Hostname: page-component-8448b6f56d-cfpbc Total loading time: 0 Render date: 2024-04-23T11:39:47.569Z Has data issue: false hasContentIssue false

Ichnofauna from coastal meandering channel systems (Upper Cretaceous Tremp Formation, South-Central Pyrenees, Spain): delineating the fluvial-tidal transition

Published online by Cambridge University Press:  01 July 2016

Davinia Díez-Canseco
Affiliation:
Departamento de Estratigrafía, Universidad Complutense de Madrid - Instituto de Geociencias, Madrid E28040, Spain 〈daviniadiezcanseco@ucm.es〉, 〈margot@geo.ucm.es〉, 〈mibenito@ucm.es〉 Catalonia Geological Research - Geoplay, Tremp E25620, Spain
Luis A. Buatois
Affiliation:
Department of Geological Sciences, University of Saskatchewan, Saskatchewan, Saskatoon SK S7N 5E2, Canada 〈luis.buatois@usask.ca〉, 〈gabriela.mangano@usask.ca〉
M. Gabriela Mángano
Affiliation:
Department of Geological Sciences, University of Saskatchewan, Saskatchewan, Saskatoon SK S7N 5E2, Canada 〈luis.buatois@usask.ca〉, 〈gabriela.mangano@usask.ca〉
Margarita Díaz-Molina
Affiliation:
Departamento de Estratigrafía, Universidad Complutense de Madrid - Instituto de Geociencias, Madrid E28040, Spain 〈daviniadiezcanseco@ucm.es〉, 〈margot@geo.ucm.es〉, 〈mibenito@ucm.es〉
M. Isabel Benito
Affiliation:
Departamento de Estratigrafía, Universidad Complutense de Madrid - Instituto de Geociencias, Madrid E28040, Spain 〈daviniadiezcanseco@ucm.es〉, 〈margot@geo.ucm.es〉, 〈mibenito@ucm.es〉

Abstract

The Upper Cretaceous “redbeds” of the lower Tremp Formation (South-Central Pyrenees, Spain) contains an ichnofauna consisting of Taenidium barretti, Taenidium bowni, Loloichnus isp., Arenicolites isp., Planolites isp., and Palaeophycus isp. This ichnofauna occurs in deposits formed in tide-influenced meander loops and their associated overbank mudflats. Evaluation of the taphonomic controls on the Tremp ichnofauna shows that (1) two morphotypes of Taenidium barretti are controlled by the substrate consistence, (2) Arenicolites may be enlarged by erosion processes, and (3) Taenidium barretti and Planolites isp. are not the same ichnotaxa showing different types of preservation. The meniscate fill in Taenidium barretti suggests that this structure was produced by deposit feeders. The Tremp ichnofauna is grouped into two trace-fossil assemblages, a depauperate subaquatic monospecific Planolites suite and an assemblage representing the Scoyenia Ichnofacies. Trace-fossil distribution reflects paleoenvironmental changes in the meandering channels along the stratigraphic section with the Planolites suite in the lowermost part of the lower interval and the Scoyenia Ichnofacies in the middle and upper intervals. The lowermost suite may be likely formed seaward of the maximum salinity limit, under extreme brackish-water conditions, whereas the Scoyenia Ichnofacies records a freshwater assemblage that was formed landward of the maximum salinity limit, reflecting deltaic progradation.

