Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-25T10:45:03.847Z Has data issue: false hasContentIssue false

Mass Extinctions and the Structure and Function of Ecosystems

Published online by Cambridge University Press:  21 July 2017

Pincelli M. Hull
Affiliation:
Department of Geology & Geophysics, Yale University, PO Box 208109, New Haven, CT 06520-8109 USA
Simon A. F. Darroch
Affiliation:
Department of Geology & Geophysics, Yale University, PO Box 208109, New Haven, CT 06520-8109 USA
Get access

Abstract

Mass extinctions shape the history of life and can be used to inform understanding of the current biodiversity crisis. In this paper, a general introduction is provided to the methods used to investigate the ecosystem effects of mass extinctions (Part I) and to explore major patterns and outstanding research questions in the field (Part II). The five largest mass extinctions of the Phanerozoic had profoundly different effects on the structure and function of ecosystems, although the causes of these differences are currently unclear. Outstanding questions and knowledge gaps are identified that need to be addressed if the fossil record is to be used as a means of informing the dynamics of future biodiversity loss and ecosystem change.

Type
Research Article
Copyright
Copyright © 2013 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

Aberhan, M., Weidemeyer, S., Kiessling, W., Scasso, R. A., and Medina, F. A. 2007. Faunal evidence for reduced productivity and uncoordinated recovery in Southern Hemisphere Cretaceous–Paleogene boundary sections. Geology, 35:227230.Google Scholar
Alegret, L., and Thomas, E. 2009. Food supply to the seafloor in the Pacific Ocean after the Cretaceous/Paleogene boundary event. Marine Micropaleontology, 73:105116.Google Scholar
Alegret, L., Thomas, E., and Lohmann, K. C. 2012. End-Cretaceous marine mass extinction not caused by productivity collapse. Proceedings of the National Academy of Sciences of the United States of America, 109:728732.Google Scholar
Algeo, T. J., and Scheckler, S. E. 1998. Terrestrial-marine teleconnections in the Devonian: Links between the evolution of land plants, weathering processes, and marine anoxic events. Philosophical Transactions of the Royal Society B-Biological Sciences, 353:113128.CrossRefGoogle Scholar
Algeo, T. J., and Twitchett, R. J. 2010. Anomalous Early Triassic sediment fluxes due to elevated weathering rates and their biological consequences. Geology, 38:10231026.Google Scholar
Allison, P. A., and Briggs, D. E. G. 1991. Taphonomy of non-mineralized tissues, p. 2569. In Allison, P. A. and Briggs, D. E. G. (eds.), Taphonomy: Releasing the Data Locked in the Fossil Record. Plenum Press, New York.Google Scholar
Alroy, J., Aberhan, M., Bottjer, D. J., Foote, M., Fursich, F. T., Harries, P. J., Hendy, A. J. W., Holland, S. M., Ivany, L. C., Kiessling, W., Kosnik, M. A., Marshall, C. R., McGowan, A. J., Miller, A. I., Olszewski, T. D., Patzkowsky, M. E., Peters, S. E., Villier, L., Wagner, P. J., Bonuso, N., Borkow, P. S., Brenneis, B., Clapham, M. E., Fall, L. M., Ferguson, C. A., Hanson, V. L., Krug, A. Z., Layou, K. M., Leckey, E. H., Nurnberg, S., Powers, C. M., Sessa, J. A., Simpson, C., Tomasovych, A., and Visaggi, C. C. 2008. Phanerozoic trends in the global diversity of marine invertebrates. Science, 321:97100.Google Scholar
Alvarez, L. W., Alvarez, W., Asaro, F., and Michel, H. V. 1980. Extraterrestrial cause for the Cretaceous–Tertiary extinction—experimental results and theoretical interpretation. Science, 208:10951108.Google Scholar
Angielczyk, K. D., Roopnarine, P. D., and Wang, S. C. 2005. Modeling the role of primary productivity disruption in end-Permian extinctions, Karoo Basin, South Africa, p. 1623. In Lucas, S. G. and Ziegler, K. E. (eds.), The Nonmarine Permian. New Mexico Museum of Natural History and Science Bulletin 30. Albuquerque, New Mexico.Google Scholar
Archibald, J. D., Clemens, W. A., Padian, K., Rowe, T., MacLeod, N., Barrett, P. M., Gale, A., Holroyd, P., Sues, H. D., Arens, N. C., Horner, J. R., Wilson, G. P., Goodwin, M. B., Brochu, C. A., Lofgren, D. L., Hurlbert, S. H., Hartman, J. H., Eberth, D. A., Wignall, P. B., Currie, P. J., Weil, A., Prasad, G. V. R., Dingus, L., Courtillot, V., Milner, A., Milner, A., Bajpai, S., Ward, D. J., and Sahni, A. 2010. Cretaceous extinctions: Multiple causes. Science, 328:973973.Google Scholar
Ausich, W. I., and Bottjer, D. J. 1982. Tiering in suspension-feeding communities on soft substrata throughout the Phanerozoic. Science, 216:173174.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:Q09008. doi:10.1029/2012GC004150 Google Scholar
Bambach, R. K. 2006. Phanerozoic biodiversity mass extinctions. Annual Review of Earth and Planetary Sciences, 34:127155.Google Scholar
Bambach, R. K., Bush, A. M., and Erwin, D. H. 2007. Autecology and the filling of ecospace: Key metazoan radiations. Palaeontology, 50:122.Google Scholar
