Hostname: page-component-848d4c4894-2pzkn Total loading time: 0 Render date: 2024-06-02T18:55:56.073Z Has data issue: false hasContentIssue false

Paleoecologic and Biostratigraphic Significance of Trace Fossils From Shallow- to Marginal-Marine Environments From the Middle Cambrian (Stage 5) of Jordan

Published online by Cambridge University Press:  20 May 2016

Richard Hofmann
Affiliation:
Freiberg University, Geological Institute, Department of Palaeontology, Bernhard-von-Cotta-street 2, Freiberg, D-09599 Germany, ; current address: Palaeontological Institute and Museum, Zurich University, Karl-Schmid Strasse 4, CH-8006, Zürich, Switzerland,
M. Gabriela Mángano
Affiliation:
Department of Geological Sciences, University of Saskatchewan, Saskatoon, SK S7N 5E2, Canada,
Olaf Elicki
Affiliation:
Freiberg University, Geological Institute, Department of Palaeontology, Bernhard-von-Cotta-street 2, Freiberg, D-09599 Germany, ;
Rafie Shinaq
Affiliation:
Al Al-Bayt University, Institute of Earth and Environmental Sciences, Mafraq, Jordan,

Abstract

The Hanneh Member (Cambrian Stage 5) of the Burj Formation and the Umm Ishrin Formation of Jordan represent a transgressive-regressive succession that contains twenty-eight ichnotaxa, including vertical burrows (Arenicolites isp., Diplocraterion isp., Gyrolithes polonicus, Rosselia isp., Skolithos linearis, escape trace fossils), horizontal simple burrows and trails (Archaeonassa fossulata, Gordia marina, Helminthoidichnites tenuis, Palaeophycus tubularis, Planolites beverleyensis, P. montanus), plug-shaped burrows (Bergaueria sucta), horizontal branched burrows (Asterosoma isp., Phycodes isp., Treptichnus cf. T. pedum), bilobate structures (various ichnospecies of Cruziana and Rusophycus), and trackways and scratch marks (Diplichnites isp., Dimorphichnus cf. D. obliquus, Monomorphichnus isp.). Eleven trace-fossil assemblages are identified. The Arenicolites isp. and Diplocraterion isp. assemblages occur in transgressive tidal dunes and bars whereas the Rosselia isp. assemblage characterizes areas between tidal dunes. The Cruziana salomonis assemblage reflects a wide variety of environmental settings including channels within tidal-bar complexes, bottomsets of tidal dunes, and interdune areas. The Gordia marina assemblage is present between dune patches. The Gyrolithes polonicus assemblage penetrates into firmground mudstone below the maximum flooding surface. The Bergaueria sucta, Archaeonassa fossulata, Rusophycus aegypticus and Cruziana problematica assemblages occur in different subenvironments of the progradational delta. Cruziana salomonis and Rusophycus burjensis, originally considered indicative of an early Cambrian age, are actually middle Cambrian in their type locality. Occurrences of Cruziana jordanica and Rusophycus aegypticus provide evidence that these ichnospecies are of the same age in Jordan and may co-exist in terms of stratigraphic distribution with C. salomonis and R. burjensis.

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

Amireh, B. S., Schneider, W., and Abed, A. M. 1994. Evolving fluvial transitional marine deposition through the Cambrian sequence of Jordan. Sedimentary Geology, 89:6590.Google Scholar
Amireh, B. S., Amaireh, M. N., and Abed, A. M. 2008. Tectono sedimentary evolution of the Umm Ghaddah Formation (late Ediacaran–early Cambrian) in Jordan. Journal of Asian Earth Sciences, 33:194218.Google Scholar
Alpert, S. P. 1974. Systematic review of genus Skolithos. Journal of Paleontology, 48:661669.Google Scholar
Bandel, K. and Shinaq, R. 2003. Sediments of the Precambrian Wadi Abu Burqa Formation influenced by life and their relation to the Cambrian sandstones in southern Jordan. Freiberger Forschungshefte C, 499:8791.Google Scholar
Bender, F. 1974. Geology of Jordan. Gebrüder Borntraeger, Berlin, 196p.Google Scholar
Bergström, J. 1976. Lower Palaeozoic trace fossils from eastern Newfoundland. Canadian Journal of Earth Sciences, 13:16131633.CrossRefGoogle Scholar