Type
Articles
Copyright
Copyright © 2016, 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

Alonso-Zarza, A.M., 2003, Palaeoenvironmental significance of palustrine carbonates and calcretes in the geological record: Earth-Science Reviews, v. 60, p. 261298.Google Scholar
Ansell, A.D., 1962, Observations on burrowing in the Veneridae (Eulamellibranchia): Biological Bulletin, Woods Hole, v. 123, p. 521530.CrossRefGoogle Scholar
Astre, G., 1937, Un annélide sabellien dans le Garumnien de Saldes: Bulletin de la Société d’Histoire Naturalle de Toulouse, v. 71, p. 192194.Google Scholar
Baceta, J.I., Pujalte, V., Serra-Kiel, J., Robador, A., and Orue-Etxebarria, X., 2004, El Maastrichtiense final, Paleoceno e Ilerdiense inferior de la Cordillera Pirenaica, in Vera, J.A., ed., Geología de España, Madrid, Sociedad Geológica de España-Instituto Geológico y Minero de España, p. 308313.Google Scholar
Baucon, A., and Felletti, F., 2013, Neoichnology of a barrier-island system: the Mula di Muggia (Grado lagoon, Italy): Palaeogeography, Palaeoclimatology, Palaeoecology, v. 375, p. 112124.Google Scholar
Bedatou, E., Melchor, R.N., Bellosi, E., and Genise, J.F., 2008, Crayfish burrows from Late Jurassic–Late Cretaceous continental deposits of Patagonia: Argentina. Their palaeoecological, palaeoclimatic and palaeobiogeographical significance: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 257, p. 169184.Google Scholar
Bottjer, D.J., and Droser, M.L., 1991, Ichnofabric and basin analysis: Palaios, v. 6, p. 199205.Google Scholar
Bown, T.M., and Kraus, M.J., 1983, Ichnofossils of the alluvial Willwood Formation (lower Eocene), Bighorn basin, northwest Wyoming, USA: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 43, p. 95128.CrossRefGoogle Scholar
Bradshaw, M.A., 1981, Paleoenvironmental interpretations and systematics of Devonian trace fossils from the Taylor Group (lower Beacon Supergroup), Antarctica: New Zealand Journal of Geology and Geophysics, v. 24, p. 615652.Google Scholar
Brady, L.F., 1947, Invertebrate tracks from the Coconino Sandstone of northern Arizona: Journal of Paleontology, v. 21, p. 466472.Google Scholar
Bromley, R.G., 1996, Trace fossils. Biology, Taphonomy and Applications, (second edition): London, UK, Chapman and Hall, 361 p.Google Scholar
Bromley, R.G., and Ekdale, A., 1984, Trace fossil preservation in flint in the European chalk: Journal of Paleontology, v. 58, no. 2, p. 298311.Google Scholar
Buatois, L.A., and Mángano, M.G., 1995, The paleoenvironmental and paleoecological significance of the lacustrine Mermia ichnofacies: an archetypical subaqueous nonmarine trace fossil assemblage: Ichnos, v. 4, p. 151161.CrossRefGoogle 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, v. 183, p. 7186.Google Scholar
Buatois, L.A., and Mángano, M.G., 2004, Animal-substrate interactions in freshwater environments: applications of ichnology in facies and sequence stratigraphic analysis of fluvio-lacustrine successions, in McIlroy, D., ed., The Application of Ichnology to Palaeoenvironmental and Stratigraphic Analysis: Geological Society Special Publication, v. 228, p. 311333.Google Scholar
Buatois, L.A., and Mángano, M.G., 2011, Ichnology: Organism-substrate interactions in space and time, New York, USA, Cambridge University Press, 358 p.Google Scholar
Buatois, L.A., Mángano, M.G., and Aceñolaza, F.G., 1996, Icnofaunas paleozoicas en sustratos firmes no marinos: evidencias del Pérmico de la cuenca de Paganzo: Ameghiniana, v. 33, p. 265270.Google Scholar
Buatois, L.A., Mangano, M.G., and Maples, C.G., 1997, The paradox of nonmarine ichnofaunas in tidal rhythmites; integrating sedimentologic and ichnologic data from the Late Cretaceous of eastern Kansas, USA: Palaios, v. 12, p. 467481.Google Scholar
Buatois, L.A., Mangano, 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, USA: Journal of Paleontology, v. 72, p. 152180.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, v. 20, p. 321347.Google Scholar
