Hostname: page-component-8448b6f56d-jr42d Total loading time: 0 Render date: 2024-04-24T16:34:05.842Z Has data issue: false hasContentIssue false

Reconstructing Rangea: new discoveries from the Ediacaran of southern Namibia

Published online by Cambridge University Press:  20 May 2016

Patricia Vickers-Rich
Affiliation:
School of Geosciences, Monash University, Clayton, Vic. 3800, Australia,
Andrey Yu. Ivantsov
Affiliation:
Paleontological Institute, Russian Academy of Sciences, Profsoyuznaya ul. 123, Moscow 117997, Russia,
Peter W. Trusler
Affiliation:
School of Geosciences, Monash University, Clayton, Vic. 3800, Australia,
Guy M. Narbonne
Affiliation:
Department of Geological Sciences and Geological Engineering, Queens University, Kingston, Ontario, Canada,
Mike Hall
Affiliation:
School of Geosciences, Monash University, Clayton, Vic. 3800, Australia,
Sasha Wilson
Affiliation:
School of Geosciences, Monash University, Clayton, Vic. 3800, Australia,
Carolyn Greentree
Affiliation:
School of Geosciences, Monash University, Clayton, Vic. 3800, Australia,
Mikhail A. Fedonkin
Affiliation:
Paleontological Institute, Russian Academy of Sciences, Profsoyuznaya ul. 123, Moscow 117997, Russia,
David A. Elliott
Affiliation:
School of Geosciences, Monash University, Clayton, Vic. 3800, Australia,
Karl H. Hoffmann
Affiliation:
Namibian Geological Survey, Ministry of Mines and Energy, Windhoek, Namibia,
Gabi I. C. Schneider
Affiliation:
Namibian Geological Survey, Ministry of Mines and Energy, Windhoek, Namibia,

Abstract

Rangea is the type genus of the Rangeomorpha, an extinct clade near the base of the evolutionary tree of large, complex organisms which prospered during the late Neoproterozoic. It represents an iconic Ediacaran taxon, but the relatively few specimens previously known significantly hindered an accurate reconstruction. Discovery of more than 100 specimens of Rangea in two gutter casts recovered from Farm Aar in southern Namibia significantly expands this data set, and the well preserved internal and external features on these specimens permit new interpretations of Rangea morphology and lifestyle. Internal structures of Rangea consist of a hexaradial axial bulb that passes into an axial stalk extending the length of the fossil. The axial bulb is typically filled with sediment, which becomes increasingly loosely packed and porous distally, with the end of the stalk typically preserved as an empty, cylindrical cone. This length of the axial structure forms the structural foundation for six vanes arranged radially around the axis, with each vane consisting of a bilaminar sheet composed of a repetitive pattern of elements exhibiting at least three orders of self-similar branching. Rangea was probably an epibenthic frond that rested upright on the sea bottom, and all known fossil specimens were transported prior to their final burial in storm deposits.

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.)

Footnotes

Article updated on May 25, 2023.