Bambach, R. K., Knoll, A. H., and Sepkoski, J. J. 2002. Anatomical and ecological constraints on Phanerozoic animal diversity in the marine realm. Proceedings of the National Academy of Sciences of the United States of America, 99:68546859.Google Scholar
Bambach, R. K., Knoll, A. H., and Wang, S. C. 2004. Origination, extinction, and mass depletions of marine diversity. Paleobiology, 30:522542.Google Scholar
Bapst, D. W., Bullock, P. C., Melchin, M. J., Sheets, H. D., and Mitchell, C. E. 2012. Graptoloid diversity and disparity became decoupled during the Ordovician mass extinction. Proceedings of the National Academy of Sciences of the United States of America, 109:34283433.CrossRefGoogle ScholarPubMed
Barnosky, A. D., Matzke, N., Tomiya, S., Wogan, G. O. U., Swartz, B., Quental, T. B., Marshall, C., McGuire, J. L., Lindsey, E. L., Maguire, K. C., Mersey, B., and Ferrer, E. A. 2011. Has the Earth's sixth mass extinction already arrived? Nature, 471:5157.Google Scholar
Barras, C. G., and Twitchett, R. J. 2007. Response of the marine infauna to Triassic–Jurassic environmental change: Ichnological data from southern England. Palaeogeography, Palaeoclimatology, Palaeoecology, 244:223241.Google Scholar
Bartolini, A., Guex, J., Spangenberg, J. E., Schoene, B., Taylor, D. G., 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, doi:10.1029/2011GC003807 Google Scholar
Baud, A., Cirilli, S., and Marcoux, J. 1997. Biotic response to mass extinction: The lowermost Triassic microbialites. Facies, 36:238242.Google Scholar
Baud, A., Richoz, S., and Pruss, S. 2007. The lower Triassic anachronistic carbonate facies in space and time. Global and Planetary Change, 55:8189.Google Scholar
Beerling, D. 2002. Palaeoclimatology—CO2 and the end-Triassic mass extinction. Nature, 415:386387.Google Scholar
Bennington, J. B., Dimichele, W. A., Badgley, C., Bambach, R. K., Barrett, P. M., Behrensmeyer, A. K., Bobe, R., Burnham, R. J., Daeschler, E. B., Van Dam, J., Eronen, J. T., Erwin, D. H., Finnegan, S., Holland, S. M., Hunt, G., Jablonski, D., Jackson, S. T., Jacobs, B. E., Kidwell, S. M., Koch, P. L., Kowalewski, M. J., Labandeira, C. C., Looy, C. V., Lyons, S. K., Novack-Gottshall, P. M., Potts, R., Roopnarine, P. D., Stromberg, C. A. E., Sues, H. D., Wagner, P. J., Wilf, P., and Wing, S. L. 2009. Critical issues of scale in paleoecology. PALAIOS, 24:14.Google Scholar
Benton, M. J. 1989. Patterns of evolution and extinction in vertebrates, p. 218241. In Allen, K. C. A. B. In Allen, D. E. G. (eds.), Evolution and the Fossil Record. John Wiley & Sons Ltd Belhaven, London.Google Scholar
Benton, M. J. 1995. Diversification and extinction in the history of life. Science, 268:5258.Google Scholar
Benton, M. J. 2009. The red queen and the court jester: species diversity and the role of biotic and abiotic factors through time. Science, 323:728732.Google Scholar
Benton, M. J. 2010. The origins of modern biodiversity on land. Philosophical Transactions of the Royal Society B-Biological Sciences, 365:36673679.Google Scholar
Benton, M. J., Dunhill, A. M., Lloyd, G. T., and Marx, F. G. 2011. Assessing the quality of the fossil record: insights from vertebrates, p. 6394. In McGowan, A. J. and Smith, A. B. (eds.), Comparing the Geological and Fossil Records: Implications for Biodiversity Studies. Geological Society Special Publications 358, Geological Society of London, London.Google Scholar
Benton, M. J., Tverdokhlebov, V. P., and Surkov, M. V. 2004. Ecosystem remodelling among vertebrates at the Permian–Triassic boundary in Russia. Nature, 432:97100.Google Scholar
Berggren, W. A., and Norris, R. D. 1997. Biostratigraphy, phylogeny and systematics of Paleocene trochospiral planktic foraminifera. Micropaleontology, 43:1116.Google Scholar
Berner, R. A. 1997. The rise of plants and their effect on weathering and atmospheric CO2 . Science, 276:544546.Google Scholar
Berner, R. A. 2002. Examination of hypotheses for the Permo–Triassic boundary extinction by carbon cycle modeling. Proceedings of the National Academy of Sciences of the United States of America, 99:41724177.CrossRefGoogle ScholarPubMed
Berner, R. A. 2006. GEOCARBSULF: A combined model for Phanerozoic atmospheric O2 and CO2 . Geochimica Et Cosmochimica Acta, 70:56535664.Google Scholar
Beurlen, K. 1956. Der Faunenschnitt an der Perm–Trias Grenze. Zeitschrift der Deutschen Geologischen Gesellschaft, 108:8899.Google Scholar
Birch, H. S., Coxall, H. K., and Pearson, P. N. 2012. Evolutionary ecology of Early Paleocene planktonic foraminifera: Size, depth habitat and symbiosis. Paleobiology, 38:374390.Google Scholar
Black, B. A., Elkins-Tanton, L. T., Rowe, M. C., and Peate, I. U. 2012. Magnitude and consequences of volatile release from the Siberian Traps. Earth and Planetary Science Letters, 317:363373.Google 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.Google Scholar