Buatois, L.A., Uba, C.E., Mángano, M.G., Hulka, C., and Heubeck, C., 2007, Deep and intense bioturbation in continental environments: evidence from Miocene fluvial deposits of Bolivia, in Bromley, R.G., Buatois, L.A., Mángano, M.G., and Genise, J.F., eds., Sediment-Organism Interactions: A Multifaceted Ichnology: Society for Sedimentary Geology Special Publication, v. 88, p. 123136.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, v. 273, p. 7586.Google Scholar
Carmona, N.B., Mángano, M.G., Buatois, L.A., and Ponce, J.J., 2010, Taphonomy and paleoecology of the bivalve trace fossil Protovirgularia in deltaic heterolithic facies of the Miocene Chenque formation, Patagonia, Argentina: Journal of Paleontology, v. 84, p. 730738.Google Scholar
Counts, J.W., and Hasiotis, S.T., 2009, Neoichnological experiments with masked chafer beetles (coleoptera: scarabaeidae): implications for backfilled continental trace fossils: Palaios, v. 24, p. 7491.Google Scholar
Crandall, K.A., and Buhay, J.E., 2008, Global diversity of crayfish (Astacidae, Cambaridae, and Parastacidae-Decapoda) in freshwater: Hydrobiologia, v. 595, no. 1, p. 295301.CrossRefGoogle Scholar
Cuevas, J.L., 1992, Estratigrafía del” Garumniense” de la Conca de Tremp. Prepirineo de Lérida: Acta Geológica Hispánica, v. 27, p. 95108.Google Scholar
D’Alessandro, A., and Bromley, R.G., 1987, Meniscate trace fossils and the Muensteria-Taenidium problem: Palaeontology, v. 30, p. 743763.Google Scholar
D’Alessandro, A., Ekdale, A., and Picard, M.D., 1987, Trace fossils in fluvial deposits of the Duchesne River Formation (Eocene), Uinta basin, Utah: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 61, p. 285301.Google Scholar
Dafoe, L.T., Gingras, M.K., and Pemberton, S.G., 2008, Analysis of mineral segregation in Euzonus mucronata burrow structures: one possible method used in the construction of ancient Macaronichnus segregates: Ichnos, v. 15, p. 91102.Google Scholar
Dalrymple, R.W., and Choi, K., 2007, Morphologic and facies trends through the fluvial-marine transition in tide-dominated depositional systems: a schematic framework for environmental and sequence-stratigraphic interpretation: Earth-Science Reviews, v. 81, p. 135174.Google Scholar
Dashtgard, S.E., 2011, Linking invertebrate burrow distributions (neoichnology) to physicochemical stresses on a sandy tidal flat: implications for the rock record: Sedimentology, v. 58, p. 13031325.Google Scholar
Dashtgard, S.E., Venditti, J.G., Hill, P.R., Sisulak, C.F., Johnson, S.M., and La Croix, A.D., 2012, Sedimentation across the tidal-fluvial transition in the lower Fraser River, Canada: The Sedimentary Record, v. 10, p. 49.Google Scholar
Díez-Canseco, D., Arz, J., Benito, M., Díaz-Molina, M., and Arenillas, I., 2014, Tidal influence in redbeds: A palaeoenvironmental and biochronostratigraphic reconstruction of the Lower Tremp Formation (South-Central Pyrenees, Spain) around the Cretaceous/Paleogene boundary: Sedimentary Geology, v. 312, p. 3149.Google Scholar
Dörjes, J., and Howard, J.D., 1975, Estuaries of the Georgia coast, U.S.A.: Sedimentology and biology. IV. Fluvial-marine transition indicators in an estuarine environment, Ogeechee River–Ossabaw Sound: Senckenbergiana Maritima, v. 7, p. 137179.Google Scholar
Ekdale, A.A., 1985, Paleoecology of the marine endobenthos: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 50, p. 6381.Google Scholar
Ekdale, A.A., and Bromley, R.G., 1991, Analysis of composite ichnofabrics: an example in uppermost Cretaceous chalk of Denmark: Palaios, v. 6, p. 232249.Google Scholar
Esteban, M., and Klappa, C.F., 1983, Subaerial exposure environments, in Scholle, P.A., Bebout, D.G., and Moore, C.H., eds., Carbonate Depositional Environments: Association of Petroleum Geologists Memoir, v. 33, p. 196.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, Canada: Palaeontographica Canadiana, v. 7, p. 1119.Google Scholar
Frey, R.W., and Pemberton, S.G., 1984, Trace fossils Facies Models, in Walker, R.G., ed., Facies Models: Geoscience Canada Reprint Series, p. 189207.Google Scholar