References

Abelson, J., Metz, J. M., McLoughlin, N., Cohen, P. A., and Tice, M. M. 2012. Deep water incised valley deposits at the Proterozoic–Cambrian boundary in southern Namibia contain abundant Treptichnus pedum . Palaios, 27:252273.Google Scholar
Bladh, K. W. 1982. The formation of goethite, jarosite, and alunite during the weathering of sulfide-bearing felsic rocks. Economic Geology, 77:176184.Google Scholar
Bouougri, E. H., Porada, H., Weber, K., and Reitner, J. 2011. Sedimentology and palaeoecology of Ernietta-bearing Ediacaran deposits in southern Namibia: implications for infaunal vendobiont communities. Advances in Stromatolite Geobiology. Lecture notes in Earth Sciences, 313:473506.Google Scholar
Brasier, M. D. 1979. The Cambrian radiation event, p. 103159. In House, M. R. (ed.), The Origin of Major Invertebrate Groups. Systematics Association Special Volume No. 12. Academic Press, London.Google Scholar
Brasier, M. D. and Antcliffe, J. 2004. Decoding the Ediacaran enigma. Science, 305:11151117.Google Scholar
Brasier, M. D. and Antcliffe, J. 2009. Evolutionary relationships within the Avalonian Ediacara biota: new insights from laser analysis. Journal of the Geological Society, London, 166:363384.Google Scholar
Cerrano, C., Calcinai, B., Gioia De Camillo, C., Valisano, L., and Bavestrello, G. 2007. How and why do sponges incorporate foreign material? Strategies in Porifera, p. 239246. In Custódio, M. R., Lôbo-Hajdu, G., Hajdu, E., and Muricy, G. (eds.), Porifera Research: Biodiversity, Innovation and Sustainability. Série Livros 28. Museu Nacional, Rio de Janeiro.Google Scholar
Chia, F. S. 1973. Sand dollar: a weight belt for the juvenile. Science, 181:7374.Google Scholar
Dzik, J. 2002. Possible ctenophoran affinities of the Precambrian “sea-pen” Rangea . Journal of Morphology, 252:315334.Google Scholar
Elliott, D. A., Vickers-Rich, P., Trusler, P. W., and Hall, M. 2011. New evidence on the taphonomic context of the Ediacaran Pteridinium . Acta Palaeontologica Polonica, 56:641650.Google Scholar
Erwin, D. H., Laflamme, M., Tweedt, S. M., Sperling, E. A., Pisani, D., and Peterson, K. J. 2011. The Cambrian conundrum: early divergence and later ecological success in the early history of animals. Science, 334:10911097.Google Scholar
Flude, L. I. and Narbonne, G. M. 2008. Taphonomy and ontogeny of a multibranched Ediacaran fossil: Bradgatia from the Avalon Peninsula of Newfoundland. Canadian Journal of Earth Sciences, 45:10951109.Google Scholar
Gehling, J. G. 1991. The case for Ediacaran roots to a metazoan tree. Geological Society of India Memoir, 20:181224.Google Scholar
Gehling, J. G. 1999. Microbial mats in terminal Neoproterozoic siliciclastics: Ediacaran death masks. Palaios, 14:4057.Google Scholar
Gehling, J. G. and Narbonne, G. M. 2007. Spindle-shaped fossils from the Mistaken Point assemblage, Avalon Zone, Newfoundland. Canadian Journal of Earth Sciences, 44:367387.Google Scholar
Germs, G. J. B. 1972. The stratigraphy and palaeontology of the lower Nama Group, South West Africa. Precambrian Research Unit, University of Cape Town, Bulletin 12, 1250.Google Scholar
Germs, G. J. B. 1973. A reinterpretation of Rangea schneiderhoehni and the discovery of a related new fossil from the Nama Group, South West Africa. Lethaia, 6:19.Google Scholar
Germs, G. J. B. 1983. Implications of a sedimentary facies and depositional analysis of the Nama Group in South West Africa, Namibia. Geological Society of South Africa, Special Publication, 11:89114.Google Scholar
Gooday, A. J., Aranda Da Silva, A., and Pawlowski, J. 2011. Xenophyophores (Rhizaria, Foraminifera) from the Nazare'Canyon (Portuguese margin, NE Atlantic). Deep Sea Reasearch II, 58:24012419.Google Scholar
Grazhdankin, D. V. 2004. Patterns of distribution in the Ediacaran biotas: facies versus biogeography and evolution. Paleobiology, 30:203221.Google Scholar
Grazhdankin, D. V. and Seilacher, A. 2002. Underground Vendobionta from Namibia. Palaeontology, 45:5778.Google Scholar
Grazhdankin, D. V. and Seilacher, A. 2005. A re-examination of the Nama-type Vendian organism Rangea schneiderhoehni . Geological Magazine, 142:571582.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, 270:598.Google Scholar
Grotzinger, J. P. and Miller, R. 2008. The Nama Group, p. 13.22913.272. In Miller, R. (ed.), The Geology of Namibia. Geological Society of Namibia Special Publication, 2.Google Scholar
Gürich, G. 1930. Die bislang altesten Spuren von Organismen in Sudafrika. International Geological Congress (XV), Pretoria, Union of South Africa, Die altesten Fossilien Sud-Afrikas, 2:670680.Google Scholar
Gürich, G. 1933. Die Kuibis-Fossilien der Nama-Formation von Sudwestafrika. Paläontologische Zeitschrift, 15:137154.Google Scholar
Haywick, D. W. and Mueller, E. M. 1997. Sediment retention in encrusting Palythoa sp.—a biological twist to a geological process. Coral Reefs, 16:3946.Google Scholar