Blois, J. L., McGuire, J. L., and Hadly, E. A. 2010. Small mammal diversity loss in response to late-Pleistocene climatic change. Nature, 465:771775.CrossRefGoogle ScholarPubMed
Bode, H. B., Zeggel, B., Silakowski, B., Wenzel, S. C., Reichenbach, H., and Muller, R. 2003. Steroid biosynthesis in prokaryotes: identification of myxobacterial steroids and cloning of the first bacterial 2,3(S)-oxidosqualene cyclase from the myxobacterium Stigmatella aurantiaca. Molecular Microbiology, 47:471481.Google Scholar
Bonis, N. R., and Kurschner, W. M. 2012. Vegetation history, diversity patterns, and climate change across the Triassic/Jurassic boundary. Paleobiology, 38:240264.CrossRefGoogle Scholar
Borths, M. R., and Ausich, W. I. 2011. Ordovician–Silurian Lilliput crinoids during the end-Ordovician biotic crisis. Swiss Journal of Paleontology, 130:718.Google Scholar
Bottjer, D. J., and Ausich, W. I. 1986. Phanerozoic development of tiering in soft substrata suspension-feeding communities. Paleobiology, 12:400420.Google Scholar
Boyce, C. K., and Lee, J. E. 2011. Could land plant evolution have fed the marine revolution? Paleontological Research, 15:100105.Google Scholar
Brasier, M. D., Antcliffe, J. B., and Callow, R. H. T. 2010. Evolutionary trends in remarkable fossil preservation across the Edicaran–Cambrian transition and the impact of metazoan mixing, p. 519567. In Allison, P. A. and Bottjer, D. J. (eds.), Taphonomy: Process and Bias Through Time: Topics in Geobiology, 32.Google Scholar
Brassell, S. C., Eglinton, G., and Maxwell, J. R. 1983. The geochemistry of terpenoids and steroids. Biochemical Society Transactions, 11:575586.Google Scholar
Brayard, A., Vennin, E., Olivier, N., Bylund, K. G., Jenks, J., Stephen, D. A., Bucher, H., Hofmann, R., Goudemand, N., and Escarguel, G. 2011. Transient metazoan reefs in the aftermath of the end-Permian mass extinction. Nature Geoscience, 4:693697.Google Scholar
Brenchley, P. J., Carden, G. A., Hints, L., Kaljo, D., Marshall, J. D., Martma, T., Meidla, T., and Nolvak, J. 2003. High-resolution stable isotope stratigraphy of Upper Ordovician sequences: Constraints on the timing of bioevents and environmental changes associated with mass extinction and glaciation. Geological Society of America Bulletin, 115:89104.Google Scholar
Brenchley, P. J., Marshall, J. D., Carden, G. A. F., Robertson, D. B. R., Long, D. G. F., Meidla, T., Hints, L., and Anderson, T. F. 1994. Bathymetric and isotopic evidence for a short-lived Late Ordovician glaciation in a greenhouse period. Geology, 22:295298.Google Scholar
Brenchley, P. J., Marshall, J. D., and Underwood, C. J. 2001. Do all mass extinctions represent an ecological crisis? Evidence from the Late Ordovician. Geological Journal, 36:329340.Google Scholar
Brocks, J. J., Logan, G. A., Buick, R., and Summons, R. E. 1999. Archean molecular fossils and the early rise of eukaryotes. Science, 285:10331036.Google Scholar
Brose, U., Berlow, E. L., and Martinez, N. D. 2005. Scaling up keystone effects from simple to complex ecological networks. Ecology Letters, 8:13171325.Google Scholar
Brown, J. H. 1995. Macroecology. The University of Chicago Press, Chicago.Google Scholar
Buatois, L. A., Solange, A., and Mangano, M. G. 2013. Onshore expansion of benthic communities after the Late Devonian mass extinction. Lethaia, 46:251261.Google Scholar
Buggisch, W., Joachimski, M. M., Lehnert, O., Bergstrom, L., Repetski, J. E., and Webers, G. F. 2010. Did intense volcanism trigger the first Late Ordovician icehouse? Geology, 38:327330.Google Scholar
Burkepile, D. E., and Hay, M. E. 2006. Herbivore vs. nutrient control of marine primary producers: Context-dependent effects. Ecology, 87:31283139.Google Scholar
Bush, A. M., Bambach, R. K., and Daley, G. M. 2007. Changes in theoretical ecospace utilization in marine fossil assemblages between the mid-Paleozoic and late Cenozoic. Paleobiology, 33:7697.Google Scholar
Bush, A. M., and Pruss, S. B. 2013. Theoretical ecospace for ecosystem paleobiology: energy, nutrients, biominerals, and macroevolution, p. 120. In Bush, A. M., Pruss, S. B., and Payne, J. L. (eds.), Ecosystem Paleobiology and Geobiology, The Paleontological Society Papers 19. Yale Press, New Haven.Google Scholar
Calner, M. 2005. A Late Silurian extinction event and anachronistic period. Geology, 33:305308.Google Scholar
Calvert, S. E., and Pedersen, T. F. 2007. Elemental proxies for palaeoclimatic and palaeoceanographic variability in marine sediments: interpretation and application, p. 567644. In Hillaire-Marcel, C. and de Vernal, A. (eds.), Proxies in Late Cenozoic Paleoceanography. Elsevier, Oxford.Google Scholar
Canfield, D. E. 1989. Sulfate reduction and oxic respiration in marine sediments – Implications for organic-carbon preservation in euxinic environments. Deep-Sea Research Part A-Oceanographic Research Papers, 36:121138.Google Scholar