Frey, R.W., and Pemberton, S.G., 1987, The Psilonichnus ichnocoenose, and its relationship to adjacent marine and nonmarine ichnocoenoses along the Georgia coast: Bulletin of Canadian Petroleum Geology, v. 35, p. 333357.Google Scholar
Frey, R.W., Curran, H.A., and Pemberton, S.G., 1984a, Tracemaking activities of crabs and their environmental significance: the ichnogenus Psilonichnus: Journal of Paleontology, v. 58, p. 333350.Google Scholar
Frey, R.W., Pemberton, S.G., and Fagerstrom, J., 1984b, Morphological, ethological, and environmental significance of the ichnogenera Scoyenia and Ancorichnus: Journal of Paleontology, v. 58, p. 511528.Google Scholar
Freytet, P., 1973, Petrography and paleo-environment of continental carbonate deposits with particular reference to the Upper Cretaceous and Lower Eocene of Languedoc (southern France): Sedimentary Geology, v. 10, no. 1, p. 2560.Google Scholar
Fürsich, F.T., 1972, Thalassinoides and the origin of nodular limestone in the Corallian beds (Upper Jurassic) of southern England: Neues Jahrbuch für Geologie, und Paläontologie, v. 3, p. 136156.Google Scholar
Fürsich, F.T., 1974, On Diplocraterion Torell 1870 and the significance of morphological features in vertical, spreiten-bearing, U-shaped trace fossils: Journal of Paleontology, v. 48, no. 5, p. 952962.Google Scholar
Fürsich, F.T., 1981, Invertebrate trace fossils from the Upper Jurassic of Portugal: Comunicaçôes dos Serviços Geológicos de Portugal, v. 67, p. 153168.Google Scholar
Gerdol, V., and Hughes, R.G., 1994, Feeding behaviour and diet of Corophium volutator in an estuary in southeastern England: Marine Ecology Progress Series, v. 114, p. 103108.Google Scholar
Gibert, J.M., and Domènech, R., 2008, Trazas fósiles de nuculoideos (Protovirgularia) del Mioceno marino de la Cuenca del Vallès-Penedès: Revista Española de Paleontología, v. 23, p. 129138.Google Scholar
Gingras, M.K., Pemberton, S.G., Saunders, T., and Clifton, H.E., 1999, The ichnology of modern and Pleistocene brackish-water deposits at Willapa Bay, Washington; variability in estuarine settings: Palaios, v. 14, p. 352374.Google Scholar
Gingras, M.K., Räsänen, M., and Ranzi, A., 2002, The significance of bioturbated inclined heterolithic stratification in the southern part of the Miocene Solimoes Formation, Rio Acre, Amazonia Brazil: Palaios, v. 17, p. 591601.Google Scholar
Goldring, R., and Pollard, J., 1996, Ichnotaxonomic revision and the importance of type material: Palaeontology Newsletter, v. 31, p. 78.Google Scholar
Goldring, R., Pollard, J., and Taylor, A., 1997, Naming trace fossils: Geological Magazine, v. 134, no. 2, p. 265268.Google Scholar
Gregory, M.R., Martin, A.J., and Campbell, K.A., 2004, Compound trace fossils formed by plant and animal interactions: quaternary of northern New Zealand and Sapelo Island, Georgia (USA): Fossils and Strata, v. 51, p. 88105.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, Philadelphia, J. Dobson.Google Scholar
Hall, J., 1847, Palaeontology of New York, Albany, C. Van Benthuysen, v. 1, 338 p.Google Scholar
Häntzschel, W., 1975, Trace fossils & Problematica, in Teichert, C., ed., Treatise on Invertebrate Paleontology, Part W, Miscellanea, 2d ed.: Boulder, Colorado, Geological Society of America (and University of Kansas Press), 269 p.Google Scholar
Hasiotis, S.T., 2004, Reconnaissance of Upper Jurassic Morrison Formation ichnofossils, Rocky Mountain Region, USA: paleoenvironmental, stratigraphic, and paleoclimatic significance of terrestrial and freshwater ichnocoenoses: Sedimentary Geology, v. 167, p. 177268.Google Scholar
Hasiotis, S.T., and Mitchell, C.E., 1993, A comparison of crayfish burrow morphologies: Triassic and Holocene fossil, paleo-and neo-ichnological evidence, and the identification of their burrowing signatures: Ichnos, v. 2, p. 291314.Google Scholar
Hasiotis, S.T., Mitchell, C.E., and Dubiel, R.F., 1993, Application of morphologic burrow interpretations to discern continental burrow architects: lungfish or crayfish?: Ichnos, v. 2, p. 315333.CrossRefGoogle Scholar