Berting, M., Braddy, R., Bromley, R. G., Demathieu, G., Genise, J., Mikuláš, R., Nielsen, J., Nielsen, K., Rindsberg, A., Schlirf, M., and Uchman, A. 2006. Names for trace fossils: A uniform approach. Lethaia, 39:265286.CrossRefGoogle Scholar
Billings, E. 1862. New species of fossils from different parts of the Lower, Middle and Upper Silurian rocks of Canada, p. 96168. InPalaeozoic Fossils, Vol. 1 (18611865). Geological Survey of Canada.Google Scholar
Brandt, D. S. 2008. Multiple Rusophycus (Arthropod ichnofossil) assemblages and their significance. Ichnos, 15:2843.Google Scholar
Bromley, R. G. 1996. Trace Fossils: Biology, Taphonomy and Applications. Chapman and Hall, London, 361p.Google Scholar
Bromleyh, R. G. and Asgaard, U. 1972. Freshwater Cruziana from the Upper Triassic of Jameson Land, East Greenland. Gr⊘nlands Geologiske Unders⊘gelse Rapport, 49:1521.Google Scholar
Buatois, L. A. and Mángano, M. G. 2002. Trace fossils from Carboniferous floodplain deposits in western Argentina: implications for ichnofacies models of continental environments. Palaeogeography Palaeoclimatology Palaeoecology, 183:7186.Google Scholar
Buatois, L. A. and Mángano, M. G. 2011. Ichnology: Organism-Substrate Interactions in Space and Time. Cambridge University Press, New York, 358p.Google Scholar
Buatois, L. A., Gingras, M. K., MacEachern, J., Mángano, M. G., Zonneveld, J. P., Pemberton, S. G., Netto, R. G., and Martin, A. 2005. Colonization of brackish-water systems through time: evidence from the trace fossil record. PALAIOS, 20:321347.Google Scholar
Buatois, L. A., Jalfin, G., and Aceñolaza, F. G. 1997. Permian nonmarine invertebrate trace fossils from southern Patagonia, Argentina: ichnologic signatures of substrate consolidation and colonization sequences. Journal of Paleontology, 71:324336.CrossRefGoogle Scholar
Buatois, L. A., Mángano, M. G., Maples, C. G., and Lanier, W. P. 1998. Ichnology of an Upper Carboniferous fluvio-estuarine paleovalley: The Tonganoxie Sandstone, Buildex Quarry, Eastern Kansas, U.S.A.Journal of Paleontology, 72:152180.Google Scholar
Buatois, L. A., Moya, M. C., Mángano, M. G., and Malanca, S. 2003. Paleoenvironmental and sequence stratigraphic framework of the Cambrian–Ordovician transition in the Angosto del Moreno area, northwest Argentina, Proceedings of the 9th International Symposium on the Ordovician System, San Juan, p. 397401.Google Scholar
Buatois, L. A., Santiago, N., Parra, K., and Steel, R. 2008. Animal-substrate interactions in an early Miocene wave-dominated tropical delta: delineating environmental stresses and depositional dynamics (Tacata field, eastern Venezuela). Journal of Sedimentary Research, 78:458479.Google Scholar
Buck, S. G. and Goldring, R. 2003. Conical sedimentary structures, trace fossils or not? Observations, experiments, and review. Journal of Sedimentary Research, 73:338353.Google Scholar
Buckmann, J. O. 1994. Archaeonassa Fenton and Fenton 1937 reviewed. Ichnos, 3:185192.CrossRefGoogle Scholar
Campbell, K. A., Nesbitt, E. A., and Bourgeois, J. 2006. Signatures of storms, oceanic floods and forearc tectonism in marine shelf strata of the Quinault Formation (Pliocene), Washington, U.S.A. Sedimentology, 53:945969.Google Scholar
Carmona, N. B., Buatois, L. A., Ponce, J. J., and Mángano, M. G. 2009. Ichnology and sedimentology of a tide-influenced delta, lower Miocene Chenque Formation, Patagonia, Argentina: trace-fossil distribution and response to environmental stresses. Palaeogeography, Palaeoclimatology, Palaeoecology, 273:7586.Google Scholar
Chamberlain, C. K. 1971. Morphology and ethology of trace fossils from the Ouachita Mountains, southeast Oklahoma. Journal of Paleontology, 45:212246.Google Scholar
Cornish, F. G. 1986. The trace-fossil Diplocraterion; evidence of animal-sediment interactions in Cambrian tidal deposits. PALAIOS, 1:478491.Google Scholar
Crimes, T. P. 1970. Trilobite tracks and other trace fossils from the upper Cambrian of North Wales. Geological Journal, 7:4768.Google Scholar