Cao, C. Q., Love, G. D., Hays, L. E., Wang, W., Shen, S. Z., and Summons, R. E. 2009. Biogeochemical evidence for euxinic oceans and ecological disturbance presaging the end-Permian mass extinction event. Earth and Planetary Science Letters, 281:188201.Google Scholar
Carpenter, S. R., Defries, R., Dietz, T., Mooney, H. A., Polasky, S., Reid, W. V., and Scholes, R. J. 2006. Millennium ecosystem assessment: Research needs. Science, 314:257258.Google Scholar
Carpenter, S. R., and Turner, M. G. 2000. Hares and tortoises: Interactions of fast and slow variables in ecosystems. Ecosystems, 3:495497.Google Scholar
Chapin, F. S., Walker, B. H., Hobbs, R. J., Hooper, D. U., Lawton, J. H., Sala, O. E., and Tilman, D. 1997. Biotic control over the functioning of ecosystems. Science, 277:500504.Google Scholar
Chen, D. Z., and Tucker, M. E. 2003. The Frasnian–Famennian mass extinction: insights from high-resolution sequence stratigraphy and cyclostratigraphy in South China. Palaeogeography, Palaeoclimatology, Palaeoecology, 193:87111.Google Scholar
Chen, Z.-Q., and Benton, M. J. 2012. The timing and pattern of biotic recovery following the end-Permian mass extinction. Nature Geoscience, 5:375383.Google Scholar
Chenet, A. L., Courtillot, V., Fluteau, F., Gerard, M., Quidelleur, X., Khadri, S. F. R., Subbarao, K. V., and Thordarson, T. 2009. Determination of rapid Deccan eruptions across the Cretaceous–Tertiary boundary using paleomagnetic secular variation: 2. Constraints from analysis of eight new sections and synthesis for a 3500-m-thick composite section. Journal of Geophysical Research-Solid Earth, 114:B06103, doi: 10.1029/2008JB005644 Google Scholar
Claeys, P., Kiessling, W., and Alvarez, W. 2002. Distribution of Chicxulub ejecta at the Cretaceous–Tertiary boundary, p. 5569. In Koeberl, C. and MacLeod, K. G. (eds.), Catastrophic Events and Mass Extinctions: Impacts and Beyond. Geological Society of America Special Paper 356, Geological Society of America, Boulder, CO.Google Scholar
Clapham, M. E., and Payne, J. L. 2011. Acidification, anoxia, and extinction: A multiple logistic regression analysis of extinction selectivity during the Middle and Late Permian. Geology, 39:10591062.Google Scholar
Clapham, M. E., Shen, S. Z., and Bottjer, D. J. 2009. The double mass extinction revisited: reassessing the severity, selectivity, and causes of the end-Guadalupian biotic crisis (Late Permian). Paleobiology, 35:3250.Google Scholar
Copper, P. 1988. Ecological succession in Phanerozoic reef ecosystems: is it real? PALAIOS, 3:136151.Google Scholar
Copper, P. 1994. Ancient Reef Ecosystem Expansion and Collapse. Coral Reefs, 13:311.Google Scholar
Copper, P. 2002. Reef development at the Frasnian/Famennian mass extinction boundary. Palaeogeography, Palaeoclimatology, Palaeoecology, 181:2765.Google Scholar
Coxall, H. K., D'hondt, S., and Zachos, J. C. 2006. Pelagic evolution and environmental recovery after the Cretaceous–Paleogene mass extinction. Geology, 34:297300.Google Scholar
Cramer, B. S., Toggweiler, J. R., Wright, J. D., Katz, M. E., and Miller, K. G. 2009. Ocean overturning since the Late Cretaceous: Inferences from a new benthic foraminiferal isotope compilation. Paleoceanography, 24:PA4216, doi: 10.1029/2008PA001683 Google Scholar
D'hondt, S. 2005. Consequences of the Cretaceous/Paleogene mass extinction for marine ecosystems. Annual Review of Ecology Evolution and Systematics, 36:295317.CrossRefGoogle Scholar
D'hondt, S., Donaghay, P., Zachos, J. C., Luttenberg, D., and Lindinger, M. 1998. Organic carbon fluxes and ecological recovery from the Cretaceous–Tertiary mass extinction. Science, 282:276279.Google Scholar
Darroch, S. A. F. 2012. Carbonate facies control on the fidelity of surface-subsurface agreement in benthic foraminiferal assemblages: Implications for index-based paleoecology. PALAIOS, 27:137150.Google Scholar
Dawson, T. P., Jackson, S. T., House, J. I., Prentice, I. C., and Mace, G. M. 2011. Beyond predictions: Biodiversity conservation in a changing climate. Science, 332:5358.Google Scholar
Donohue, I., Petchey, O. L., Montoya, J. M., Jackson, A. L., McNally, L., Viana, M., Healy, K., Lurgi, M., O'Connor, N. E., and Emmerson, M. C. 2013. On the dimensionality of ecological stability. Ecology Letters, 16:421429.Google Scholar
Droser, M. L., and Bottjer, D. J. 1986. A semiquantitative field classification of ichnofabric. Journal of Sedimentary Petrology, 56:558559.Google Scholar
Droser, M. L., Bottjer, D. J., and Sheehan, P. M. 1997. Evaluating the ecological architecture of major events in the Phanerozoic history of marine invertebrate life. Geology, 25:167170.Google Scholar
Droser, M. L., Bottjer, D. J., Sheehan, P. M., and McGhee, G. R. 2000. Decoupling of taxonomic and ecologic severity of Phanerozoic marine mass extinctions. Geology, 28:675678.Google Scholar
Dunne, J. A., and Williams, R. J. 2009. Cascading extinctions and community collapse in model food webs. Philosophical Transactions of the Royal Society B-Biological Sciences, 364:17111723.Google Scholar
