Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-25T14:10:11.171Z Has data issue: false hasContentIssue false

Marine Ecological State-Shifts Following the Triassic–Jurassic Mass Extinction

Published online by Cambridge University Press:  21 July 2017

Kathleen A. Ritterbush
Affiliation:
University of Chicago, 5734 S Ellis Ave, Chicago, IL 60637 USA
Yadira Ibarra
Affiliation:
Stanford University, Department of Environmental and Earth System Science, Stanford, CA 94305 USA
David J. Bottjer
Affiliation:
University of Southern California, 1657 Trousdale Pkwy., Los Angeles, CA, 90089, USA
Frank A. Corsetti
Affiliation:
University of Southern California, 1657 Trousdale Pkwy., Los Angeles, CA, 90089, USA
Silvia Rosas
Affiliation:
Pontificia Universidad Católica del Perú, Section for Mining Engineering, San Miguel, Lima, Peru
A. Joshua West
Affiliation:
University of Southern California, 1657 Trousdale Pkwy., Los Angeles, CA, 90089, USA
William M. Berelson
Affiliation:
University of Southern California, 1657 Trousdale Pkwy., Los Angeles, CA, 90089, USA
Joyce A. Yager
Affiliation:
University of Southern California, 1657 Trousdale Pkwy., Los Angeles, CA, 90089, USA
Get access

Abstract

One of the most severe extinction events in Earth history, the Triassic–Jurassic extinction, struck against a backdrop of radical increases in atmospheric CO2 and supercontinent breakup. This juxtaposition of first-order geophysical and biotic changes produced excellent case studies in Earth-Life Transitions. Recent recognition of a worldwide “carbonate gap” following the extinction has focused attention on causes, often invoked as eustacy or ocean acidification, but the ecology of the extinction aftermath remains poorly understood. Results from paleoecological studies on three separate Triassic–Jurassic records are presented and incorporated into regional depositional models. Examination of the Penarth Group of Great Britain reveals a widespread, laterally homogenous, level-bottom microbial stromatolite regime across the innermost ramp. The Sunrise Formation in Nevada, USA, was deposited during a biosiliceous (“glass”) regime dominated by demosponges across the inner ramp that lasted at least two million years. Investigations of the Pucará group in the central Andes of Peru revealed a demosponge-dominated level-bottom glass ramp with many similarities to the Nevada deposits, but offering broader regional extent and variation in recorded depositional settings. This suite of studies demonstrates state-shifts in marine ecological systems that also profoundly altered regional sedimentation regimes. The sponge-dominated systems produced glass ramp conditions instead of carbonate ramps, and indicate the importance of marine silica concentrations. The post-extinction changes in regional marine ecology demonstrate connectivity to changes in global climate and terrigenous weathering driven by global-scale geophysical processes.