Heer, O., 1876–77, Flora fossilis Hevetiae. Die vorweltliche Flora der Schweiz: J. Würster & Co, 182 p.Google Scholar
Heinberg, C., 1974, A dynamic model for a meniscus filled tunnel (Ancorichnus n. ichnogen.) from the Jurassic Pecten Sandstone of Milne Land, East Greenland: Grønlands Geologiske Undersøgelse Rapport, v. 62, 20 p.Google Scholar
Hovikoski, J., Gingras, M.K., Räsänen, M., Rebata, L.A., Guerrero, J., Ranzi, A., Melo, J., Romero, L., del Prado, H.N., and Jaimes, F., 2007, The nature of Miocene Amazonian epicontinental embayment: High-frequency shifts of the low-gradient coastline: Geological Society of America Bulletin, v. 119, p. 15061520.CrossRefGoogle Scholar
Hovikoski, J., Lemiski, R., Gingras, M.K., Pemberton, S.G., and MacEachern, J.A., 2008a, Ichnology and sedimentology of a mud-dominated deltaic coast: Upper Cretaceous Alderson Member (Lea Park Fm), western Canada: Journal of Sedimentary Research, v. 78, p. 803824.Google Scholar
Hovikoski, J., Räsänen, M., Gingras, M.K., Ranzi, A., and Melo, J., 2008b, Tidal and seasonal controls in the formation of Late Miocene inclined heterolithic stratification deposits, western Amazonian foreland basin: Sedimentology, v. 55, p. 499530.CrossRefGoogle Scholar
Howard, J.D., and Frey, R.W., 1975, Estuaries of the Georgia coast, U.S.A.: Sedimentology and biology, II. Regional animal-sediment characteristics of Georgia estuaries: Senckenbergiana Maritima, v. 7, p. 33103.Google Scholar
Howard, J.D., Elders, C.A., and Heinbokel, J.F., 1975, Estuaries of the Georgia Coast, U.S.A.: Sedimentology and biology, V. Animal-sediment relationships in estuarine point bar deposits, Ogeechee River–Ossabaw Sound: Senckenbergiana Maritima, v. 7, p. 181203.Google Scholar
Johnson, S.M., and Dashtgard, S.E., 2014, Inclined heterolithic stratification in a mixed tidal-fluvial channel: Differentiating tidal versus fluvial controls on sedimentation: Sedimentary Geology, v. 301, p. 4153.Google Scholar
Keighley, D., and Pickerill, R., 1994, The ichnogenus Beaconites and its distinction from Ancorichnus and Taenidium: Palaeontology, v. 37, p. 305338.Google Scholar
Kluijver, M.J., and Ingalsuo, S.S., 1999, Macrobenthos of the North Sea-Crustacea, Marine Species Indentification Portal. http://www.species-identification.org/species.php?species_group=crustacea&id=303&menuentry=soorten (accessed November 2014).Google Scholar
Krapovickas, V., Ciccioli, P.L., Mángano, M.G., Marsicano, C.A., and Limarino, C.O., 2009, Paleobiology and paleoecology of an arid–semiarid Miocene South American ichnofauna in anastomosed fluvial deposits: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 284, p. 129152.Google Scholar
Krapovickas, V., Mancuso, A.C., Arcucci, A.B., and Caselli, A.T., 2010, Fluvial and eolian ichnofaunas from the Lower Permian of South America (Patquía Formation, Paganzo Basin): Geologica Acta, v. 8, no. 4, p. 449462.Google Scholar
Larsen, R.J., and Marx, M.L., 1990, Statistics, Upper Saddle River, New Jersey, Prentice Hall, 800 p.Google Scholar
Lewis, D.W., and Ekdale, A.A., 1992, Composite ichnofabric of a mid-Tertiary unconformity on a pelagic limestone: Palaios, v. 7, p. 222235.Google Scholar
Linck, O., 1949, Lebens-Spuren aus dem Schilfsandstein (Mittl. Keuper km 2) NW-Württemberfs und ihre Bedeutung für die Bildungsgeschichte der Stufe: Jahreshefte des Vereins für vaterländische Naturkunde in Württemberg, v. 97–101, p. 1100.Google Scholar
López-Martínez, N., 2001, La extinción de los dinosaurios y su registro en los Pirineos Meridionales: Actas de las II Jornadas Internacionales sobre Paleontología de Dinosaurios y su Entorno, Salas de los Infantes, Spain, Colectivo Arqueológico-Paleontológico de Salas, p. 7198.Google Scholar
López-Martínez, N., Arribas, M., Robador, A., Vicens, E., and Ardèvol, L., 2006, Los carbonatos Danienses (Unidad 3) de la Fm Tremp (Pirineos Sur-Centales): Paleogeografía y relación con el límite Cretácico-Terciario: Revista de la Sociedad Geológica de España, v. 19, p. 233255.Google Scholar