Dunne, J. A., Williams, R. J., and Martinez, N. D. 2004. Network structure and robustness of marine food webs. Marine Ecology-Progress Series, 273:291302.Google Scholar
Dunne, J. A., Williams, R. J., Martinez, N. D., Wood, R. A., and Erwin, D. H. 2008. Compilation and network analyses of Cambrian food webs . PLoS Biology, 6:693708.Google Scholar
Dutkiewicz, A., Rasmussen, B., and Buick, R. 1998. Oil preserved in fluid inclusions in Archaean sandstones. Nature, 395:885888.Google Scholar
Elick, J. M., Driese, S. G., and Mora, C. I. 1998. Very large plant and root traces from the Early to Middle Devonian: Implications for early terrestrial ecosystems and atmospheric pCO2 . Geology, 26:143146.Google Scholar
Ellwood, B. B., Benoist, S. L., El Hassani, A., Wheeler, C., and Crick, R. E. 2003. Impact ejecta layer from the mid-Devonian: Possible connection to global mass extinctions. Science, 300:17341737.Google Scholar
Elmqvist, T., Folke, C., Nystrom, M., Peterson, G., Bengtsson, J., Walker, B., and Norberg, J. 2003. Response diversity, ecosystem change, and resilience. Frontiers in Ecology and the Environment, 1:488494.Google Scholar
Erwin, D. H. 1990. The end-Permian mass extinction. Annual Review of Ecology and Systematics, 21:6991.Google Scholar
Erwin, D. H. 1994. The Permo–Triassic extinction. Nature, 367:231236.Google Scholar
Erwin, D. H. 1998. The end and the beginning: recoveries from mass extinctions. Trends in Ecology & Evolution, 13:344349.Google Scholar
Erwin, D. H. 2001. Lessons from the past: biotic recoveries from mass extinctions. Proceedings of the National Academy of Sciences of the United States of America, 98:53995403.Google Scholar
Erwin, D. H. 2006. Extinction: How Life on Earth Nearly Ended 250 Million Years Ago. Princeton University Press, Princeton, New Jersey.Google Scholar
Erwin, D. H. 2007. Disparity: Morphological pattern and developmental context. Palaeontology, 50:5773.Google Scholar
Erwin, D. H. 2008. Extinction as the loss of evolutionary history. Proceedings of the National Academy of Sciences of the United States of America, 105:1152011527.Google Scholar
Erwin, D. H. 2009. A call to the custodians of deep time. Nature, 462:282283.Google Scholar
Erwin, D. H., Bowring, S. A., and Yugan, J. 2002. End-Permian mass extinctions: A review, p. 363383. In Köberl, C. and MacLeod, K. G. (eds.), Catastrophic Events and Mass Extinctions: Impacts and Beyond. Geological Society of America Special Paper 356, Geological Society of America, Boulder, CO Google Scholar
Erwin, D. H., and Tweedt, S. M. 2012. Ecological drivers of the Ediacaran–Cambrian diversification of Metazoa. Evolutionary Ecology, 26:417433.Google Scholar
Estes, J. A., Tinker, M. T., Williams, T. M., and Doak, D. F. 1998. Killer whale predation on sea otters linking oceanic and nearshore ecosystems. Science, 282:473476.Google Scholar
Ezaro, T. H. G., Aze, T., Pearson, P. N., and Purvis, A. 2011. Interplay between changing climate and species' ecology drives macroevolutionary dynamics. Science, 332:349351.Google Scholar
Fastovsky, D. E., and Sheehan, P. M. 2005. The extinction of the dinosaurs in North America. GSA Today, 15:410.Google Scholar
Fastovsky, D. E., Sheehan, P. M., Isbell, J. L., and Grandpre, R. 2008. Anomalous, temporary terrestrial sedimentary environments following Cretaceous–Tertiary ecosystems disruptions: North America, Asia, and Europe [Abstract], Geological Society of America Abstracts with Program 40, p. 322.Google Scholar
Faul, K. L., and Delaney, M. L. 2000. Nutrient and paleoproductivity dynamics in the early Paleogene. GFF, 122:4647, doi: 10.1080/11035890001221046 Google Scholar
Field, D. B., Baumgartner, T. R., Charles, C. D., Ferreira-Bartrina, V., and Ohman, M. D. 2006. Planktonic foraminifera of the California Current reflect 20th Century warming. Science, 311:6366.Google Scholar
Finnegan, S., Bergmann, K., Eiler, J. M., Jones, D. S., Fike, D. A., Eisenman, I., Hughes, N. C., Tripati, A. K., and Fischer, W. W. 2011. The magnitude and duration of Late Ordovician–Early Silurian glaciation. Science, 331:903906.Google Scholar
Finnegan, S., and Droser, M. L. 2008. Body size, energetics, and the Ordovician restructuring of marine ecosystems. Paleobiology, 34:342359.Google Scholar
Finnegan, S., Heim, N. A., Peters, S. E., and Fischer, W. W. 2012. Climate change and the selective signature of the Late Ordovician mass extinction. Proceedings of the National Academy of Sciences of the United States of America, 109:68296834.Google Scholar
Fischlin, A., Midgley, G. F., Price, J. T., Leemans, R., Gopal, B., Turley, C., Rounsevell, M. D. A., Dube, O. P., Tarazona, J., and Velichko, A. A. 2007. Ecosystems, their properties, goods, and services, p. 211272. In Parry, M. L., Canziani, O. F., Palutikof, J. P., van der Linden, P. J., and Hanson, C. E. (eds.), Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge.Google Scholar
