. L. 1962. Occurrence and Significance of Marine Animal Remains in American Coal Balls. United States Geological Survey Professional Paper 354-I. U. S. Government Printing Office, Washington, DC.Google Scholar
Matysová, P., Rössler, R., Götze, J., Leichmann, J., Forbes, G., Taylor, E. L., Sakala, J., and Grygar, T. 2010. Alluvial and volcanic pathways to silicified plant stems (upper Carboniferous–Triassic) and their taphonomic and palaeoenvironmental meaning. Palaeogeography, Palaeoclimatology, Palaeoecology, 292:127143.Google Scholar
McCoy, V 2014. Concretions as agents of soft-tissue preservation: a review, p. 147161. 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
Nickrent, D. L., Parkinson, C. L., Palmer, J. D., and Duff, R. J. 2000. Multigene phylogeny of land plants with special reference to bryophytes and the earliest land plants. Molecular Biology and Evolution, 17:18851895.Google Scholar
Niklas, K. J. 1978. Morphometric relationships and rates of evolution among Paleozoic vascular plants, p. 509543. In Hecht, M. K., Steere, W. C., and Wallace, B. (eds.), Evolutionary Biology: Volume 11. Springer, New York.Google Scholar
Niklas, K. J. 1984. Size-related changes in the primary xylem anatomy of some early tracheophytes. Paleobiology, 10:487506.Google Scholar
Niklas, K. J., and Brown, R. M. Jr. 1981. Ultrastructural and paleobiochemical correlations among fossil leaf tissues from the St. Maries River (Clarkia) area, northern Idaho, USA. American Journal of Botany, 68:332341.Google Scholar
Nip, M., Tegelaar, E., Brinkhuis, H., de Leeuw, J., Schenck, P., and Holloway, P. 1986. Analysis of modern and fossil plant cuticles by Curie point Py-GC and Curie point Py-GC-MS: recognition of a new, highly aliphatic and resistant biopolymer. Organic Geochemistry, 10:769778.Google Scholar
Nishiyama, T., Wolf, P. G., Kugita, M., Sinclair, R. B., Sugita, M., Sugiura, C., Wakasugi, T., Yamada, K., Yoshinaga, K., and Yamaguchi, K. 2004. Chloroplast phylogeny indicates that bryophytes are monophyletic. Molecular Biology and Evolution, 21:18131819.Google Scholar
Phillips, T. L., and Peppers, R. A. 1984. Changing patterns of Pennsylvanian coal-swamp vegetation and implications of climatic control on coal occurrence. International Journal of Coal Geology, 3:205255.Google Scholar
Phillips, T. L., Peppers, R. A., and DiMichele, W. A. 1985. Stratigraphic and interregional changes in Pennsylvanian coal-swamp vegetation: environmental inferences. International Journal of Coal Geology, 5:43109.Google Scholar
Pittman, E. D., and Larese, R. E. 1991. Compaction of lithic sands: experimental results and applications (1). AAPG Bulletin, 75:12791299.Google Scholar
Pyne, S. J., Andrews, P. L., and Laven, R. D. 1996. Introduction to Wildland Fire, Edition 2. John Wiley and Sons, New York.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
Reid, E. M., and Chandler, M. E. J. 1933. London Clay Flora. British Museum of Natural History, London.Google Scholar
Rex, G., and Chaloner, W. 1983. The experimental formation of plant compression fossils. Palaeontology, 26:231252.Google Scholar
Rice, C., Trewin, N., and Anderson, L. 2002. Geological setting of the Early Devonian Rhynie cherts, Aberdeenshire, Scotland: an early terrestrial hot spring system. Journal of the Geological Society, 159:203214.Google Scholar
Rickard, D., Grimes, S., Butler, I., Oldroyd, A., and Davies, K. L. 2007. Botanical constraints on pyrite formation. Chemical Geology, 236:228246.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
Rieley, G., Collier, R. J., Jones, D. M., Eglinton, G., Eakin, P. A., and Fallick, A. E. 1991. Sources of sedimentary lipids deduced from stable carbon-isotope analyses of individual compounds. Nature, 352:425427.Google Scholar
Rowe, N. P., and Speck, T. 1998. Biomechanics of plant growth forms: the trouble with fossil plants. Review of Palaeobotany and Palynology, 102:4362.Google Scholar
Royer, D. L., Wing, S. L., Beerling, D. J., Jolley, D. W., Koch, P. L., Hickey, L. J., and Berner, R. A. 2001. Paleobotanical evidence for near present-day levels of atmospheric CO2 during part of the Tertiary. Science, 292:23102313.Google Scholar
Schönenberger, J. 2005. Rise from the ashes—the reconstruction of charcoal fossil flowers. Trends in Plant Science, 10:436443.Google Scholar
Schopf, J. M. 1975. Modes of fossil preservation. Review of Palaeobotany and Palynology, 20:2753.Google Scholar
Scott, A., and Rex, G. 1985. The formation and significance of Carboniferous coal balls. Philosophical Transactions of the Royal Society of London B-Biological Sciences, 311:123137.Google Scholar
Scott, A. C. 1989. Observations on the nature and origin of fusain. International Journal of Coal Geology, 12:443475.Google Scholar
Scott, A. C. 2001. Preservation by fire, p. 277280. In Briggs, D. E. G. and Crowther, P. R. (eds.). Paleobiology II. Blackwell Science, Oxford.Google Scholar
Scott, A. C. 2010. Charcoal recognition, taphonomy and uses in palaeoenvironmental analysis. Palaeogeography, Palaeoclimatology, Palaeoecology, 291:1139.Google Scholar
