Hostname: page-component-8448b6f56d-dnltx Total loading time: 0 Render date: 2024-04-18T20:45:55.432Z Has data issue: false hasContentIssue false

Smithian shoreline migrations and depositional settings in Timpoweap Canyon (Early Triassic, Utah, USA)

Published online by Cambridge University Press:  17 January 2014

NICOLAS OLIVIER*
Affiliation:
Laboratoire de Géologie de Lyon, Terre, Planètes, Environnement, UMR CNRS 5276, Université Lyon 1, 69622 Villeurbanne cedex, France
ARNAUD BRAYARD
Affiliation:
UMR CNRS 6282 Biogéosciences, Université de Bourgogne, 6 boulevard Gabriel, 21000 Dijon, France
EMMANUEL FARA
Affiliation:
UMR CNRS 6282 Biogéosciences, Université de Bourgogne, 6 boulevard Gabriel, 21000 Dijon, France
KEVIN G. BYLUND
Affiliation:
140 South 700 East, Spanish Fork, Utah 84660, USA
JAMES F. JENKS
Affiliation:
1134 Johnson Ridge Lane, West Jordan, Utah 84084, USA
EMMANUELLE VENNIN
Affiliation:
UMR CNRS 6282 Biogéosciences, Université de Bourgogne, 6 boulevard Gabriel, 21000 Dijon, France
DANIEL A. STEPHEN
Affiliation:
140 South 700 East, Spanish Fork, Utah 84660, USA
GILLES ESCARGUEL
Affiliation:
Laboratoire de Géologie de Lyon, Terre, Planètes, Environnement, UMR CNRS 5276, Université Lyon 1, 69622 Villeurbanne cedex, France
*
Author for correspondence: Nicolas.Olivier@univ-lyon1.fr

Abstract

In Timpoweap Canyon near Hurricane (Utah, USA), spectacular outcrop conditions of Early Triassic rocks document the geometric relationships between a massive Smithian fenestral-microbial unit and underlying, lateral and overlying sedimentary units. This allows us to reconstruct the evolution of depositional environments and high-frequency relative sea-level fluctuations in the studied area. Depositional environments evolved from a coastal plain with continental deposits to peritidal settings with fenestral-microbial limestones, which are overlain by intertidal to shallow subtidal marine bioclastic limestones. This transgressive trend of a large-scale depositional sequence marks a long-term sea-level rise that is identified worldwide after the Permian–Triassic boundary. The fenestral-microbial sediments were deposited at the transition between continental settings (with terrigenous deposits) and shallow subtidal marine environments (with bioturbated and bioclastic limestones). Such a lateral zonation questions the interpretation of microbial deposits as anachronistic and disaster facies in the western USA basin. The depositional setting may have triggered the distribution of microbial deposits and contemporaneous marine biota. The fenestral-microbial unit is truncated by an erosional surface reflecting a drop in relative sea level at the scale of a medium depositional sequence. The local inherited topography allowed the recording of small-scale sequences characterized by clinoforms and short-distance lateral facies changes. Stratal stacking pattern and surface geometries allow the reconstruction of relative sea-level fluctuations and tracking of shoreline migrations. The stacking pattern of these small-scale sequences and the amplitude of corresponding high-frequency sea-level fluctuations are consistent with climatic control. Large- and medium-scale sequences suggest a regional tectonic control.