Type
Research Article
Copyright
Copyright © 2015 by 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

Alroy, J. 2010. Fair sampling of taxonomic richness and unbiased estimation of origination and extinction rates. Paleontological Society Special Papers, 16:126.Google Scholar
Bachan, A., van de Schootbrugge, B., Fiebig, J., McRoberts, C. A., Ciarapica, G., and Payne, J. L. 2012. Carbon cycle dynamics following the end-Triassic mass extinction: Constraints from paired δ13Ccarb and δ13Corg records. Geochemistry Geophysics Geosystems, 13:doi: 10.1029/2012GC004150.Google Scholar
Bartolini, A., Guex, J., Spangenberg, J., Schoene, B., Taylor, D., Schaltegger, U., and Atudorei, V. 2012. Disentangling the Hettangian carbon isotope record: Implications for the aftermath of the end-Triassic mass extinction. Geochemistry Geophysics Geosystems, 13:Q01007.Google Scholar
Beerling, D. J., and Berner, R. A. 2002. Biogeochemical constraints on the Triassic–Jurassic boundary carbon cycle event. Global Biogeochemical Cycles, 16:10–110–13.CrossRefGoogle Scholar
Berner, R. A., and Beerling, D. J. 2007. Volcanic degassing necessary to produce a CaCO3 undersaturated ocean at the Triassic–Jurassic boundary. Palaeogeography, Palaeoclimatology, Palaeoecology, 244:368373.CrossRefGoogle Scholar
Blackburn, T. J., Olsen, P. E., Bowring, S. A., McLean, N. M., Kent, D. V., Puffer, J., McHone, G., Rasbury, E. T., and Et-Touhami, M. 2013. Zircon U-Pb geochronology links the end-Triassic extinction with the Central Atlantic Magmatic Province. Science, 340:941945.CrossRefGoogle ScholarPubMed
Blomeier, D., Dustira, A. M., Forke, H., and Scheibner, C. 2013. Facies analysis and depositional environments of a storm-dominated, temperate to cold, mixed siliceous–carbonate ramp:the Permian Kapp Starostin Formation in NE Svalbard. Norwegian Journal of Geology, 93:7594.Google Scholar
Bodzioch, A. 1993. Paleoecology of hexactinellid sponges from the epicontinental Triassic of Poland, p. 3544 In van Soest, R. W. M., van Kempen, T. M. G., and Braekman, J. C. (eds.), Sponges in Time and Space. Routledge, Rotterdam.Google Scholar
Bonis, N. R., and Kürschner, W. M. 2012. Vegetation history, diversity patterns, and climate change across the Triassic/Jurassic boundary. Paleobiology, 38:240264.Google Scholar
Bonis, N. R., Ruhl, M., and Kürschner, W. M. 2010. Milankovitch-scale palynological turnover across the Triassic–Jurassic transition at St. Audrie's Bay, SW UK. Journal of the Geological Society, 167:877888.Google Scholar
Carter, J. G., Barrera, E., and Tevesz, M. J. S. 1998. Thermal potentiation and mineralogical evolution in the Bivalvia (Mollusca). Journal of Paleontology, 72:9911010.Google Scholar
Corsetti, F. A., Ritterbush, K. A., Bottjer, D. J., Greene, S. E., Ibarra, Y., Yager, J. A., West, A. J., Berelson, W. M., Rosas, S., Becker, T. W., Levine, N. M., Loyd, S. J., Martindale, R. C., Petryshyn, V. A., Carroll, N. R., Petsios, E., Piazza, O., Pietsche, C., Stellmann, J. L., Thompson, J. R., Washington, K. A., and Wilmeth, D. T. 2015. Investigating the paleoecological consequences of supercontinent breakup: sponges clean up in the early Jurassic. The Sedimentary Record, 13:410.Google Scholar
Corso, J. D., Marzoli, A., Tateo, F., Jenkyns, H. C., Bertrand, H., Youbi, N., Mahmoudi, A., Font, E., Buratti, N., and Chilli, S. 2014. The dawn of CAMP volcanism and its bearing on the end-Triassic carbon cycle disruption. Journal of the Geological Society, 171:153164.Google Scholar