Flessa, K. W., Cutler, A. H., and Meldahl, K. H. 1993. Time and taphonomy: Quantitative estimates of time-averaging and stratigraphic disorder in a shallow marine habitat. Paleobiology, 19:266286.Google Scholar
Flügel, E., and Kiessling, W. 2002. Patterns of Phanerozoic reef crises, p. 691723. In Kiessling, W., Flügel, E., and Golonka, J. (eds.), Phanerozoic Reef Patterns. SEPM Special Publication 72, Tulsa, OK.Google Scholar
Foote, M. 1999. Morphological diversity in the evolutionary radiation of Paleozoic and post-Paleozoic crinoids. Paleobiology, 25:1115.Google Scholar
Friedman, M. 2009. Ecomorphological selectivity among marine teleost fishes during the end-Cretaceous extinction. Proceedings of the National Academy of Sciences of the United States of America, 106:52185223.Google Scholar
Friedman, M. 2010. Explosive morphological diversification of spiny-finned teleost fishes in the aftermath of the end-Cretaceous extinction. Proceedings of the Royal Society B Biological Sciences, 277:16751683.Google Scholar
Friedrich, O., Schiebel, R., Wilson, P. A., Weldeab, S., Beer, C. J., Cooper, M. J., and Fiebig, J. 2012. Influence of test size, water depth, and ecology on Mg/Ca, Sr/Ca, δO18 and δC13 in nine modern species of planktic foraminifers. Earth and Planetary Science Letters, 319:133145.Google Scholar
Ganino, C., and Arndt, N. T. 2009. Climate changes caused by degassing of sediments during the emplacement of large igneous provinces. Geology, 37:323326.Google Scholar
Gibbs, S. J., Poultin, A. J., Bown, P. R., Daniels, C. J., Hopkins, J., Young, J. R., Jones, H. L., Thiemann, G. J., O'dea, S. A., and Newsam, C. 2013. Species-specific growth response of coccolithophores to Palaeocene–Eocene environmental change. Nature Geoscience, 6:218222.Google Scholar
Godderis, Y., and Joachimski, M. M. 2004. Global change in the Late Devonian: Modelling the Frasnian–Famennian short-term carbon isotope excursions. Palaeogeography, Palaeoclimatology, Palaeoecology, 202:309329.Google Scholar
Gould, S. J., and Calloway, C. B. 1980. Clams and brachiopods: ships that pass in the night. Paleobiology, 6:383396.Google Scholar
Greene, S. E., Bottjer, D. J., Corsetti, F. A., Berelson, W. M., and Zonneveld, J.-P. 2012. A subseafloor carboante factory across the Triassic–Jurrasic transition. Geology, 40:10431046.Google Scholar
Greene, S. E., Bottjer, D. J., Hagdorn, H., and Zonneveld, J.-P. 2011. The Mesozoic return of Paleozoic faunal constituents: A decoupling of taxonomic and ecological dominance during the recovery from the end-Permian mass extinction. Palaeogeography, Palaeoclimatology, Palaeoecology, 308:224232.Google Scholar
Grice, K., Cao, C. Q., Love, G. D., Bottcher, M. E., Twitchett, R. J., Grosjean, E., Summons, R. E., Turgeon, S. C., Dunning, W., and Jin, Y. G. 2005. Photic zone euxinia during the Permian–Triassic superanoxic event. Science, 307:706709.Google Scholar
Grotzinger, J. P., and Knoll, A. H. 1995. Anomalous carbonate precipitates: Is the precambrian the key to the Permian? PALAIOS, 10:578596.Google Scholar
Hallam, A. 1965. Environmental causes of stunting in living and fossil marine benthonic invertebrates. Palaeontology, 8:132155.Google Scholar
Hallam, A., and Wignall, P. B. 1997. Mass Extinctions and Their Aftermath. Oxford University Press, Oxford.Google Scholar
Hallmann, C., Kelly, A. E., Neal Gupta, S., and Summons, R. E. 2011. Reconstructing deep–time biology with molecular fossils, p. 355401. In Laflamme, M., Schiffbauer, J. D., and Dornbos, S. Q. (eds.), Quantifying the Evolution of Early Life: Numerical Approaches to the Evaluation of Fossils and Ancient Ecosystems. Springer, Dordrecht.Google Scholar
Halpern, B. S., Walbridge, S., Selkoe, K. A., Kappel, C. V., Micheli, F., D'agrosa, C., Bruno, J. F., Casey, K. S., Ebert, C., Fox, H. E., Fujita, R., Heinemann, D., Lenihan, H. S., Madin, E. M. P., Perry, M. T., Selig, E. R., Spalding, M., Steneck, R., and Watson, R. 2008. A global map of human impact on marine ecosystems. Science, 319:948952.Google Scholar
Hammarlund, E. U., Dahl, T. W., Harper, D. A. T., Bond, D. P. G., Nielsen, A. T., Bjerrum, C. J., Schovsbo, N. H., Schonlaub, H. P., Zalasiewicz, J. A., and Canfield, D. E. 2012. A sulfidic driver for the end-Ordovician mass extinction. Earth and Planetary Science Letters, 331:128139.Google Scholar
Harmon, J. P., Moran, N. A., and Ives, A. R. 2009. Species response to environmental change: Impacts of food web interactions and evolution. Science, 323:13471350.Google Scholar
Harnik, P. G., Lotze, H. K., Anderson, S. C., Finkel, Z. V., Finnegan, S., Lindberg, D. R., Liow, L. H., Lockwood, R., McClain, C. R., McGuire, J. L., O'dea, A., Pandolfi, J. M., Simpson, C., and Tittensor, D. P. 2012. Extinctions in ancient and modern seas. Trends in Ecology & Evolution, 27:608617.Google Scholar