Type
Original Articles
Copyright
Copyright © Cambridge University Press 2014 

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

Baud, A., Richoz, S. & Pruss, S. B. 2007. The Lower Triassic anachronistic carbonate facies in space and time. Global and Planetary Change 55, 81–9.Google Scholar
Beatty, T. W., Zonneveld, J.-P. & Henderson, C. M. 2008. Anomalously diverse Early Triassic ichnofossil assemblages in northwest Pangea: a case for a shallow-marine habitable zone. Geology 36, 771–4.Google Scholar
Berra, F., Balini, M., Levera, M., Nicora, A. & Salamati, R. 2012. Anatomy of carbonate mounds from the Middle Anisian of Nakhlak (Central Iran): architecture and age of a subtidal microbial-bioclastic carbonate factory. Facies 58, 685705.Google Scholar
Blakey, R. C. 1974. Stratigraphic and depositional analysis of the Moenkopi Formation, southeastern Utah. Utah Geological and Mineral Survey, Bulletin 104, 81 pp.Google Scholar
Blakey, R. C. 1977. Petroliferous lithosomes in the Moenkopi Formation, southern Utah. Utah Geology 4, 6784.Google Scholar
Blakey, R. C. 1979. Oil impregnated carbonate rocks of the Timpoweap Member, Moenkopi Formation, Hurricane Cliffs area, Utah and Arizona. Utah Geology 6, 4554.Google Scholar
Brayard, A., Bylund, K. G., Jenks, J., Stephen, D. A., Olivier, N., Escarguel, G., Fara, E. & Vennin, E. 2013. Smithian ammonoid faunas from Utah: implications for Early Triassic biostratigraphy, correlations and basinal paleogeography. Swiss Journal of Paleontology 132, 141–219.Google Scholar
Brayard, A., Nützel, A., Kaim, A., Escarguel, G., Hautmann, M., Stephen, D. A., Bylund, K. G., Jenks, J. & Bucher, H. 2011 a. Gastropod evidence against the Early Triassic Lilliput effect: Reply. Geology 39, e233 pp.Google Scholar
Brayard, A., Nützel, A., Stephen, D. A., Bylund, K. G., Jenks, J. & Bucher, H. 2010. Gastropod evidence against the Early Triassic Lilliput effect. Geology 38, 147–50.Google Scholar
Brayard, A., Vennin, E., Olivier, N., Bylund, K. G., Jenks, J., Stephen, D. A., Bucher, H., Hofmann, R., Goudemand, N. & Escarguel, G. 2011 b. Transient metazoan reefs in the aftermath of the end-Permian mass extinction. Nature Geoscience 4, 693–7.Google Scholar
Brühwiler, T., Bucher, H., Brayard, A. & Goudemand, N. 2010. High-resolution biochronology and diversity dynamics of the Early Triassic ammonoid recovery: The Smithian faunas of the Northern Indian Margin. Palaeogeography, Palaeoclimatology, Palaeoecology 297, 491501.Google Scholar
Chen, Z. Q., Fraiser, M. L. & Bolton, C. 2012. Early Triassic trace fossils from Gondwana Interior Sea: implication for ecosystem recovery following the end-Permian mass extinction in south high-latitude region. Gondwana Research 22, 238–55.Google Scholar
Collinson, J. W., Kendall, C. G. S. C. & Marcantel, J. B. 1976. Permian-Triassic boundary in eastern Nevada and west-central Utah. Bulletin of the Geological Society of America 87, 821–4.Google Scholar
Dickinson, W. R. 2006. Geotectonic evolution of the Great Basin. Geosphere 2, 353–68.Google Scholar
Embry, A. F. 1997. Global sequence boundaries of the Triassic and their identification in the western Canada sedimentary basin. Canadian Petroleum Geology Bulletin 45, 415–33.Google Scholar
Esteban, M. & Pray, L.C. 1983. Pisoids and pisolite facies (Permian), Guadalupe Mountains, New Mexico and West Texas. In Coated Grains (ed. Peryt, T.), pp. 503–37. Berlin: Springer-Verlag.Google Scholar
Flügel, E. 2004. Microfacies of carbonate rocks. Analysis, interpretation and application. Berlin, Heidelberg, New York: Springer, 976 pp.Google Scholar