Crimes, T. P. 1975. The production and preservation of trilobite resting and furrowing traces. Lethaia, 8:3548.Google Scholar
Crimes, T. P. and Anderson, M. M. 1985. Trace fossils from late Precambrian–early Cambrian strata of southeastern Newfoundland (Canada)—temporal and environmental implications. Journal of Paleontology, 59:310343.Google Scholar
Dahmer, G. 1937. Lebensspuren aus dem Taunusquarzit und den Siegener Schichten (Unterdevon). Preussischen Geologischen Landesanstalt zu Berlin Jahrbuch, 57:523539.Google Scholar
Dawson, J. W. 1873. Impressions and footprints of aquatic animals and imitative markings on Carboniferous rocks. American Journal of Science and Art, 105:1624.Google Scholar
Desjardins, P. R., Buatois, L. A., Mángano, M. G., and Pratt, B. R. 2012. Subtidal sandbody architecture and ichnology in the early Cambrian Gog Group of western Canada: Implications for an integrated sedimentologic-ichnologic model of tide-dominated shelf settings. Sedimentology, 59:14521477.Google Scholar
Desjardins, P. R., Mángano, M. G., Buatois, L. A., and Pratt, B. R. 2010. Skolithos pipe rock and associated ichnofabrics from the southern Rocky Mountains, Canada: Colonization trends and environmental controls in an early Cambrian sand-sheet complex. Lethaia, 43:507528.Google Scholar
Donovan, S. K. 2010. Cruziana and Rusophycus: Trace fossils produced by trilobites … in some cases? Lethaia, 43:283284.Google Scholar
Dornbos, S. Q., Bottjer, D. J., and Chen, J. Y. 2004. Evidence for seafloor microbial mats and associated metazoan lifestyles in lower Cambrian phosphorites of southwest China. Lethaia, 37:127137.Google Scholar
Dornbos, S. Q., Bottjer, D. J., and Chen, J. Y. 2005. Paleoecology of benthic metazoans in the early Cambrian Maotianshan Shale biota and the middle Cambrian Burgess Shale biota: Evidence for the Cambrian substrate revolution. Palaeogeography, Palaeoclimatology, Palaeoecology, 220:4767.Google Scholar
Droser, M. L., Jensen, S., and Gehling, J. G. 2002. Trace fossils and substrates of the terminal Proterozoic–Cambrian transition: implications for the record of early bilaterians and sediment mixing. Proceedings of the National Academy of Sciences of the United States of America, 99:1257212576.Google Scholar
Droser, M. L., Jensen, S., and Gehling, J. G. 2004. Development of early Palaeozoic ichnofabrics: Evidence from shallow marine siliciclastics. Geological Society, London, Special Publications, 228:383396.Google Scholar
Elicki, O. 2007. Facies development during late early–middle Cambrian (Tayan Member, Burj Formation) transgression in the Dead Sea Rift valley, Jordan. Carnets de Géologie / Notebooks on Geology, 2007 (7):120.Google Scholar
Elicki, O. 2011. First skeletal microfauna from the Cambrian Series 3 of the Jordan Rift Valley (Middle East). Memoirs of the Association of Australasian Palaeontologists, 42:153173.Google Scholar
Elicki, O. and Geyer, G. In press. The Cambrian trilobites of Jordan—taxonomy, systematic and stratigraphic significance. Acta Geologica Polonica.Google Scholar
Elicki, O., Schneider, J., and Shinaq, R. 2002. Prominent facies from the lower/middle Cambrian of the Dead Sea area (Jordan) and their palaeodepositional significance. Bulletin de la Societe Geologique de France, 173:547552.Google Scholar
Emmons, E. 1844. The Taconic system; based on Observations in New-York, Massachusetts, Maine, Vermont and Rhode-Island. Carroll and Cook, Albany, 65p.Google Scholar
Eerdoğan, B., Uchman, A., Gungor, T., and Ozgul, N. 2004. Lithostratigraphy of the lower Cambrian metaclastics and their age based on trace fossils in the Sandikli region, southwestern Turkey. Geobios, 37:346360.Google Scholar
Fedonkin, M. A. 1981. Belomorskaya biota venda. Trudy Akademii Nauk SSSR, 342:1100.Google Scholar
Fenton, C. L. and Fenton, M. A. 1937. Trilobite ‘nests' and feeding burrows. The American Midland Naturalist, 18:446451.Google Scholar
