Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-19T13:48:01.593Z Has data issue: false hasContentIssue false

Taphonomy of the Ediacaran Fossil Pteridinium Simplex Preserved Three-Dimensionally in Mass Flow Deposits, Nama Group, Namibia

Published online by Cambridge University Press:  15 October 2015

Mike Meyer
Affiliation:
Department of Geosciences, Virginia Tech, Blacksburg, VA 24061, USA, ; Department of Geosciences and Natural Resources, Western Carolina University, Cullowhee, NC 28723, USA,
David Elliott
Affiliation:
School of Geosciences, Monash University, Clayton, Victoria, 3800, Australia, ;
James D. Schiffbauer
Affiliation:
Department of Geological Sciences, University of Missouri, Columbia, MO 65211, USA,
Michael Hall
Affiliation:
School of Geosciences, Monash University, Clayton, Victoria, 3800, Australia, ;
Karl H. Hoffman
Affiliation:
Geological Survey of Namibia, Windhoek, Namibia, ;
Gabi Schneider
Affiliation:
Geological Survey of Namibia, Windhoek, Namibia, ;
Patricia Vickers-Rich
Affiliation:
School of Geosciences, Monash University, Clayton, Victoria, 3800, Australia, ;
Shuhai Xiao
Affiliation:
Department of Geosciences, Virginia Tech, Blacksburg, VA 24061, USA, ;

Abstract

Ediacara-type fossils are found in a diverse array of preservational styles, implying that multiple taphonomic mechanisms might have been responsible for their preservational expression. For many Ediacara fossils, the “death mask” model has been invoked as the primary taphonomic pathway. The key to this preservational regime is the replication or sealing of sediments around the degrading organisms by microbially induced precipitation of authigenic pyrite, leading toward fossil preservation along bedding planes. Nama-style preservation, on the other hand, captures Ediacaran organisms as molds and three-dimensional casts within coarse-grained mass flow beds, and has been previously regarded as showing little or no evidence of a microbial preservational influence. To further understand these two seemingly distinct taphonomic pathways, we investigated the three-dimensionally preserved Ediacaran fossil Pteridinium simplex from mass flow deposits of the upper Kliphoek Member, Dabis Formation, Kuibis Subgroup, southern Namibia. Our analysis, using a combination of petrographic and micro-analytical methods, shows that Pteridinium simplex vanes are replicated with minor pyrite, but are most often represented by open voids that can be filled with secondary carbonate material; clay minerals are also found in association with the vanes, but their origin remains unresolved. The scarcity of pyrite and the development of voids are likely related to oxidative weathering and it is possible that microbial activities and authigenic pyrite may have contributed to the preservation of Pteridinium simplex; however, any microbes growing on P. simplex vanes within mass flow deposits were unlikely to have formed thick mats as envisioned in the death mask model. Differential weathering of replicating minerals and precipitation of secondary minerals greatly facilitate fossil collection and morphological characterization by allowing Pteridinium simplex vanes to be parted from the massive hosting sandstone.