Jambor, J. L. and Blowes, D. W. 1998. Theory and applications of mineralogy in environmental studies of sulfide-bearing mine wastes, p. 367401. In Cabri, L. J. and Vaughan, D. J. (eds.), Modern Approaches to Ore and Environmental Mineralogy, 27. Mineralogical Association of Canada Short Course Series, Mineralogical Association of Canada, Ottawa, Canada.Google Scholar
Jenkins, R. J. F. 1985. The enigmatic Ediacaran (Late Precambrian) genus Rangea and related forms. Paleobiology, 11:336355.Google Scholar
Laflamme, M. and Narbonne, G. M. 2008a. Ediacaran fronds. Palaeogeography, Palaeoclimatology, Palaeoecology, 258:162179.Google Scholar
Laflamme, M. and Narbonne, G. M. 2008b. Competition in a Precambrian world: palaeoecology of Ediacaran fronds. Geology Today, 24:182187.Google Scholar
Laflamme, M., Xiao, S., and Kowalewski, M. 2009. Osmotrophy in modular Ediacaran organisms. Proceedings of the National Academy of Sciences of U.S.A., 106:1443814443.Google Scholar
Lottermoser, B. G. 2010. Mine Wastes: Characterization, Treatment and Environmental Impacts (third edition). Springer-Verlag, Berlin, Germany, 400 p.Google Scholar
Mooi, R. and Chen, C. P. 1996. Weight belts, diverticula, and the phylogeny of the sand dollars. Bulletin of Marine Science, 58:186195.Google Scholar
Myrow, P. M. 1992. Pot and gutter casts from the Chapel Island Formation, southeast Newfoundland. Journal of Sedimentary Research, 62:9921007.Google Scholar
Myrow, P. M. and Southard, J. B. 1996. Tempestite deposition. Journal of Sedimentary Research, 66:875887.Google Scholar
Narbonne, G. M. 2004. Modular construction of early Ediacaran complex life forms. Science, 305:11411144.Google Scholar
Narbonne, G. M., Saylor, B. Z., and Grotzinger, J. P. 1997. The youngest Ediacaran fossils from southern Africa. Journal of Paleontology, 71:953969.Google Scholar
Narbonne, G. M., Laflamme, M., Greentree, C., and Trusler, P. 2009. Reconstructing a lost world: Ediacaran rangeomorphs from Spaniard's Bay, Newfoundland. Journal of Paleontology, 83:503523.Google Scholar
Narbonne, G.M., Xiao, S., and Shields, G. 2012. The Ediacaran Period, p. 427449. In Gradstein, F., Ogg, J., Schmitz, M. D., and Ogg, G. (eds), Geologic Timescale 2012. Elsevier, Boston.Google Scholar
Pflüg, H. D. 1970. Zur fauna der Nama-Schichten in Südwest-Afrika; II. Rangeidae, Bau und systematische Zugehörigkeit. Palaeontographica Abteilung A, 135:198231.Google Scholar
Pflüg, H. D. 1972. Systematik der jung-präkambrischen Petalonamae. Paläontologische Zeitschrift, 46:5667.Google Scholar
Plint, A. G. 2010. Wave- and storm-dominated shoreline and shallow-marine systems, p. 167199. In James, N. P. and Dalrymple, R. W. (eds.), Facies Models 4. Geological Association of Canada.Google Scholar
Reimer, J. D., Nakachi, S., Hirose, M., and Shinji, H. 2010. Using hydrofluoric acid for morphological investigations of Zoanthids (Cnidaria: Anthozoa): a critical assessment of methodology and necessity. Marine Biotechnology, 12:605617.Google Scholar
Richter, R. 1955. Die ältesten Fossilien Süd-Afrikas. Senckenbergiana Lethaea, 36:243289.Google Scholar
Savazzi, E. 2007. A new reconstruction of Protolyellia (Early Cambrian psammocoral), p. 339353. In Vickers-Rich, P. and Komarower, P. (eds.), The Rise and Fall of the Ediacaran Biota. Geological Society of London Special Publication 286.Google Scholar
Saylor, B. Z., Grotzinger, J. P., and Germs, G. J. B. 1995. Sequence stratigraphy and sedimentology of the Neoproterozoic Kuibis and Schwarzrand subgroups (Nama Group), southwestern Namibia. Precambrian Research, 73:153172.Google Scholar
Saylor, B. Z., Kaufman, A. J., Grotzinger, J. P., and Urban, F. 1998. A composite reference section for terminal Proterozoic strata of southern Namibia. Journal of Sedimentary Research, 68:12231235.Google Scholar
Saylor, B. Z., Poling, J. M., and Huff, W. D. 2005. Stratigraphic and chemical correlation of volcanic ash beds in the terminal Proterozoic Nama Group, Namibia. Geological Magazine, 142:519538.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. 1979. Constructional morphology of sand dollars. Paleobiology, 5:191221.Google Scholar
Seilacher, A. 1992. Vendobionta and Psammocorallia: lost constructions of Precambrian evolution. The Journal of the Geological Society of London, 149:607614.Google Scholar
Seilacher, A. 1999. Biomat-related lifestyles in the Precambrian. Palaios, 14:8693.Google Scholar
Seilacher, A. 2007. The nature of Vendobionts. Geological Society of London, Special Publication 286 (1):387397.Google Scholar
Tendal, O. S. 1972. A monography of the Xenophyophoria (Rhizopodea, Protozoa). Galathea Report, 12:799.Google Scholar
Vickers-Rich, P. 2007. The Nama fauna of southern Africa, p. 6988. In Fedonkin, M. A., Gehling, J. G., Grey, K., Narbonne, G. M., and Vickers-Rich, P. (eds.), The Rise of Animals: Evolution and Diversification of the Kingdom Animalia. Johns Hopkins University Press.Google Scholar
Wood, D. A., Dalrymple, R. W., Narbonne, G. M., Gehling, J. G., and Clapham, M. E. 2003. Paleoenvironmental analysis of the late Neoproterozoic Mistaken Point and Trepassey formations, southeastern Newfoundland. Canadian Journal of Earth Sciences, 40:13751391.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.Google Scholar