Harries, P. J., Kauffman, E. G., and Hansen, T. A. 1996. Models for biotic survival following mass extinction, p. 4160. In Hart, M. B. (ed.), Biotic Recovery from Mass Extinction Events. Geological Society of London Special Publication, 102, The Geological Society, London.Google Scholar
Harries, P. J., and Knorr, P. O. 2009. What does the ‘Lilliput Effect’ mean? Palaeogeography, Palaeoclimatology, Palaeoecology, 284:410.Google Scholar
Hassan, R., Scholes, R., and Ash, N. 2005. Ecosystems and human well-being: current state and trends. The Millennium Ecosystem Assessment Series, 1. Island Press, Washington.Google Scholar
Hays, J. D., Imbrie, J., and Shackleton, N. J. 1976. Variations in Earth's orbit: pacemaker of ice ages. Science, 194:11211132.Google Scholar
Hays, L., Beatty, T., Henderson, C. M., Love, G. D., and Summons, R. E. 2007. Evidence for photic zone euxinia through the end-Permian mass extinction in the Panthalassic Ocean (Peace River Basin, Western Canada). Palaeoworld, 16:3950.Google Scholar
He, Q., Bertness, M. D., and Altieri, A. H. 2013. Global shifts towards positive species interactions with increasing environmental stress. Ecology Letters, 16:695706.Google Scholar
Hedges, J. I., and Keil, R. G. 1995. Sedimentary organic-matter preservation: An assessment and speculative synthesis. Marine Chemistry, 49:81115.Google Scholar
Herringshaw, L. G., and Davies, N. S. 2008. Bioturbation levels during the end-Ordovician extinction event: a case study of shallow marine strata from the Welsh Basin. Aquatic Biology, 2:279287.Google Scholar
Herrmann, A. D., Patzkowsky, M. E., and Pollard, D. 2003. Obliquity forcing with 8–12 times preindustrial levels of atmospheric pCO2 during the Late Ordovician glaciation. Geology, 31:485488.Google Scholar
Hesselbo, S. P., McRoberts, C. A., and Palfy, J. 2007. Triassic–Jurassic boundary events: Problems, progress, possibilities. Palaeogeography, Palaeoclimatology, Palaeoecology, 244:110.Google Scholar
Hildebrand, A. R., Penfield, G. T., Kring, D. A., Pilkington, M., Camargo, A., Jacobsen, S. B., and Boynton, W. V. 1991. Chicxulub crater: a possible Cretaceous–Tertiary Boundary impact crater on the Yucatan Peninsula, Mexico. Geology, 19:867871.Google Scholar
Hillaire-Marcel, C., and De Vernal, A. 2007. Proxies in Late Cenozoic Paleoceanography. Elsevier, Amsterdam.Google Scholar
Hilting, A. K., Kump, L. R., and Bralower, T. J. 2008. Variations in the oceanic vertical carbon isotope gradient and their implications for the Paleocene–Eocene biological pump. Paleoceanography, 23:PA3222, doi: 10.1029/2007PA001458 Google Scholar
Hironaga, R., and Yamamura, N. 2010. Effects of extinction on food web structures on an evolutionary time scale. Journal of Theoretical Biology, 263:161168.Google Scholar
Hönisch, B., Ridgwell, A., Schmidt, D. N., Thomas, E., Gibbs, S., Sluijs, A., Zeebe, R. E., Kump, L., Martindale, R. C., Greene, S. E., Kiessling, W., Ries, J., Zachos, J. C., Royer, D. L., Barker, S., Marchitto, T. M. Jr., Moyer, R., Pelejero, C., Ziveri, P., Foster, G. L., and Williams, B. 2012. The geological record of ocean acidification. Science, 335:10581063.Google Scholar
Hooper, D. U., Chapin, F. S., Ewel, J. J., Hector, A., Inchausti, P., Lavorel, S., Lawton, J. H., Lodge, D. M., Loreau, M., Naeem, S., Schmid, B., Setala, H., Symstad, A. J., Vandermeer, J., and Wardle, D. A. 2005. Effects of biodiversity on ecosystem functioning: A consensus of current knowledge. Ecological Monographs, 75:335.Google Scholar
House, M. R. 2002. Strength, timing, setting and cause of mid-Palaeozoic extinctions. Palaeogeography, Palaeoclimatology, Palaeoecology, 181:525.Google Scholar
Hsu, K. J., and McKenzie, J. A. 1985. A “Strangelove Ocean” in the earliest Tertiary, p. 487492. in Sundquist, E. T. and Broecker, W. (eds.), The Carbon Cycle and Atmospheric CO2: Natural Variations Archaean to Present. Geophysical Monograph Series 32, American Geophysical Union, Washington, D.C. Google Scholar
Huang, B., Harper, D. A. T., Zhan, R. B., and Rong, J. Y. 2010. Can the Lilliput Effect be detected in the brachiopod faunas of South China following the terminal Ordovician mass extinction? Palaeogeography, Palaeoclimatology, Palaeoecology, 285:277286.Google Scholar
Hull, P. M., and Norris, R. D. 2011. Diverse patterns of ocean export productivity change across the Cretaceous–Paleogene boundary: New insights from biogenic barium. Paleoceanography, 26:PA3205, doi: 10.1029/2010PA002082 Google Scholar
Hull, P. M., Norris, R. D., Bralower, T. J., and Schueth, J. D. 2011. A role for chance in marine recovery from the end-Cretaceous extinction. Nature Geoscience, 4:856860.Google Scholar
Irmis, R. B., and Whiteside, J. H. 2012. Delayed recovery of non-marine tetrapods after the end-Permian mass extinction tracks global carbon cycle. Proceedings of the Royal Society B-Biological Sciences, 279:13101318.Google Scholar