MacEachern, J.A., and Gingras, M.K., 2007, Recognition of brackish-water trace fossil suites in the Cretaceous Western interior seaway of Alberta, Canada, in Bromley, R.G., Buatois, L.A., Mangano, G., Genise, J.F., and Melchor, R.N., eds., Sediment-Organism Interactions: A Multifacted Ichnology: Society for Sedimentary Geology Core Workshop, v. 88, p. 149194.Google Scholar
MacEachern, J.A., and Pemberton, S.G., 1994, Ichnological aspects of incised valley fill systems from the Viking Formation of the Western Canada Sedimentary Basin, Alberta, Canada, in Boyd, R., Zaitlin, B.A., and Dalrymple, R.W., eds., Incised Valley Systems: Origin and Sedimentary Sequences: Society for Sedimentary Geology Special Publication, v. 51, p. 129157.Google Scholar
Mángano, M.G., and Buatois, L.A., 2004, Ichnology of Carboniferous tide-influenced environments and tidal flat variability in the North American Midcontinent, in McIlroy, D., ed., The application of ichnology to palaeoenvironmental and stratigraphic analysis, London, UK, Geological Society Special Publications, v. 228, p. 157178.Google Scholar
Mángano, M.G., Buatois, L.A., West, R.R., and Maples, C.G., 1998, Contrasting behavioral and feeding strategies recorded by tidal-flat bivalve trace fossils from the Upper Carboniferous of eastern Kansas: Palaios, v. 13, p. 335351.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, v. 245, p. 1133.Google Scholar
Maples, C.G., and West, R.R., 1990, Trace-fossil taxonomy and substrate fluidity: An example using Late Carboniferous bivalve traces: 13th International Sedimentological Congress, Abstracts of Papers, 328 p.Google Scholar
Mayoral, E., and Calzada., S, 1998, Reinterpretación de Spirographites ellipticus Astre, 1937 como una pista fósil de Artrópodos no marinos en el Cretácico superior (facies Garumn) del Prepirineo Catalán (NE de España): Geobios, v. 31, p. 633643.Google Scholar
McIlroy, D., Worden, R., and Needham, S., 2003, Faeces, clay minerals and reservoir potential: Journal of the Geological Society, v. 160, p. 489493.Google Scholar
Meadows, P., and Reid, A., 1966, The behaviour of Corophium volutator (Crustacea: Amphipoda): Journal of Zoology, v. 150, p. 387399.Google Scholar
Melchor, R.N., Genise, J.F., Farina, J.L., Sánchez, M.V., Sarzetti, L., and Visconti, G., 2010, Large striated burrows from fluvial deposits of the Neogene Vichina Formation, La Rioja, Argentina: A crab origin suggested by neoichnology and sedimentology: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 291, p. 400418.Google Scholar
Mey, P., Nagtegaal, P., Roberti, K., and Hartevelt, J., 1968, Lithostratigraphic subdivision of post-hercynian deposits in the south-central Pyrenees, Spain: Leidse Geologische Mededelingen, v. 41, p. 221228.Google Scholar
Minter, N.J., Braddy, S.J., and Davis, R.B., 2007, Between a rock and a hard place: arthropod trackways and ichnotaxonomy: Lethaia, v. 40, p. 365375.Google Scholar
Morrissey, L.B., and Braddy, S.J., 2004, Terrestrial trace fossils from the Lower Old Red Sandstone, southwest Wales: Geological Journal, v. 39, no. 3–4, p. 315336.Google Scholar
Nagtegaal, P.J.C., Vanvliet, A., and Brouwer, J., 1983, Syntectonic coastal offlap and concurrent turbidite deposition-the Upper Cretaceous Aren Sandstone in the South-Central Pyrenees, Spain: Sedimentary Geology, v. 34, p. 185218.Google Scholar
Needham, S., Worden, R., and Cuadros, J., 2006, Sediment ingestion by worms and the production of bio-clays: a study of macrobiologically enhanced weathering and early diagenetic processes: Sedimentology, v. 53, p. 567579.Google Scholar
Netto, R.G., and Rossetti, D.F., 2003, Ichnology and salinity fluctuations: A case study from the early Miocene (Lower Barreiras Formation) of Sao Luís Basin, Maranhano, Brazil: Revista brasileira de Paleontología, v. 6, p. 518.Google Scholar