Fillion, D. and Pickerill, R. K. 1990. Ichnology of the upper Cambrian? to Lower Ordovician Bell Island and Wabana groups of eastern Newfoundland. Palaeontographica Canadiana, 7, 119p.Google Scholar
Fitch, A. 1850. A historical, topographical and agricultural survey of the County of Washington, Part 2–5. Transactions of the New York Agricultural Society, 9:753944.Google Scholar
Frey, R. W. and Howard, J. D. 1985. Trace Fossils from the Panther Member, Star Point Formation (Upper Cretaceous), Coal Creek Canyon, Utah. Journal of Paleontology, 59:370404.Google Scholar
Fürsich, F. T. 1974 a. Diplocraterion Torell 1870 and significance of morphological features in vertical, spreiten-bearing, U-shaped trace fossils. Journal of Paleontology, 48:952962.Google Scholar
Fürsich, F. T. 1974 b. Corallian (Upper Jurassic) trace fossils from England and Normandy. Stuttgarter Beiträge zur Naturkunde, Serie B (Geologie und Paläontologie), 13:115.Google Scholar
Gaigalas, A. and Uchman, A. 2004. Trace fossils form the upper Pleistocene varved clays S of Kaunas, Lithuania. Geologija, 45:1626.Google Scholar
Gernant, R. E. 1972. Paleoenvironmental significance of Gyrolithes (Lebensspur). Journal of Paleontology, 46:735741.Google Scholar
Geyer, G. and Landing, E. 2004. A unified lower–middle Cambrian chronostratigraphy for West Gondwana. Acta Geologica Polonica, 54:179218.Google Scholar
Geyer, G. and Uchman, A. 1995. Ichnofossil assemblages from the Nama Group (Neoproterozoic–lower Cambrian) in Namibia and the Proterozoic–Cambrian boundary problem revisited, p. 175202. InGeyer, G. and Landing, E.(eds.), Morocco95The Lower Cambrian–Middle Cambrian Standard of Western Gondwana, Beringeria Special Issue 2.Google Scholar
Gingras, M. K., MacEachern, J. A., and Pemberton, S. G. 1998. A comparative analysis of the ichnology of wave- and river-dominated allomembers of the Upper Cretaceous Dunvegan Formation. Bulletin of Canadian Petroleum Geology, 46:5173.Google Scholar
Golding, R. and Seilacher, A. 1971. Limulid undertracks and their sedimentological implications. Neues Jahrbuch Fur Geologie Und Palaontologie–Abhandlungen, 137:422442.Google Scholar
Haldeman, S. S. 1840. Supplement to number one of ‘A monograph of the Limniades, and other freshwater bivalve shells of the apparently new animals in different classes, and names and characters of the subgenera in Paludina and Anculosa.’ J. Dobson, Philadelphia, 3p.Google Scholar
Hall, J. 1847. Palæontology of New York, Vol. I. Van Benthuysen, Albany, NY, 1, 338p.Google Scholar
Hall, J. 1852. Palæontology of New York, Vol. II. Van Benthuysen, Albany, NY, 2, 362p.Google Scholar
Han, Y. J. and Pickerill, R. K. 1994. Phycodes templus isp. nov. from the Lower Devonian of Northwestern New-Brunswick, Eastern Canada. Atlantic Geology, 30:3746.Google Scholar
Häntzschel, W. 1975. Miscellanea. Supplement 1, Trace fossils and problematica, 269 p. InTeichert, C.(ed.), Treatise on invertebrate paleontology. Geological Society of America, University of Kansas Press, Boulder, Colorado, Lawrence.Google Scholar
Hofmann, H. J. 1990. Computer simulation of trace fossils with random patterns, and the use of goniograms. Ichnos, 1:1522.Google Scholar
Hofmann, H. J. and Patel, I. M. 1989. Trace fossils from the type Etcheminian Series (lower Cambrian Ratcliffe Brook Formation), Saint John Area, New Brunswick, Canada. Geological Magazine, 126:139157.Google Scholar
Husseini, M. I. 1989. Tectonic and deposition model of late Precambrian–Cambrian Arabian and adjoining plates. American Association of Petroleum Geologists Bulletin, 73:11171131.Google Scholar
Jarrar, G., Wachendorf, H., and Zellmer, H. 1991. The Saramuj conglomerate: evolution of a Pan-African molasses sequence from southwest Jordan. Neues Jahrbuch für Geologie und Paläontologie, Monatshefte, 6:335356.CrossRefGoogle Scholar
Jensen, S. 1997. Trace fossils from the lower Mickwitzia sandstone, south-central Sweden. Fossils and Strata, 42:1110.Google Scholar