Delecat, S., Arp, G., and Reitner, J. 2011. Aftermath of the Triassic–Jurassic Boundary Crisis: Spiculite formation on drowned Triassic Steinplatte reef-slope by communities of hexactinellid sponges (northern Calcareous Alps, Austria). Advances in Stromatolite Geobiology, 355390.CrossRefGoogle Scholar
Erwin, D. 2008. Macroevolution of ecosystem engineering, niche construction and diversity. Trends in Ecology & Evolution, 23:304310.Google Scholar
Gates, L. M., James, N. P., and Beauchamp, B. 2004. A glass ramp: shallow-water Permian spiculitic chert sedimentation, Sverdrup Basin, Arctic Canada. Sedimentary Geology, 168:125147.Google Scholar
Gammon, P. R., James, N. P., and Pisera, A. 2000. Eocene spiculites and spongolites in southwestern Australia: not deep, not polar, but shallow and warm. Geology, 28:855858.Google Scholar
Gammon, P. R., and James, N. P. 2003. Paleoenvironmental controls on upper Eocene biosiliceous neritic sediments, southern Australia. Journal of Sedimentary Research, 73:157172.Google Scholar
Götz, A. E., Ruckwied, K., Pálfy, J., and Haas, J. 2009. Palynological evidence of synchronous changes within the terrestrial and marine realm at the Triassic/Jurassic boundary (Csővár section, Hungary). Review of Palaeobotany and Palynology, 156:401409.Google Scholar
Greene, S. E., Martindale, R. C., Ritterbush, K. A., Bottjer, D. J., Corsetti, F. A., and Berelson, W. M. 2012. Recognising ocean acidification in deep time: An evaluation of the evidence for acidification across the Triassic–Jurassic boundary. Earth-Science Reviews, 113:7293.Google Scholar
Guex, J., Bartolini, A., Atudorei, V., and Taylor, D. 2004. High-resolution ammonite and carbon isotope stratigraphy across the Triassic–Jurassic boundary New York Canyon (Nevada). Earth and Planetary Science Letters, 225:2941.Google Scholar
Hallam, A. 1988. A reevaluation of Jurassic eustasy in the light of new data and the revised Exxon curve, p. 261273 In Wilgus, C. K., Hastings, B. S., Kendall, C. G. St. C., Posamatir, H. W., Ron, C. A., and van Wagner, J. C. (eds.), Sea-Level Changes—An Integrated Approach. SEPM Special Publication 42. SEPM, Tulsa, OK.Google Scholar
Hallam, A., and Wignall, P. B. 1999. Mass extinctions and sea-level changes. Earth-Science Reviews, 48:217250.Google Scholar
Hallam, A., and Wignall, P. B. 2000. Facies changes across the Triassic–Jurassic boundary in Nevada, USA. Journal of the Geological Society, 157:4954.CrossRefGoogle Scholar
Hallam, A., Wignall, P. B., Yin, J., and Riding, J. B. 2000. An investigation into possible facies changes across the Triassic–Jurassic boundary in southern Tibet. Sedimentary Geology, 137:101106.Google Scholar
Hallam, T., Wignall, P., Hesselbo, S. P., Robinson, S. A., and Surlyk, F. 2004. Discussion on sea-level changes and facies development across potential Triassic–Jurassic boundary horizons, SW Britain. Journal of the Geological Society, 161:10531056.Google Scholar
Hamilton, D. 1961. Algal growths in the Rhaetic Cotham Marble of southern England. Palaeontology, 4:324333.Google Scholar
Hautmann, M. 2004. Effect of end-Triassic CO 2 maximum on carbonate sedimentation and marine mass extinction. Facies, 50:257261.Google Scholar
Hautmann, M., Benton, M. J., and Tomašových, A. 2008. Catastrophic ocean acidification at the Triassic–Jurassic boundary. Neues Jahrbuch Für Geologie Und Paläontologie-Abhandlungen, 249:119127.CrossRefGoogle Scholar
Hesselbo, S. P., Robinson, S. A., and Surlyk, F. 2004. Sea-level change and facies development across potential Triassic–Jurassic boundary horizons, SW Britain. Journal of the Geological Society, London, 161:365379.Google Scholar