Type
Research Article
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

Anderson, E., Schiffbauer, J. D., and Xiao, S. 2011. Taphonomic study of organic-walled microfossils confirms the importance of clay minerals and pyrite in Burgess Shale-type preservation. Geology, 39:643646.Google Scholar
Awonusi, A., Morris, M. D., and Tecklenburg, M. M. J. 2007. Carbonate assignment and calibration in the Raman spectrum of apatite. Calcified Tissue International, 81:4652.CrossRefGoogle ScholarPubMed
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
Butterfield, N. J., Balthasar, U., and Wilson, L. A. 2007. Fossil diagenesis in the Burgess Shale. Palaeontology, 50:537543.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. 2012. Preservational modes in the Ediacaran Gaojiashan Lagerstätte: Pyritization, aluminosilicification, and carbonaceous compression. Palaeogeography, Palaeoclimatology, Palaeoecology, 326–28:109117.Google Scholar
Crimes, T. P. and Fedonkin, M. A. 1996. Biotic changes in platform communities across the Precambrian–Phanerozoic boundary. Rivista Italiana di Paleontologia e Stratigrafia, 102:317332.Google Scholar
Elliott, D. A., Vickers-Rich, P., Trusler, P., and Hall, M. 2011. New evidence on the taphonomic context of the Ediacaran Pteridinium . Acta Palaeontologica Polonica, 56:641650.Google Scholar
Farmer, J., Vidal, G., Moczydlowska, M., Strauss, H., Ahlberg, P., and Siedlecka, A. 1992. Ediacaran fossils from the Innerelv Member (late Proterozoic) of the Tanafjorden area, northeastern Finnmark. Geological Magazine, 129:181195.CrossRefGoogle Scholar
Fedonkin, M. A., Gehling, J. G., Grey, K., Narbonne, G. M., and Vickers-Rich, P. 2007. The Rise of Animals: Evolution and Diversification of the Kingdom Animalia. Johns Hopkins University Press, Baltimore, 326 p.Google Scholar
Gaines, R. R., Briggs, D. E. G., and Zhao, Y. L. 2008. Cambrian Burgess Shale-type deposits share a common mode of fossilization. Geology, 36:755758.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
Gehling, J. G. 1999. Microbial mats in terminal Proterozoic siliciclastics: Ediacaran death masks. Palaios, 14:4057.CrossRefGoogle Scholar
Gehling, J. G. and Droser, M. L. 2009. Textured organic surfaces associated with the Ediacara biota in South Australia. Earth Science Reviews, 96:196206.Google Scholar
Gehling, J. G. and Droser, M. L. 2013. How well do fossil assemblages of the Ediacara Biota tell time? Geology, 41:447450.Google Scholar
Gehling, J. G., Narbonne, G. M., and Anderson, M. M. 2000. The first named Ediacaran body fossil, Aspidella terranovica . Palaeontology, 43:427456.Google Scholar
Germs, G. J. B. 1973. The Nama Group in South-West Africa and its relationship to the pan-African geosyncline. The Journal of Geology, 82:301317.Google Scholar
Gibson, G. G., Teeter, S. A., and Fedonkin, M. A. 1984. Ediacarian fossils from the Carolina slate belt, Stanly County, North Carolina. Geology, 12:387390.Google Scholar
Glaessner, M. F. W. and Wade, M. 1966. The late Precambrian fossils from Ediacara, South Australia. Palaeontology, 9:599628.Google Scholar
Grazhdankin, D. and Seilacher, A. 2002. Underground Vendobionta from Namibia. Palaeontology, 45:5778.Google Scholar
Gresse, P. G. and Germs, G. J. B. 1993. The Nama foreland basin: Sedimentation, major unconformity bounded sequences and multisided active margin advance. Precambrian Research, 63:247272.Google Scholar
Grotzinger, J. P., Bowring, S. A., Saylor, B. Z., and Kaufman, A. J. 1995. Biostratigraphic and geochronologic constraints on early animal evolution. Science, 270:598604.Google Scholar
Gürich, G. 1930. Die bislang ältesten Spuren von Organismen in Südafrika. International Geological Congress. South Africa, 1929 (XV), 2:670680.Google Scholar
Hall, M., Kaufman, A. J., Vickers-Rich, P., Ivanstov, A. Y., Trusler, P., Linnemann, U., Hofmann, M., Elliott, D. A., Cui, H., Fedonkin, M. A., Hoffman, K., Wilson, S. A., Schneider, G., and Smith, J. 2013. Stratigraphy, palaeontology and geochemistry of the late Neoproterozoic Aar Member, southwest Namibia: Reflecting environmental controls on Ediacara fossil preservation during the terminal Proterozoic in African Gondwana. Precambrian Research, 238:214232.Google Scholar
Jenkins, R. J. F. 1992. Functional and ecological aspects of Ediacaran assemblages, p. 131176. In Lipps, J. H. and Signor, P. W. (eds.), Origin and Early Evolution of Metazoa. Plenum Press, New York.Google Scholar
Jenkins, R. J. F., Ford, C. H., and Gehling, J. G. 1983. The Ediacara member of the Rawnsley quartzite: The context of the Ediacara assemblage (late Precambrian, Flinders Ranges). Journal of the Geological Society of Australia 30.Google Scholar
Keller, B. M., Menner, V. V., Stepanov, V. A., and Chumakov, N. M. 1974. New findings of Metazoan in the Vendian of the Russian Platform. Izvestiya Akademii Nauk SSSR, Seriya Geologicheskaya, 12:130134.Google Scholar