Jensen, S. 2003. The Proterozoic and earliest Cambrian trace fossil record; patterns, problems and perspectives. Integrative and Comparative Biology, 43:219228.Google Scholar
Jensen, S. and Bergström, J. 2000. Cheiichnus gothicus igen. et isp n., a new Bergaueria-like arthropod trace fossil from the lower Cambrian of Vastergotland, Sweden. Gff, 122:293296.CrossRefGoogle Scholar
Jensen, S., Droser, M. L., and Gehling, J. G. 2006. A critical look at the Ediacaran trace fossil record, p. 115157. InXiao, S. and Kaufman, A. J.(eds.), Neoproterozoic Geobiology and Paleobiology. Volume 27. Springer Netherlands.CrossRefGoogle Scholar
Keighley, D. G. and Pickerill, R. K. 1995. The ichnotaxa Palaeophycus and Planolites: Historical perspectives and recommendations. Ichnos, 3:301309.Google Scholar
Keighley, D. G. and Pickerill, R. K. 1996. Small Cruziana, Rusophycus, and related ichnotaxa from eastern Canada: The nomenclatural debate and systematic ichnology. Ichnos, 4:261285.Google Scholar
Linnarssön, J. G. O. 1869. Om några försteningar från Vestergötlands sandstenlager. Ovfersigt af Kongliga. Vetenskaps-Akademiens Förhandlingar, 3:337357.Google Scholar
MacEachern, J. A., Raychaudhuri, I., and Pemberton, S. G. 1992. Stratigraphic applications of the Glossifungites ichnofacies: Delineating discontinuities in the rock record, p. 169198. InPemberton, S. G.(ed.), In Applications of Ichnology to Petroleum Exploration: A Core Workshop. Volume 17. Society for Sedimentary Geology.Google Scholar
MacEachern, J. A., Bann, K. L., Pemberton, S. G., and Gingras, M. K. 2007. The Ichnofacies paradigm: High-resolution paleoenvironmental interpretation of the rock record, p. 4985. InMacEachern, J. A., Bann, K. L., Gingras, M. K., and Pemberton, S. G.(eds.), Applied Ichnology, Society for Sedimentary Geology Special Publication 83.Google Scholar
MacEachern, J. A., Bann, K. L., Bhattacharya, J. P., and Howell, C. D. 2005. Ichnology of deltas: organism responses to the dynamic interplay of rivers, waves, storms, and tides, p. 4985. InGiosan, L. and Bhattacharya, J. P.(eds.), River deltas: Concepts, Models, and Examples, SEPMSpecial Publication. SEPM (Society for Sedimentary Geology), Tulsa, Oklahoma.Google Scholar
Maklhlouf, I. M. 2003. Braided river model and associated facies of lower Cambrian age in South Jordan. Africa Geoscience Review, 10:289300.Google Scholar
Maklhlouf, I. M. and A. M. Abed 1991. Depositional facies and environments in the Umm Ishrin Sandstone Formation, Dead Sea area, Jordan, Sedimentary Geology, 71:177187.Google Scholar
Mángano, M. G. and Buatois, L. A. 2003 a. Rusophycus leifeirikssoni en la Formación Campanario: Implicancias paleobiológicas y paleoambientales, p. 6584. InBuatois, L. A. and Mángano, M. G.(eds.), Icnología: Hacia una convergencia entre geología y biología. Volume 9. Publicación Especial de la Asociación Paleontológica, Argentina.Google Scholar
Mángano, M. G. and Buatois, L. A. 2003 b. Trace fossils, p. 507553. InBenedetto, J. L.(ed.), Ordovician Fossils of Argentina. Universidad Nacional de Córdoba, Secretaría de Ciencia y Tecnología.Google Scholar
Mángano, M. G. and Buatois, L. A. 2004. Reconstructing early Phanerozoic intertidal ecosystems: Ichnology of the Cambrian Campanario Formation in northwest Argentina. Fossils and Strata, 51:122.Google Scholar
Mángano, M. G., Buatois, L. A., and Aceñolaza, G. F. 1996. Trace fossils and sedimentary facies from a late Cambrian–Early Ordovician tide-dominated shelf (Santa Rosita Formation, northwest Argentina): Implications for ichnofacies models of shallow marine successions. Ichnos, 5:5388.Google Scholar
Mángano, M. G., Buatois, L. A., and Guinea, F. M. 2005. Ichnology of the Alfarcito Member (Santa Rosita Formation) of northwestern Argentina: Animal-substrate interactions in a lower Paleozoic wave-dominated shallow sea. Ameghiniana, 42:641668.Google Scholar
Mángano, M. G., Buatois, L. A., Maples, C. G., and West, R. R. 2002. Ichnology of a Pennsylvanian equatorial tidal flat; the Stull Shale Member at Waverly, eastern Kansas. Kansas Geological Survey 245, 133p.Google Scholar