Forel, M.-B., Crasquin, S., Kershaw, S. & Collin, P.-Y. 2013. In the aftermath of the end-Permian extinction: the microbialite refuge? Terra Nova 25, 137–43.Google Scholar
Frakes, L. A., Francis, J. E. & Syktus, J. I. 1992. Climate Modes of the Phanerozoic: The History of the Earth's Climate Over the Past 600 million Years. Cambridge: Cambridge University Press, 274 pp.Google Scholar
Goodspeed, T. H. & Lucas, S. G. 2007. Stratigraphy, sedimentology, and sequence stratigraphy of the Lower Triassic Sinbad Formation, San Rafael Swell, Utah. In Triassic of the American West (eds Lucas, S. G. & Spielmann, J. A.), pp. 91102. New Mexico Museum of Natural History and Science Bulletin no. 40.Google Scholar
Gregory, H. E. 1950. Geology and geography of the Zion Park Region, Utah and Arizona. Geological Survey Professional Paper 220, 1200.Google Scholar
Hallam, A. & Wignall, P. B. 1999. Mass extinctions and sea-level changes. Earth Science Reviews 48, 217–50.Google Scholar
Haq, B. U. & Al-Qahtani, A. M. 2005. Phanerozoic cycles of sea-level change on the Arabian Platform. Geoarabia 10, 127–60.Google Scholar
Haq, U. B., Hardenbol, J. & Vail, P. R. 1987. Chronology of fluctuating sea levels since the Triassic. Science 235, 1156–67.Google Scholar
Haq, B. U. & Shutter, S. R. 2008. A chronology of Paleozoic sea-level changes. Science 322, 64–8.Google Scholar
Heydari, E., Hassanzadeh, J., Wade, W. J. & Ghazi, A. M. 2003. Permian–Triassic boundary interval in the Abadeh section of Iran with implications for mass extinction: Part 1. Sedimentology. Palaeogeography, Palaeoclimatology, Palaeoecology 193, 405–23.Google Scholar
Hofmann, R., Goudemand, N., Wasmer, M., Bucher, H. & Hautmann, M. 2011. New trace fossil evidence for an early recovery signal in the aftermath of the end-Permian mass extinction. Palaeogeography, Palaeoclimatology, Palaeoecology 310, 216–26.Google Scholar
Hofmann, R., Hautmann, M., Wasmer, M. & Bucher, H. 2013. Palaeoecology of the Spathian Virgin Formation (Utah, USA) and its implications for the Early Triassic recovery. Acta Palaeontologica Polonica 58, 149–73.Google Scholar
Huang, C., Tong, J., Hinnov, L. & Chen, Z. 2011. Did the great dying of life take 700 k.y.? Evidence from global astronomical correlation of the Permian-Triassic boundary interval. Geology 39, 779–82.Google Scholar
Jenson, J. 1986. Stratigraphy and facies analysis of the Upper Kaibab and Lower Moenkopi formations in Southwest Washington County, Utah. Brigham Young University Geology Studies 33, 121.Google Scholar
Kelley, N. P., Motani, R., Jiang, D. Y., Rieppel, O. & Schmitz, L. 2013. Selective extinction of Triassic marine reptiles during long-term sea-level changes illuminated by seawater strontium isotopes. Palaeogeography, Palaeoclimatology, Palaeoecology, published online 3 August 2012. doi: 10.1016/j.palaeo.2012.07.026.Google Scholar
Kershaw, S., Crasquin, S., Li, Y., Collin, P.-Y., Forel, M.-B., Mu, X., Baud, A., Wang, Y., Xie, S., Maurer, F. & Guo, L. 2012. Microbialites and global environmental change across the Permian-Triassic boundary: a synthesis. Geobiology 10, 2547.Google Scholar
Kershaw, S., Li, Y., Crasquin-Soleau, S., Feng, Q., Mu, X., Collin, P.-Y., Reynolds, A. & Guo, L. 2007. Earliest Triassic microbialites in the South China block and other areas: controls on their growth and distribution. Facies 53, 409–25.CrossRefGoogle Scholar
Kidder, D. L. & Worsley, T. R. 2004. Causes and consequences of extreme Permo-Triassic warming to globally equable climate and relation to the Permo-Triassic extinction and recovery. Palaeogeography, Palaeoclimatology, Palaeoecology 203, 207–37.Google Scholar