Hillebrandt, von, A., Krystyn, L., and Kuerschner, W. 2007. A candidate GSSP for the base of the Jurassic in the Northern Calcareous Alps (Kuhjoch section, Karwendel Mountains, Tyrol, Austria). International Subcommission on Jurassic Stratigraphy Newsletter, 34:220.Google Scholar
Ibarra, Y., Corsetti, F. A., Greene, S. E., and Bottjer, D. J. 2014. Microfacies of the Cotham Marble: A tubestone carbonate microbialite from the Upper Triassic, southwestern U.K. Palaios, 29:885899.Google Scholar
Irmis, R. B., Nesbitt, S. J., Padian, K., Smith, N. D., Turner, A. H., Woody, D., and Downs, A. 2007. A Late Triassic Dinosauromorph assemblage from New Mexico and the rise of dinosaurs. Science, 317:358361.Google Scholar
Jackson, J. B. 2008. Ecological extinction and evolution in the brave new ocean. Proceedings of the National Academy of Sciences, 105:1145811465.Google Scholar
Ketner, K. B. 2009. Mid-Permian Phosphoria Sea in Nevada and the upwelling model. US Geological Survey Professional Paper, 1764:121.Google Scholar
Kiessling, W., and Simpson, C. 2011. On the potential for ocean acidification to be a general cause of ancient reef crises. Global Change Biology, 17:5667.Google Scholar
Kiessling, W., and Aberhan, M. 2007. Environmental determinants of marine benthic biodiversity dynamics through Triassic–Jurassic time. Paleobiology, 33:414434.CrossRefGoogle Scholar
Kiessling, W., Roniewicz, E., Villier, L., Léonide, P., and Struck, U. 2009. An early Hettangian coral reef in southern France: Implications for the end-Triassic reef crisis. Palaios, 24:657671.CrossRefGoogle Scholar
Kiessling, W., Aberhan, M., Brenneis, B., and Wagner, P. J. 2007. Extinction trajectories of benthic organisms across the Triassic–Jurassic boundary. Palaeogeography, Palaeoclimatology, Palaeoecology, 244:201222.Google Scholar
Kidder, D. L., and Erwin, D. H. 2001. Secular distribution of biogenic silica through the Phanerozoic: comparison of silica–replaced fossils and bedded cherts at the Series level. The Journal of Geology, 109:509522.Google Scholar
Korte, C., Hesselbo, S. P., Jenkyns, H. C., Rickaby, R. E. M., and Spötl, C. 2009. Palaeoenvironmental significance of carbon- and oxygen-isotope stratigraphy of marine Triassic–Jurassic boundary sections in SW Britain. Journal of the Geological Society, London, 166:431445.Google Scholar
Kuerschner, W. M., Bonis, N. R., and Krystyn, L. 2007. Carbon-isotope stratigraphy and palynostratigraphy of the Triassic–Jurassic transition in the Tiefengraben section —Northern Calcareous Alps (Austria). Palaeogeography, Palaeoclimatology, Palaeoecology, 244:257280.Google Scholar
Laws, R. A. 1982. Late Triassic depositional environments and molluscan associations from west-central Nevada. Palaeogeography, Palaeoclimatology, Palaeoecology, 37:131148.Google Scholar
Loughman, D., and Hallam, A. 1982. A facies analysis of the Pucará Group (Norian to Toarcian carbonates, organic-rich shale and phosphate) of central and northern Peru. Sedimentary Geology, 32:161194.Google Scholar
Lucas, S. G., Taylor, D. G., Guex, J., Tanner, L. H., and Krainer, K. 2007. The Proposed Global Stratotype Section and Point for the base of the Jurassic system in the New York Canyon Area, Nevada, USA. Triassic of the American West, New Mexico Museum of Natural History and Science Bulletin, 40:139161.Google Scholar
Mander, L., Twitchett, R. J., and Benton, M. J. 2008. Palaeoecology of the Late Triassic extinction event in the SW UK. Journal of the Geological Society, London, 165:319332.Google Scholar