Mángano, M. G., Buatois, L. A., West, R. R., and Maples, C. G. 1999. The origin and paleoecologic significance of the trace fossil Asteriacites in the Pennsylvanian of Kansas and Missouri. Lethaia, 32:1730.Google Scholar
Maples, C. G. and Archer, A. W. 1989. The potential of Paleozoic nonmarine trace fossils for paleoecological interpretations. Palaeogeography, Palaeoclimatology, Palaeoecology, 73:185195.Google Scholar
Martinsson, A. 1965. Aspects of a middle Cambrian thanatotope on Öland. Geologiska Foereningan i Stockholm. Foerhandlingar, 87:181230.Google Scholar
McCarthy, B. 1979. Trace fossils from a Permian shoreface-foreshore environment, eastern Australia. Journal of Paleontology, 53:345366.Google Scholar
Miller, S. A. 1889. North American geology and palæontology for the use of amateurs, students, and scientists. Western Methodist book concern, Cincinnati, Ohio, 664p.Google Scholar
Miller, M. F. and Smail, S. E. 1997. A semiquantitative field method for evaluating bioturbation on bedding planes. PALAIOS, 12:391396.Google Scholar
Minter, N. J., Krainer, K., Lucas, S. G., Braddy, S. J., and Hunt, A. P. 2007. Palaeoecology of an Early Permian playa lake trace fossil assemblage from Castle Peak, Texas, U.S.A. Palaeogeography, Palaeoclimatology, Palaeoecology, 246:390423.Google Scholar
Müller, A. H. 1971. Zur Kenntnis von Asterosoma (Vestigia invertebratorum). Freiberg Forschungshefte C, 267:717.Google Scholar
Nara, M. 1995. Rosselia socialis—a dwelling structure of a probable terebellid polychaete. Lethaia, 28:171178.Google Scholar
Nara, M. 1997. High-resolution analytical method for event sedimentation using Rosselia socialis. PALAIOS, 12:489494.Google Scholar
Nara, M. 2002. Crowded Rosselia socialis in Pleistocene inner shelf deposits: Benthic paleoecology during rapid sea-level rise. PALAIOS, 17:268276.2.0.CO;2>CrossRefGoogle Scholar
Neto de Carvalho, C. 2006. Roller coaster behavior in the Cruziana rugosa Group from Penha Garcia (Portugal): Implications for the feeding program of trilobites. Ichnos, 13:255265.Google Scholar
Neto de Carvalho, C. and Rodrigues, N. P. C. 2007. Compound Asterosoma ludwigae Schlirf, 2000 from the Jurassic of the Lusitanian Basin (Portugal): Conditional strategies in the behaviour of Crustacea. Journal of Iberian Geology, 33:295310.Google Scholar
Netto, R. G., Buatois, L. A., Mángano, M. G., and Alisteri, P. 2007. Gyrolithes as a multipurpose burrow: An ethologic approach. Revista Brasileira de Paleontologia, 10:157168.Google Scholar
Nicholson, H. A. 1873. Contributions to the study of the errant annelids of the older Palaeozoic rocks. Royal Society of London Proceedings, 21:288290.Google Scholar
D‘Orbigny, A. D. 1842. Voyages darts l'Amérique méridionale—le Bresil, la République orientale d'Uruguay, la République Argentine, la Patagonie, la République du Chili, la République de Bolivia, la republique du Pérou—éxécuté pendant les annés 1826, 1827, 1828, 1829, 1830, 1831, 1832, et 1833. Pitois-Levrault, Ve Levrault, Paris, Strasbourg, 3.Google Scholar
Orlowski, S. 1992. Trilobite trace fossils and their stratigraphical significance in the Cambrian sequence of the Holy Cross Mountains, Poland. Geological Journal, 27:1534.CrossRefGoogle Scholar
Osgood, R. G. J. 1970. Trace fossils of the Cincinnati Area. Palaeontographica Americana, 6 (41):277444.Google Scholar
Otto, E. von 1854. Additamente zur Flora des Quadergebirges in Sachsen. Teil 2. G. Mayer, Leipzig, 53p.Google Scholar
Pemberton, S. G. and Frey, R. W. 1982. Ichnological nomenclature and the Palaeophycus-Planolites dilemma. Journal of Paleontology, 56:843881.Google Scholar
Pemberton, S. G., Frey, R. W., and Bromley, R. G. 1988. The ichnotaxonomy of Conostichus and other plug-shaped ichnofossils. Canadian Journal of Earth Sciences, 25:866892.Google Scholar