Scott, A. C., Cripps, J. A., Collinson, M. E., and Nichols, G. J. 2000. The taphonomy of charcoal following a recent heathland fire and some implications for the interpretation of fossil charcoal deposits. Palaeogeography, Palaeoclimatology, Palaeoecology, 164:131.Google Scholar
Scott, A. C., and Glasspool, I. J. 2007. Observations and experiments on the origin and formation of inertinite group macerals. International Journal of Coal Geology, 70:5366.Google Scholar
Scott, A. C., Mattey, D. P., and Howard, R. 1996. New data on the formation of Carboniferous coal balls. Review of Palaeobotany and Palynology, 93:317331.Google Scholar
Sigleo, A. C. 1978. Organic geochemistry of silicified wood, Petrified Forest National Park, Arizona. Geochimica et Cosmochimica Acta, 42:13971405.Google Scholar
Smoot, E. L. 1984. Phloem anatomy of the Carboniferous pteridosperm Medullosa and evolutionary trends in gymnosperm phloem. Botanical Gazette, 145:550564.Google Scholar
Smoot, E. L., and Taylor, T. N. 1986. Structurally preserved fossil plants from Antarctica: II. A Permian moss from the Transantarctic Mountains. American Journal of Botany, 72:16831691.Google Scholar
Sørensen, J. 1982. Reduction of ferric iron in anaerobic marine sediment and interaction with reduction of nitrate and sulfate. Applied and Environmental Microbiology, 43:319324.Google Scholar
Spicer, R. A. 1977. The pre-depositional formation of some leaf impressions. Palaeontology, 20:907912.Google Scholar
Spicer, R. A. 1989. The Formation and Interpretation of Plant Fossil Assemblages. Academic Press, London.Google Scholar
Stein, C. 1982. Silica recrystallization in petrified wood. Journal of Sedimentary Research, 52:12771282.Google Scholar
Stopes, M., and Watson, D. 1909. On the present distribution and origin of the calcareous concretions in coal seams, known as “coal balls.” Royal Society of London Philosophical Transactions Series B, 200:167218.Google Scholar
Taylor, T. N., and Taylor, E. L. 2000. The Rhynie Chert ecosystem: a model for understanding fungal interactions, p. 3147. In Bacon, C. W. and White, J. F. (eds.), Microbial Endophytes. Marcel Decker Inc., New York.Google Scholar
Tegelaar, E. W., Kerp, H., Visscher, H., Schenck, P. A., and de Leeuw, J. W. 1991. Bias of the paleobotanical record as a consequence of variations in the chemical composition of higher vascular plant cuticles. Paleobiology, 17:133144.Google Scholar
Tegelaar, E., Matthezing, R., Jansen, J., Horsfield, B., and De Leeuw, J. 1989. Possible origin of n-alkanes in high-wax crude oils. Nature, 342:529531.Google Scholar
Treese, T. N., Owen, R. M., and Wilkinson, B. H. 1981. SrCa and MgCa ratios in polygenetic carbonate allochems from a Michigan marl lake. Geochimica et Cosmochimica Acta, 45:439445.Google Scholar
Trewin, N. H. 1996. The Rhynie cherts: an early Devonian ecosystem preserved by hydrothermal activity, p. 131145. In Bock, G. R. and Goode, J. (eds.), Evolution of Hydrothermal Ecosystems on Earth (and Mars?). Ciba Foundation Symposium 202, Wiley, Chichester.Google Scholar
Tyler, S. A., and Barghoorn, E. S. 1954. Occurrence of structurally preserved plants in pre-Cambrian rocks of the Canadian Shield. Science, 119:606608.Google Scholar
Wakeham, S. G., and Canuel, E. A. 2006. Degradation and preservation of organic matter in marine sediments, p. 295321. In Hitzinger, O. and Volkham, J. K. (eds.), The Handbook of Environmental Chemistry, Vol. 2: Reactions and Processess, Part N: Marine Organic Matter: Biomarkers, Isotopes and DNA. Springer-Verlag, Berlin.Google Scholar
Walton, J. 1936. On the factors which influence the external form of fossil plants, with descriptions of the foliage of some species of the Palaeozoic equisetalean genus Annularia Sternberg. Philosophical Transactions of the Royal Society of London Series B-Biological Sciences, 226:219237.Google Scholar
Wellman, C. H., Osterloff, P. L., and Mohiuddin, U. 2003. Fragments of the earliest land plants. Nature, 425:282285.Google Scholar
Wilson, J. P., Knoll, A. H., Holbrook, N. M., and Marshall, C. R. 2009. Modeling fluid flow in Medullosa, an anatomically unusual Carboniferous seed plant. Paleobiology, 34:472493.Google Scholar
Wing, S. L., DiMichele, W. A., Phillips, T., Taggart, R., Tiffney, B., and Mazer, S. 1992. Ecological characterization of fossil plants, p. 139180. In Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A., Potts, R., Sues, H.-D., and Wing, S. L. (eds.), Terrestrial Ecosystems Through Time: Evolutionary Paleoecology of Terrestrial Plants and Animals. University of Chicago Press, Chicago.Google Scholar
Yang, H., and Huang, Y. 2003. Preservation of lipid hydrogen isotope ratios in Miocene lacustrine sediments and plant fossils at Clarkia, northern Idaho, USA. Organic Geochemistry, 34:413423.Google Scholar
Zhou, Y.-L., Wang, S.-J., Hilton, J., and Tian, B.-L. 2008. Anatomically preserved lepidodendralean plants from lower Permian coal balls of northern China: Achlamydocarpon intermedium sp. nov. Plant Systematics and Evolution, 273:7185.Google Scholar