Marzoli, A., Renne, P. R., Piccirillo, E. M., Ernesto, M., Bellieni, G., and DeMin, A. 1999. Extensive 200-million-year-old continental flood basalts of the central Atlantic Magmatic province. Science, 284:616618.Google Scholar
Mayall, M. J., and Wright, V. P. 1981. Algal tuft structures in stromatolites from the Upper Triassic of South-West England. Palaeontology, 24:655660.Google Scholar
Mazzullo, S. J., Wilhite, B. W., and Woolsey, I. W. 2009. Petroleum reservoirs within a spiculite-dominated depositional sequence: Cowley Formation (Mississippian: Lower Carboniferous), south-central Kansas. AAPG Bulletin, 93:16491689.Google Scholar
McElwain, J., Beerling, D., and Woodward, F. 1999. Fossil plants and global warming at the Triassic–Jurassic boundary. Science, 285:13861390.Google Scholar
McElwain, J. C., Wagner, P. J., and Hesselbo, S. P. 2009. Fossil plant relative abundances indicate sudden loss of Late Triassic biodiversity in East Greenland. Science, 324:15541556.Google Scholar
Michalik, J., Biron, A., Lintenerova, O., Gotz, A., and Ruckwied, K. 2010. Climate change at the Triassic/Jurassic boundary in the northwestern Tethyan realm, inferred from sections in the Tatra Mountains (Slovakia). Acta Geologica Polonica, 60:535548.Google Scholar
McHone, J. G. 2003. Volatile emissions from central Atlantic Magmatic Province basalts: mass assumptions and environmental consequences. The Central Atlantic Magmatic Province: Perspectives from the Rifted Fragments of Pangea, American Geophysical Union Monograph, 136:241254.Google Scholar
Morton, S. 2008. Newsletter. Morton, N. and Hesselbo, S. (eds.), International Subcommission on Jurassic Stratigraphy, 35/2:51 pp.Google Scholar
Muller, S. W., and Ferguson, H. G. 1939. Mesozoic stratigraphy of the Hawthorne and Tonopah quadrangles, Nevada. GSA Bulletin, 50:15731624.Google Scholar
Murchey, B. L., and Jones, D. J. 2002. A mid-Permian chert event: widespread deposition of biogenic siliceous sediments in coastal, island arc and oceanic basins. Palaeogeography, 96:161174.Google Scholar
Nomade, S., Knight, K., Beutel, E., Renne, P., Vérati, C., Féraud, G., Marzoli, A., Youbi, N., and Bertrand, H. 2007. Chronology of the Central Atlantic Magmatic Province: Implications for the Central Atlantic rifting processes and the Triassic–Jurassic biotic crisis. Palaeogeography, Palaeoclimatology, Palaeoecology, 244:326344.Google Scholar
Norström, A. V., Nyström, M., Lokrantz, J., and Folke, C. 2009. Alternative states on coral reefs: beyond coral–macroalgal phase shifts. Marine Ecology Progress Series, 376:295306.Google Scholar
Olsen, P. E. 2002. Ascent of dinosaurs linked to an iridium anomaly at the Triassic–Jurassic boundary. Science, 296:13051307.CrossRefGoogle Scholar
Pisera, A. 2006. Palaeontology of sponges—a review. Canadian Journal of Zoology, 84:242261.Google Scholar
Prinz, P. 1985. Zur Stratigraphie und Ammonitenfauna der Pucará-Gruppe bei San Vicente (Depto. Junín, Peru). Newsletters on Stratigraphy, 14:129141.Google Scholar
Racki, G., and Cordey, F. 2000: Radiolarian palaeoecology and radiolarites: is the present the key to the past? Earth-Science Reviews, 52:83120.Google Scholar
Radley, J. D., Twitchett, R. J., Mander, L., and Cope, J. C. W. 2008. Discussion on palaeoecology of the Late Triassic extinction event in the SW UK. Journal of the Geological Society, London, 165:988992.Google Scholar
Raup, D. M., and Sepkoski, J. J. Jr. 1982. Mass extinctions in the marine fossil record. Science, 215:15011503.Google Scholar