Konstantinov, A.G. 2008. Triassic ammonoids of Northeast Asia: diversity and evolutionary stages. Stratigraphy and Geological Correlation 16, 490502.Google Scholar
Lehrmann, D. J., Payne, J. L., Pei, D., Enos, P., Druke, D., Steffen, K., Zhang, J., Wei, J., Orchard, M. J. & Ellwood, B. 2007. Record of the End-Permian extinction and Triassic biotic recovery in the Chongzuo-Pingguo platform, southern Nanpanjiang basin, Guangxi, south China. Palaeogeography, Palaeoclimatology, Palaeoecology 252, 200–17.Google Scholar
Lucas, S. G., Krainer, K. & Milner, A. R. 2007. The type section and age of the Timpoweap Member and stratigraphic nomenclature of the Triassic Moenkopi Group in Southwestern Utah. In Triassic of the American West (eds Lucas, S. G. & Spielmann, J. A.), pp. 109–17. New Mexico Museum of Natural History and Science Bulletin no. 40.Google Scholar
Marenco, P. J., Griffin, J. M., Fraiser, M. L. & Clapham, M. E. 2012. Paleoecology and geochemistry of Early Triassic (Spathian) microbial mounds and implications for anoxia following the end-Permian mass extinction. Geology 40, 715–8.Google Scholar
Mata, S. A. & Bottjer, D. J. 2011. Origin of Lower Triassic microbialites in mixed carbonate-siliciclastic successions: ichnology, applied stratigraphy, and the end-Permian mass extinction. Palaeogeography, Palaeoclimatology, Palaeoecology 300, 158178.Google Scholar
McKee, E. D. 1954. Stratigraphy and History of the Moenkopi Formation of Triassic Age. Geological Society of America, Memoir 61, 133 pp.Google Scholar
Nielson, R. L. 1991. Petrology, sedimentology and stratigraphic implications of the Rock Canyon conglomerate, southwestern Utah. Utah Geological Survey, Miscellaneous Publication 91, 1–65.Google Scholar
Paull, R. A. & Paull, R. K. 1993. Interpretation of Early Triassic nonmarine-marine relations, Utah, USA. New Mexico Museum of Natural History and Science Bulletin 3, 403–9.Google Scholar
Paull, R. K. & Paull, R. A. 1997. Transgressive conodont faunas of the early Triassic: an opportunity for correlation in the Tethys and the circum-Pacific. In Late Palaeozoic and Early Mesozoic Circum-Pacific Events and their Global Correlation (eds Dickins, J. M., Zunyi, Y., Hongfu, Y., Lucas, S. G. & Acharyya, S. K.), pp. 158–67. Cambridge University Press, World and Regional Geology 10.Google Scholar
Pérez-López, A. & Pérez-Valera, F. 2011. Tempestite facies model for the epicontinental Triassic carbonates of the Betic Cordillera (southern Spain). Sedimentology 59, 646–78.Google Scholar
Peryt, T. M. 1983. Vadoids. In Coated Grains (ed. Peryt, T. M.), pp. 437–49. Berlin, Heidelberg, New York: Springer.Google Scholar
Pruss, S. B. & Bottjer, D. J. 2004. Late Early Triassic microbial reefs of the western United States; a description and model for their deposition in the aftermath of the end-Permian mass extinction. Palaeogeography, Palaeoclimatology, Palaeoecology 211, 127–37.Google Scholar
Pruss, S. B., Bottjer, D. J., Corsetti, F. A. & Baud, A. 2006. A global marine sedimentary response to the end-Permian mass extinction: examples from southern Turkey and the western United States. Earth-Science Reviews 78, 193206.Google Scholar
Quiquerez, A. & Dromart, G. 2006. Environmental control on granular clinoforms of ancient carbonate shelves. Geological Magazine 143, 343–65.Google Scholar
Reeside, J. B. Jr. & Bassler, H. 1922. Stratigraphic sections in southwestern Utah and northwestern Arizona. US Geological Survey, Professional Paper 129-D, pp. 53–77.Google Scholar
Ridente, D., Petrungaro, R., Falese, F. & Chiocci, F. L. 2012. Middle–Upper Pleistocene record of 100-ka depositional cycles on the Southern Tuscany continental margin (Tyrrhenian Sea, Italy). Sequence architecture and margin growth pattern. Marine Geology 326–8, 113.Google Scholar