Pemberton, S. G., Spila, M., Pulham, A. J., Saunders, T., MacEachern, J. A., Robbins, D., and Sinclair, I. K. 2001. Ichnology and sedimentology of shallow to marginal marine systems: Ben Nevis and Avalon Reservoirs, Jeanne d'Arc Basin. Geological Association of Canada, Short Course Notes, St. John's, Newfoundland, 15:1343.Google Scholar
Pickerill, R. K., Fyffe, L. R., and Forbes, W. H. 1987. Late Ordovician–Early Silurian trace fossils from the Matapedia Group, Tobique River, western New Brunswick. Atlantic Geology, 23:7788.Google Scholar
Pickerill, R. K., Romano, M., and Meléndez, B. 1984. Arenig trace fossils from the Salamanca area, western Spain. Geological Journal, 19:249269.Google Scholar
Plummer, P. S. and Gostin, V. A. 1981. Shrinkage cracks; desiccation or synaeresis? Journal of Sedimentary Research, 51:11471156.Google Scholar
Pollard, J. E., Goldring, R., and Buck, S. G. 1993. Ichnofabrics containing Ophiomorpha—significance in shallow-water facies interpretation. Journal of the Geological Society, 150:149164.Google Scholar
Powell, E. N. 1977. Relationship of trace fossil Gyrolithes (=Xenohelix) to Family Capitellidae (Polychaeta). Journal of Paleontology, 51:552556.Google Scholar
Powell, J. H. 1989. Stratigraphy and sedimentation of the Phanerozoic rocks in Central and South, Jordan, Part A: Ram and Khreim Groups. Bulletin, 11. Natural Resources Authority, Geology Directorate, Geological Mapping Division: 172.Google Scholar
Prantl, F. 1945. Two new problematic trails from the Ordovician of Bohemia. Academie tcheque des sciences, Bulletin international, Classe des sciences mathematiques et naturalles et de la medicine, 46:4959.Google Scholar
Richter, R. 1850. Aus der thüringischen Grauwacke. Deutsche Geologische Gesellschaft, Zeitschrift 2:198206.Google Scholar
Richter, R. 1937. Marken und Spuren aus allen Zeiten. I–II. Senckenbergiana, 19:150163.Google Scholar
Rushton, A. W. A. and Powell, J. H. 1998. A review of the stratigraphy and trilobite faunas from the Cambrian Burj Formation in Jordan. Bulletin of the Natural History Museum, Geology, 54:131146.Google Scholar
Salter, J. W. 1857. On annelide-burrows and surface markings from the Cambrian rocks of the Longmynd and North Wales. Quarterly Journal of the Geological Society of London, 13:199205.Google Scholar
Saporta, G. de. 1884. Les organismes problématiques des anciennes mers, Paris, 100p.Google Scholar
Schatz, E. R., Mángano, M. G., Buatois, L. A., and Limarino, C. O. 2011. Life in the Late Paleozoic Ice Age: Trace fossils from glacially influenced deposits in a Late Carboniferous fjord of western Argentina. Journal of Paleontology, 85:502518.Google Scholar
Schindewolf, O. H. 1928. Studien aus dem Marburger Buntsandstein III–VI. Senckenbergiana, 10:1654.Google Scholar
Schlirf, M. and Uchman, A. 2005. Revision of the Ichnogenus Sabellarifex Richter, 1921 and its relationship to Skolithos Haldeman, 1840 and Polykladichnus Fürsich, 1981. Journal of Systematic Palaeontology, 3:115131.Google Scholar
Schlirf, M., Uchman, A., and Kümmel, M. 2001. Upper Triassic (Keuper) non-marine trace fossils from the Haβberge area (Franconia, south-eastern Germany). Paläontologische Zeitschrift, 75:7196.Google Scholar
Schmalfuss, H. 1981. Structure, pattern, and function of cuticular terraces in trilobites. Lethaia, 14:331341.Google Scholar
Schneider, W., Amireh, B. S., and Abed, A. M. 2007. Sequence analysis of the early Paleozoic sedimentary systems of Jordan. Zeitschrift der Deutschen Gesellschaft fur Geowissenschaften, 158:225247.Google Scholar
Seilacher, A. 1953. Studien zur Palichnologie. II. Die fossilen Ruhespuren (Cubichnia). Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen, 98:87124.Google Scholar
Seilacher, A. 1955. Spuren und Lebensweise der Trilobiten, p. 86116. InSchindewolf, O. H.(ed.), Beiträge zur Kenntnis des Kambriums in der Salt Range (Pakistan). Akademie der Wissenschaften und der Literatur in Mainz, Wiesbaden.Google Scholar
Seilacher, A. 1967. Bathymetry of trace fossils. Marine Geology, 5:413428.Google Scholar