Nicholson, H.A., 1873, Contributions to the study of the errant annelides of the older Palaeozoic rocks: Royal Society London Proceedings, v. 21, p. 288290.Google Scholar
O’Geen, A.T., and Busacca, A.J., 2001, Faunal burrows as indicators of paleo-vegetation in eastern Washington, USA: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 169, p. 2337.Google Scholar
Pallas, P.S., 1766, Miscellanea zoologica: Hagae Comitum, Apud Petrum Van Cleef, 224 p.Google Scholar
Pemberton, S.G., and Frey, R.W., 1982, Trace fossil nomenclature and the Planolites-Palaeophycus dilemma: Journal of Paleontology, v. 56, p. 843881.Google Scholar
Pemberton, S.G., and Gingras, M.K., 2005, Classification and characterizations of biogenically enhanced permeability: AAPG bulletin, v. 89, p. 14931517.Google Scholar
Pemberton, S.G., and Wightman, D.M., 1992, Ichnological characteristics of brackish water deposits, in Pemberton, S.G., ed., Applications of ichnology to petroleum exploration: A core workshop: Society for Sedimentary Geology Core Workshop, v. 17, p 141167.Google Scholar
Pemberton, S.G., Spila, M., Pulham, A.J., Saunders, T., MacEachern, J.A., Robbins, D., and Sinclair, I.K., 2001, The application of Ichnofacies analysis to the evaluation of marginal marine reservoirs: Ichnology & Sedimentology of Shallow to Marginal Marine Systems. Geological Association of Canada Short Course, v. 15, p. 155179.Google Scholar
Percy, J.A., 1999, Keystone Corophium: Master of the Mudflats: Fundy Issues Series BoFEP, v. 13, http://www.bofep.org/corophium.htm (accessed June 2016).Google Scholar
Puigdefàbregas, C., Muñoz, J., and Vergés, J., 1992, Thrusting and foreland basin evolution in the southern Pyrenees, in McClay, K.R., ed., Thrust Tectonics, Berlin, Germany, Springer, p. 247254.Google Scholar
Rebata, L.A., Gingras, M.K., Räsänen, M.E., and Barberi, M., 2006a, Tidal-channel deposits on a delta plain from the Upper Miocene Nauta Formation, Marañón Foreland Sub-basin, Peru: Sedimentology, v. 53, p. 9711013.Google Scholar
Rebata, L.A., Räsänen, M.E., Gingras, M.K., Vieira, V. Jr, Barberi, M., and Irion, G., 2006b, Sedimentology and ichnology of tide-influenced Late Miocene successions in western Amazonia: The gradational transition between the Pebas and Nauta formations: Journal of South American Earth Sciences, v. 21, p. 96119.CrossRefGoogle Scholar
Riera, V., Oms, O., Gaete, R., and Galobart, A., 2009, The end-Cretaceous dinosaur succession in Europe: The Tremp Basin record (Spain): Palaeogeography, Palaeoclimatology, Palaeoecology, v. 283, p. 160171.Google Scholar
Rosell, J., 1965, Estudio geológico del sector del Prepirineo com-prendido entre los ríos Segre y Noguera Ribagorzana (Provincia de Lérida): Pirineos, v. 21, p. 1225.Google Scholar
Rosell, J., Linares, R., and Llompart, C., 2001, El “Garumniense” Prepirenaico: Revista de la Sociedad Geológica de España, v. 14, no. 1–2, p. 4756.Google Scholar
Rossi, C., 1997, Microcodium y trazas fósiles de invertebrados en facies continentales (Paleoceno de la Cuenca de Áger, Lérida): Revista de la Sociedad Geológica de España, v. 10, p. 371391.Google Scholar
Salter, J.W., 1857, On annelide-burrows and surface-markings from the Cambrian rocks of the Longmynd. Quarterly: Journal of the Geological Society of London, v. 13, p. 199206.Google Scholar
Sarkar, S., and Chaudhuri, A.K., 1992, Trace fossils in Middle to late triassic fluvial redbeds, Pranhita-Godavari Valley, South India: Ichnos, v. 2, p. 719.Google Scholar
Savrda, C.E., Blanton-Hooks, A.D., Collier, J.W., Drake, R.A., Graves, R.L., Hall, A.G., Nelson, A.I., Slone, J.C., Williams, D.D., and Wood, H.A., 2000, Taenidium and associated ichnofossils in fluvial deposits, Cretaceous Tuscaloosa Formation, eastern Alabama, southeastern USA: Ichnos, v. 7, p. 227242.Google Scholar
Schlirf, M., Uchman, A., and Kümmel, M., 2001, Upper Triassic (Keuper) non-marine trace fossils from the Hassberge area (Franconia, south-eastern Germany): Paläontologische Zeitschrift, v. 75, p. 7196.Google Scholar
Scotese 2001, Paleomap project. http://www.Scotese.com/K/t.htm (accessed November 2014).Google Scholar