Romano, C., Goudemand, N., Vennemann, T. W., Ware, D., Schneebeli-Hermann, E., Hochuli, P. A., Brühwiler, T., Brinkmann, W. & Bucher, H. 2013. Climatic and biotic upheavals following the end-Permian mass extinction. Nature Geoscience 6, 5760.Google Scholar
Ruban, D. A. 2008. Evolutionary rates of the Triassic marine macrofauna and sea-level changes: Evidences from the Northwestern Caucasus, Northern Neotethys (Russia). Palaeoworld 17, 115–25.Google Scholar
Ruddiman, W. F. 2003. Orbital insolation, ice volume, and greenhouse gases. Quaternary Science Reviews 22, 1597–629.Google Scholar
Sano, H. & Nakashima, K. 1997. Lowermost Triassic (Griesbachian) microbial bindstone-cementstone facies, southwest Japan. Facies 36, 124.Google Scholar
Sano, H., Onoue, T., Orchard, M. J. & Martini, R. 2012. Early Triassic peritidal carbonate sedimentation on a Panthalassan seamount: the Jesmond succession, Cache Creek Terrane, British Columbia. Facies 58, 113–30.Google Scholar
Schubert, J. K. & Bottjer, D. J. 1992. Early Triassic stromatolites as post-mass extinction disaster forms. Geology 20, 883–6.Google Scholar
Shackleton, N. J. 1987. Oxygen isotopes, ice volume and sea level. Quaternary Science Reviews 6, 183–90.Google Scholar
Stewart, J. H., Poole, F. G. & Wilson, R. F. 1972. Stratigraphy and origin of the Triassic Moenkopi Formation and related strata in the Colorado Plateau region. Geological Survey Professional Paper 691, 1195.Google Scholar
Twitchett, R. J. 2006. The palaeoclimatology, palaeoecology and palaeoenvironmental analysis of mass extinction events. Palaeogeography, Palaeoclimatology, Palaeoecology 232, 190213.CrossRefGoogle Scholar
Weidlich, O. 2007. PTB mass extinction and earliest Triassic recovery overlooked? New evidence for a marine origin of Lower Triassic mixed carbonate–siliciclastic sediments (Rogenstein Member), Germany. Palaeogeography, Palaeoclimatology, Palaeoecology 252, 259–69.Google Scholar
Woods, A. D. 2009. Anatomy of an anachronistic carbonate platform: Lower Triassic carbonates of the southwestern United States. Australian Journal of Earth Sciences 56, 825–39.Google Scholar
Woods, A. D. 2013. Microbial ooids and cortoids from the Lower Triassic (Spathian) Virgin Limestone, Nevada, USA: evidence for an Early Triassic microbial bloom in shallow depositional environments. Global and Planetary Change 105, 91101.Google Scholar
Wu, H., Zhang, S., Feng, Q., Jiang, G., Li, H. & Yang, T., 2012. Milankovitch and sub-Milankovitch cycles of the early Triassic Daye Formation, South China and their geochronological and paleoclimatic implications. Gondwana Research 22, 748–59.Google Scholar
Yang, H., Chen, Z. Q., Wang, Y., Tong, J., Song, H. & Chen, J. 2011. Composition and structure of microbialite ecosystems following the End-Permian mass extinction in South China. Palaeogeography, Palaeoclimatology, Palaeoecology 308, 111–28.Google Scholar
Yang, W. & Lehrmann, D. J. 2003. Milankovitch climatic signals in Lower Triassic (Olenekian) peritidal carbonate successions, Nanpanjiang Basin, South China. Palaeogeography, Palaeoclimatology, Palaeoecology 201, 283306.Google Scholar
Zatoń, M., Taylor, P. D. & Vinn, O. 2013. Early Triassic (Spathian) post-extinction microconchids from Western Pangea. Journal of Paleontology 87, 159–65.Google Scholar
Zonneveld, J.-P., Gingras, M. K. & Beatty, T. W. 2010. Diverse ichnofossil assemblages following the P-T mass extinction, Lower Triassic, Alberta and British Columbia, Canada: evidence for shallow marine refugia on the northwestern coast of Pangaea. Palaios 25, 368–92.Google Scholar