Seilacher, A. 1970. Cruziana stratigraphy of “nonfossiliferous” Palaeozoic sandstones, p. 447476. InCrimes, T. P. and Harper, J. C.(eds.), Trace Fossils, Geological Journal Special Issue. Volume 3.Google Scholar
Seilacher, A. 1985. Trilobite palaeobiology and substrate relationships. Transactions of the Royal Society of Edinburgh Earth Sciences, 76:231238.Google Scholar
Seilacher, A. 1990. Chapter 32: Paleozoic trace fossils, p. 649722. InSaid, R.(ed.), The geology of Egypt. A. A. Balkema, Rotterdam, Brookfield, VT.Google Scholar
Seilacher, A. 1992. An updated Cruziana stratigraphy of Gondwanan Paleozoic sandstones, p. 15651581. InSalem, M. J.(ed.), The Geology of Lybia. Volume part 8. Elsevier, Amsterdam.Google Scholar
Seilacher, A. 2000. Ordovician and Silurian arthrophycid ichnostratigraphy, p. 237258. InSola, M. A., and Worsley, D.(eds.), Geological Exploration in Murzuk Basin, Elsevier, Amsterdam.Google Scholar
Seilacher, A. 2005. Silurian trace fossils from Africa and South America mapping a trans-Gondwanan seaway. Neues Jahrbuch für Geololgie und Paläontologie–Monatshefte, 2005:129141.Google Scholar
Seilacher, A. 2007. Trace Fossil Analysis. Springer, Berlin, 226p.Google Scholar
Seilacher, A., Buatois, L. A., and Mángano, M. G. 2005. Trace fossils in the Ediacaran–Cambrian transition: behavioral diversification, ecological turnover and environmental shift. Palaeogeography, Palaeoclimatology, Palaeoecology, 227:323356.CrossRefGoogle Scholar
Sharland, P. R., Archer, R., Casey, D. M., Davies, R. B., Hall, S. H., Heward, A. P., Horbury, A. D., and Simmons, M. D. 2001. Arabian plate sequence stratigraphy. GeoArabia Special Publications 2, p. 371.Google Scholar
Shinaq, R. and Bandel, K. 1992. Microfacies of Cambrian limestones in Jordan. Facies, 27:263283.Google Scholar
Shinaq, R. and Elicki, O. 2007. The Cambrian sedimentary succession from the Wadi Zerqa Ma'in (northeastern Dead Sea area, Jordan): Lithology and fossil content. Neues Jahrbuch für Geologie und Palaontologie, Abhandlungen, 243:255271.Google Scholar
Taylor, A. M. and Goldring, R. 1993. Description and analysis of bioturbation and ichnofabric. Journal of the Geological Society, 150:141148.Google Scholar
Teimeh, M., Taani, Y., Aulithie, O., and Saad, L. Abu. 1990. A study of the Palaeozoic formations of Jordan at outcrop and in the subsurface including measured sections and regional isopach maps. Natural Resources Authority, Geology Directorate, Subsurface Geology Division, Bulletin, Amman, 1:172.Google Scholar
Torell, O. M. 1870. Petrificata Suecana Formationis Cambricae. Lunds Universitets Årskrift, 6:114.Google Scholar
Uchman, A., Kazakauskas, V., and Gaigalas, A. 2009. Trace fossils from late Pleistocene varved lacustrine sediments in eastern Lithuania. Palaeogeography, Palaeoclimatology, Palaeoecology, 272:199211.Google Scholar
Vannier, J., Calandra, I., Gaillard, C., and Zylinska, A. 2010. Priapulid worms: pioneer horizontal burrowers at the Precambrian–Cambrian boundary. Geology, 38:711714.CrossRefGoogle Scholar
Walker, E. F. 1985. Arthropod ichnofauna of the Old Red sandstone at Dunure and Montrose, Scotland. Transactions of the Royal Society of Edinburgh, Earth Sciences, 76:287297.Google Scholar
Webby, B. D. 1983. Lower Ordovician arthropod trace fossils from western New South Wales. Proceedings of the Linnean Society of New South Wales, 107:5974.Google Scholar
Weiss, A. 1958. Die Innerkristalline Quellung als Allgemeines Modell fur Quellungsvorgänge. Chemische Berichte, 91:481502.Google Scholar
Westergård, A. H. 1931. Diplocraterion, Monocraterion and Scolithus from the lower Cambrian of Sweden. Sveriges Geologiska Undersköping, serie C, Avhandlingar och uppsatser, 372:115.Google Scholar
Wetzel, A., Tjallingii, R., and Stattegger, K. 2010. Gyrolithes in Holocene estuarine incised-valley fill deposits, offshore southern Vietnam. PALAIOS, 25:239246.Google Scholar
Yochelson, E. L. and Fedonkin, M. A. 1997. The type specimens (middle Cambrian) of the trace fossil Archaeonassa Fenton and Fenton. Canadian Journal of Earth Sciences, 34:12101219.CrossRefGoogle Scholar