Ritterbush, K. A., Bottjer, D. J., Corsetti, F. A., and Rosas, S. 2014. New evidence on the role of siliceous sponges in ecology and sedimentary facies development in eastern Panthalassa following the Triassic–Jurassic mass extinction. Palaios, 29:652668.Google Scholar
Ritterbush, K. A., Rosas, S., Corsetti, F. A., Bottjer, D. J., and West, A. J. 2015. Andean sponges reveal long-term benthic ecosystem shifts following the end-Triassic mass extinction. Palaeogeography, Palaeoclimatology, Palaeoecology, 420:193209.Google Scholar
Rosas, S. 1994. Facies, diagenetic evolution, and sequence analysis along a SW-NE profile in the southern Pucará Basin (Upper Triassic–Lower Jurassic), Central Peru. Heidelberger Geowissenschaftliche Abhandlungen 80:1295.Google Scholar
Rosas, S., Fontbote, L., and Tankard, A. 2007. Tectonic evolution and paleogeography of the Mesozoic Pucará Basin, central Peru. Journal of South American Earth Sciences, 24:124.Google Scholar
Ruhl, M., Bonis, N. R., Reichart, G. J., and Damsté, J. 2011. Atmospheric carbon injection linked to end-Triassic mass extinction. Science.Google Scholar
Schaller, M. F., Wright, J. D., and Kent, D. V. 2011. Atmospheric pCO2 perturbations associated with the Central Atlantic Magmatic Province. Science, 331:14041409.CrossRefGoogle ScholarPubMed
Schaller, M. F., Wright, J. D., Kent, D. V., and Olsen, P. E. 2012. Rapid emplacement of the Central Atlantic Magmatic Province as a net sink for CO2 . Earth and Planetary Science Letters, 323–324:2739.Google Scholar
Schaltegger, U., Guex, J., Bartolini, A., Schoene, B., and Ovtcharova, M. 2008. Precise U–Pb age constraints for end-Triassic mass extinction, its correlation to volcanism and Hettangian post-extinction recovery. Earth and Planetary Science Letters, 267:266275.Google Scholar
Schoene, B., Guex, J., Bartolini, A., Schaltegger, U., and Blackburn, T. J. 2010. Correlating the end-Triassic mass extinction and flood basalt volcanism at the 100,000-year level. Geology, 38:387390.Google Scholar
Senowbari-Daryan, B. 1994. Mesozoic sponges of the Pucará Group, Peru. Palaeontographica Abt. A, 233:5774.Google Scholar
Senowbari-Daryan, B., and Stanley, G. D. 1994. Mesozoic sponge assemblage in Peru. Zentralblatt für Geologie und Paläontologie Teil 1:403412.Google Scholar
Swift, A., and Martill, D. M. 1999. Fossils of the Rhaetian Penarth Group. The Paleontological Association, Blackwell, London.Google Scholar
Szekely, T. S., and Grose, L. T. 1972. Stratigraphy of the carbonate, black shale, and phosphate of the Pucara Group (Upper Triassic–Lower Jurassic), Central Andes, Peru. Bulletin of the Geological Society of America, 83:407428.Google Scholar
Tackett, L. S., Kaufman, A. J., Corsetti, F. A., and Bottjer, D. J. 2014. Strontium isotope stratigraphy of the Gabbs Formation (Nevada): implications for global Norian–Rhaetian correlations and faunal turnover. Lethaia, 47:500511.Google Scholar
Taylor, D. G., Smith, P. L., Laws, R. A., and Guex, J. 1983. The stratigraphy and biofacies trends of the lower Mesozoic Gabbs and Sunrise formations, west-central Nevada. Canadian Journal of Earth Sciences, 20:15981608.Google Scholar
Taylor, D., and Guex, J. 2002. The Triassic/Jurassic system boundary in the John Day inlier, east-central Oregon. Oregon Geology, 64:328.Google Scholar
Twitchett, R. J., and Barras, C. G. 2004. Trace fossils in the aftermath of mass extinction event, p. 397418 In McIlroy, D. (ed.), The Application of Ichnology to Paleoenvironmental and Stratigraphic Analysis, Geological Society London Special Publication 228.Google Scholar