Seilacher, A., 1964, Biogenic sedimentary structures, in Imbrie, J., and Newell, N., eds., Approaches to Paleoecology, New York, John Wiley and Sons, p. 296316.Google Scholar
Seilacher, A., 1967, Bathymetry of trace fossils: Marine Geology, v. 5, p. 413428.Google Scholar
Seilacher, A., and Seilacher, E., 1994, Bivalvian trace fossils: a lesson from actuopaleontology: Courier Forschungsinstitut Institut Senckenberg, v. 169, p. 515.Google Scholar
Shiers, M.N., Mountney, N.P., Hodgson, D.M., and Cobain, S.L., 2014, Depositional controls on tidally influenced fluvial successions, Neslen Formation, Utah, USA: Sedimentary Geology, v. 311, p. 116.Google Scholar
Sisulak, C.F., and Dashtgard, S.E., 2012, Seasonal controls on the development and character of Inclined Heterolithic Stratification in a tide-influenced, fluvially dominated channel: Fraser River, Canada: Journal of Sedimentary Research, v. 82, p. 244257.Google Scholar
Smith, J.J., Hasiotis, S.T., Kraus, M.J., and Woody, D.T., 2008, Naktodemasis bowni: New Ichnogenus and Ichnospecies for Adhesive Meniscate Burrows (AMB), and Paleoenvironmental Implications, Paleogene Willwood Formation, Bighorn Basin, Wyoming: Journal of Paleontology, v. 82, p. 267278.Google Scholar
Squires, R.L., and Advocate, D.M., 1984, Meniscate burrows from Miocene lacustrine-fluvial deposits, Diligencia Formation, Orocopia Mountains, Southern California: Journal of Paleontology, v. 58, p. 593597.Google Scholar
Stanley, K.O., and Fagerstrom, J.A., 1974, Miocene invertebrate trace fossils from a braided river environment, western Nebraska, USA: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 15, p. 6382.CrossRefGoogle Scholar
Taylor, A., and Goldring, R., 1993, Description and analysis of bioturbation and ichnofabric: Journal of the Geological Society, v. 150, no. 1, p. 141148.Google Scholar
Thomas, R.G., Smith, D.G., Wood, J.M., Visser, J., Calverley-Range, E.A., and Koster, E.H., 1987, Inclined heterolithic stratification-terminology, description, interpretation and significance: Sedimentary Geology, v. 53, p. 123179.Google Scholar
Torell, O, 1870, Petrificata Suecana Formationis Cambricae, Lunds, Univ. Arsskr. 6. Avdel. 2, v. 8, p. 114.Google Scholar
Trueman, E.R., 1966, Bivalve mollusks: fluid dynamics of burrowing: Science, v. 152, p. 523525.Google Scholar
Uchman, A., 1995, Taxonomy and palaeoecology of flysch trace fossils: the Marnoso-arenacea Formation and associated facies (Miocene, Northern Apennines, Italy): Beringeria, v. 15, p. 3115.Google Scholar
Vialov, O.S., 1962, Problematica of the Beacon Sandstone at Beacon Height West, Antarctica: New Zealand Journal of Geology and Geophysics, v. 5, p. 718732.Google Scholar
Vila, B., Galobart, À., Canudo, J., Le Loeuff, J., Dinarès-Turell, J., Riera, V., Oms, O., Tortosa, T., and Gaete, R., 2012, The diversity of sauropod dinosaurs and their first taxonomic succession from the latest Cretaceous of southwestern Europe: clues to demise and extinction: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 350–352, p. 1938.Google Scholar
Vila, B., Oms, O., Fondevilla, V., Gaete, R., Galobart, À., Riera, V., and Canudo, J.I., 2013, The latest succession of dinosaur tracksites in Europe: hadrosaur ichnology, track production and palaeoenvironments: PloS one, v. 8, p. e72579.Google Scholar
White, D., 1929, Flora of the Hermit Shale, Grand Canyon, Arizona: Publications Carnegie Institution 405, 221 p.Google Scholar
Zenker, J. C., 1836, Historisch-topographisches Taschenbuch von Jena und seiner Umgebung besonders in naturwissenschaftlicher und medizinischer Bezjehung, Wackenhoder, Jena, Germany338 p.Google Scholar
Zonneveld, J.P., Lavigne, J.M., Bartels, W.S., and Gunnell, G.F., 2006, Lunulichnus tuberosus ichnogen. and ichnosp. nov. from the Early Eocene Wasatch Formation, Fossil Butte National Monument, Wyoming: an arthropod-constructed trace fossil associated with alluvial firmgrounds: Ichnos, v. 13, p. 8794.Google Scholar