Laflamme, M., Darroch, S. A. F., Tweedt, S. M., Peterson, K. J., and Erwin, D. H. 2013. The end of the Ediacara biota: Extinction, biotic replacement, or Cheshire Cat? Gondwana Research, 23:558573.Google Scholar
Laflamme, M., Schiffbauer, J. D., and Narbonne, G. M. 2011 a. Deep-water microbially induced sedimentary structures (MISS) in deep time: The Ediacaran fossil Ivesheadia , p. 111123. In Noffke, N. and Chafetz, H. S. (eds.), Microbial Mats in Siliciclastic Depositional Systems Through Time. Volume 101, SEPM Special Publications.Google Scholar
Laflamme, M., Schiffbauer, J. D., Narbonne, G. M., and Briggs, D. G. 2011 b. Microbial biofilms and the preservation of the Ediacara biota. Lethaia, 44:203213.CrossRefGoogle Scholar
McKeown, D. A., Bell, M. I., and Etz, E. S. 1999. Vibrational analysis of the dioctahedral mica: 2M1 muscovite. American Mineralogist, 84:10411048.Google Scholar
Meyer, M., Schiffbauer, J. D., Xiao, S., Cai, Y., and Hua, H. 2012. Taphonomy of the upper Ediacaran enigmatic ribbonlike fossil Shaanxilithes . Palaios, 27:354372.Google Scholar
Narbonne, G. M. 2005. The Ediacara Biota: Neoproterozoic origin of animals and their ecosystems. Annual Review of Earth and Planetary Sciences, 33:421442.Google Scholar
Narbonne, G. M. and Hofmann, H. J. 1987. Ediacaran biota of the Wernecke Mountains, Yukon, Canada. Palaeontology, 30:647676.Google Scholar
Nicola, J. H., Scott, J. F., Couto, R. M., and Correa, M. M. 1976. Raman spectra of dolomite [CaMg(CO3)2]. Physical Review B, 14:46764678.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
Page, A., Gabbott, S. E., Wilby, P. R., and Zalasiewicz, J. A. 2008. Ubiquitous Burgess Shale-style “clay templates” in low-grade metamorphic mudrocks. Geology, 36:855858.CrossRefGoogle Scholar
Petrovich, R. 2001. Mechanisms of fossilization of the soft-bodied and lightly armored faunas of the Burgess Shale and of some other classical localities. American Journal of Science, 301:683726.Google Scholar
Saylor, B. Z., Kaufman, A. J., Grotzinger, J. P., and Urban, F. 1995. Sequence stratigraphy and sedimentology of the Neoproterozoic Kuibis and Schwarzrand Subgroups (Nama Group), southwestern Namibia. Precambrian Research, 73:153171.Google Scholar
Schiffbauer, J. D., Xiao, S., Sen Sharma, K., and Wang, G. 2012. The origin of intracellular structures in Ediacaran metazoan embryos. Geology, 40:223226.Google Scholar
Schmitz, M. D. 2012. Appendix 2—Radiometric ages used in GTS2012, p. 10451082. In Gradstein, F., Ogg, J., Schmitz, M. D., and Ogg, G. (eds.), The Geologic Time Scale 2012. Elsevier, Boston.Google Scholar
Seilacher, A. 1992. Vendobionta and Psammocorallia: Lost constructions of Precambrian evolution. Journal of the Geological Society of London, 149:607613.Google Scholar
St. Jean, J. 1973. A new Cambrian trilobite from the Piedmont of North Carolina. American Journal of Science, 273–A:196216.Google Scholar
Vickers-Rich, P. 2007. The Nama fauna of southern Africa, p. 6988. The Rise of Animals: Evolution and Diversification of the Kingdom Animalia. Johns Hopkins University Press.Google Scholar
Vickers-Rich, P., Ivantsov, A. Y., Trusler, P. W., Narbonne, G. M., Hall, M., Wilson, S. A., Greentree, C., Fedonkin, M. A., Elliott, D. A., Hoffmann, K. H., and Schneider, G. I. C. 2013. Reconstructing Rangea: New discoveries from the Ediacaran of southern Namibia. Journal of Paleontology, 87:115.Google Scholar
Wada, N. and Kamitakahara, W. A. 1991. Inelastic neutron- and Raman-scattering studies of muscovite and vermiculite layered silicates. Physical Review B, 45:23912397.Google Scholar
Wang, A., Freeman, J., and Kuebler, K. E. 2002. Raman Spectroscopic Characterization of Phyllosilicates. Lunar and Planetary Science, XXXIII:13741375.Google Scholar
Wopenka, B. and Pasteris, J. D. 1993. Structural characterization of kerogens to granulite-facies graphite: Applicability of Raman microprobe spectroscopy. American Mineralogists, 78:533557.Google Scholar
Xiao, S. and Knoll, A. H. 2000. Phosphatized animal embryos from the Neoproterozoic Doushantuo Formation at Weng'an, Guizhou, South China. Journal of Paleontology, 74:767788.Google Scholar
Xiao, S. and Laflamme, M. 2009. On the eve of animal radiation: Phylogeny, ecology and evolution of the Ediacara biota. Trends in Ecology and Evolution, 24:3140.CrossRefGoogle ScholarPubMed
Xiao, S. and Schiffbauer, J.D. 2009. Fossil preservation through phosphatization and its astrobiological implications, p. 89118. In Seckbach, J. and Walsh, M. (eds.), From Fossils to Astrobiology. Springer, Heidelberg.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.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, 76:347376.2.0.CO;2>CrossRefGoogle Scholar
Yuan, X., Chen, Z., Xiao, S., Zhou, C., and Hua, H. 2011. An early Ediacaran assemblage of macroscopic and morphologically differentiated eukaryotes. Nature, 470:390393.Google Scholar
Supplementary material: File

Meyer et al. supplementary material

Meyer et al. supplementary material

Download Meyer et al. supplementary material(File)
File 7.4 MB