Uriz, M. J., Turon, X., Becerro, M. A., and Agell, G. 2003. Siliceous spicules and skeleton frameworks in sponges: origin, diversity, ultrastructural patterns, and and biological functions. Microscopy Research and Technique Special Issue: Biology of Silica Deposition in Sponges, 62:279299.Google Scholar
van de Schootbrugge, B., Tremolada, F., Rosenthal, Y., Bailey, T. R., Feist-Burkhardt, S., Brinkhuis, H., Pross, J., Kent, D. V., and Falkowski, P. G. 2007. End-Triassic calcification crisis and blooms of organic-walled ‘disaster species’. Palaeogeography, Palaeoclimatology, Palaeoecology, 244:126141.Google Scholar
Walker, L. J., Wilkinson, B. H., and Ivany, L. C. 2002: Continental drift and Phanerozoic carbonate accumulation in shallow–shelf and deep–marine settings. The Journal of Geology, 110:7587.Google Scholar
Ward, P. D., Haggart, J. W., Carter, E. S., Wilbur, D., Tipper, H. W., and Evans, T. 2001. Sudden productivity collapse associated with the Triassic–Jurassic boundary mass extinction. Science, 292:11481151.Google Scholar
Ward, P. D., Garrison, G. H., Haggart, J. W., Kring, D. A., and Beattie, M. J. 2004. Isotopic evidence bearing on late Triassic extinction events, Queen Charlotte Islands, British Columbia, and implications for the duration and cause of the Triassic/Jurassic mass extinction. Earth and Planetary Science Letters, 224:589600.Google Scholar
Ward, P. D., Garrison, G. H., Williford, K. H., Kring, D. A., Goodwin, D., Beattie, M. J., and McRoberts, C. A. 2007. The organic carbon isotopic and paleontological record across the Triassic–Jurassic boundary at the candidate GSSP section at Ferguson Hill, Muller Canyon, Nevada, USA. Palaeogeography, Palaeoclimatology, Palaeoecology, 244:281289.Google Scholar
Warrington, G., Cope, J. C. W., and Ivimey-Cook, H. C. 2008. The St Audrie's Bay-Doniford Bay section, Somerset, England: updated proposal for a candidate Global Stratotype Section and Point for the base of the Hettangian Stage, and of the Jurassic Section. International Subcomission on Jurassic Stratigraphy Newsletter, 35(1):266.Google Scholar
Whiteside, J. H., Olsen, P. E., Eglinton, T., Brookfield, M. E., and Sambrotto, R. N. 2010. From the cover: Compound-specific carbon isotopes from Earth's largest flood basalt eruptions directly linked to the end-Triassic mass extinction. Proceedings of the National Academy of Sciences, 107:67216725.Google Scholar
Wignall, P. 2001. Large igneous provinces and mass extinctions. Earth-Science Reviews, 53:133.Google Scholar
Wignall, P. B., and Bond, D. P. G. 2008. The end-Triassic and Early Jurassic mass extinction records in the British Isles. Proceedings of the Geologists' Association, 119:7384.Google Scholar
Wilmsen, M., and Neuweiler, F. 2008. Biosedimentology of the Early Jurassic post-extinction carbonate depositional system, central High Atlas rift basin, Morocco. Sedimentology, 55:773807.Google Scholar
Wotzlaw, J. F., Guex, J., Bartolini, A., Gallet, Y., Krystyn, L., McRoberts, C. A., Taylor, D., Schoene, B., and Schaltegger, U. 2014. Towards accurate numerical calibration of the Late Triassic: High-precision U-Pb geochronology constraints on the duration of the Rhaetian. Geology. Google Scholar
Wyld, S. J., Rogers, J. W., and Wright, J. E. 2001. Structural evolution within the Luning–Fencemaker fold-thrust belt, Nevada: progression from back-arc basin closure to intra-arc shortening. Journal of Structural Geology, 23:1971–199.Google Scholar