IUCN. International Union for Conservation of Nature Red List of Threatened Species. www.iucnredlist.org [accessed 9-30-13] Google Scholar
Ivany, L. C., and Huber, B. (eds). 2012. Reconstructing Earth's Deep-Time Climate. The Paleontological Society Papers 18. Yale Press, New Haven.Google Scholar
Ives, A. R., and Carpenter, S. R. 2007. Stability and diversity of ecosystems. Science, 317:5862.Google Scholar
Jablonski, D. 1991. Extinctions: a paleontological perspective. Science, 253:754757.Google Scholar
Jablonski, D. 1998. Geographic variation in the molluscan recovery from the end-Cretaceous extinction. Science, 279:13271330.Google Scholar
Jablonski, D. 2008. Biotic interactions and macroevolution: Extensions and mismatches across scales and levels. Evolution, 62:715739.Google Scholar
Jablonski, D., and Raup, D. M. 1995. Selectivity of End-Cretaceous marine bivalve extinctions. Science, 268:389391.Google Scholar
Jablonski, D., and Sepkoski, J. J. 1996. Paleobiology, community ecology, and scales of ecological pattern. Ecology, 77:13671378.Google Scholar
Jackson, J. B. C., and Erwin, D. H. 2006. What can we learn about ecology and evolution from the fossil record? Trends in Ecology & Evolution, 21:322328.Google Scholar
Jackson, J. B. C., Kirby, M. X., Berger, W. H., Bjorndal, K. A., Botsford, L. W., Bourque, B. J., Bradbury, R. H., Cooke, R., Erlandson, J., Estes, J. A., Hughes, T. P., Kidwell, S., Lange, C. B., Lenihan, H. S., Pandolfi, J. M., Peterson, C. H., Steneck, R. S., Tegner, M. J., and Warner, R. R. 2001. Historical overfishing and the recent collapse of coastal ecosystems. Science, 293:629638.Google Scholar
Jiang, S. J., Bralower, T. J., Patzkowsky, M. E., Kump, L. R., and Schueth, J. D. 2010. Geographic controls on nannoplankton extinction across the Cretaceous/Palaeogene boundary. Nature Geoscience, 3:280285.Google Scholar
Joachimski, M. M., Pancost, R. D., Freeman, K. H., Ostertag-Henning, C., and Buggisch, W. 2002. Carbon isotope geochemistry of the Frasnian–Famennian transition. Palaeogeography, Palaeoclimatology, Palaeoecology, 181:91109.Google Scholar
Katz, M. E., Cramer, B. S., Franzese, A., Honisch, B., Miller, K. G., Rosenthal, Y., and Wright, J. D. 2010. Traditional and emerging geochemical proxies in foraminifera. Journal of Foraminiferal Research, 40:165192.Google Scholar
Keller, G., Adatte, T., Pardo, A., Bajpai, S., Khosla, A., and Samant, B. 2010. Cretaceous extinctions: Evidence overlooked. Science, 328:974975.Google Scholar
Kershaw, S., Crasquin, S., Li, Y., Collin, P. Y., Forel, M. B., Mu, X., Baud, A., Wang, Y., Xie, S., Maurer, F., and Guo, L. 2012. Microbialites and global environmental change across the Permian–Triassic boundary: a synthesis. Geobiology, 10:2547.Google Scholar
Kidwell, S. M. 2013. Time-averaging and fidelity of modern death assemblages: Building a taphonomic foundation for conservation paleobiology. Palaeontology, 56:487522.Google Scholar
Kidwell, S. M., and Flessa, K. W. 1995. The quality of the fossil record: populations, species, and communities. Annual Review of Ecology and Systematics, 26:269299.Google Scholar
Kiessling, W. 2002. Secular variations in the Phanerozoic reef ecosystem, p. 625690. In Kiessling, W., Flügel, E., and Golonka, J. (eds.), Phanerozoic Reef Patterns. SEPM Special Publication 72, Tulsa, OK.Google Scholar
Kiessling, W. 2006. Towards an unbiased estimate of fluctuations in reef abundance and volume during the Phanerozoic. Biogeosciences, 3:1527.Google Scholar
Kiessling, W. 2009. Geologic and biologic controls on the evolution of reefs. Annual Review of Ecology Evolution and Systematics, 40:173192.Google Scholar
Kiessling, W., and Flügel, E. 2002. Paleoreefs: a database on Phanerozoic reefs, p. 7795. In Kiessling, W., Flügel, E., and Golonka, J. (eds.), Phanerozoic Reef Patterns. SEPM Special Publication 72, Tulsa, OK.Google Scholar
Kiessling, W., Flügel, E., and Golonka, J. 1999. Paleoreef maps: Evaluation of a comprehensive database on Phanerozoic reefs. AAPG Bulletin, 83:15521587.Google Scholar
Kiessling, W., Flügel, E., and Golonka, J. 2000. Fluctuations in the carbonate production of Phanerozoic reefs, p. 191215. In Insalaco, E., Skelton, P. W., and Palmer, T. J. (eds.), Carbonate Platform Systems: Components and Interactions, Geological Society of London Special Publication 178, the Geological Society of London, London.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
Klug, J. L., Fischer, J. M., Ives, A. R., and Dennis, B. 2000. Compensatory dynamics in planktonic community responses to pH perturbations. Ecology, 81:387398.Google Scholar
Knoll, A. H., Bambach, R. K., Canfield, D. E., and Grotzinger, J. P. 1996. Comparative earth history and Late Permian mass extinction. Science, 273:452457.Google Scholar
Knoll, A. H., Bambach, R. K., Payne, J. L., Pruss, S., and Fischer, W. W. 2007. Paleophysiology and end-Permian mass extinction. Earth and Planetary Science Letters, 256:295313.Google Scholar
Korte, C., Kozur, H. W., Joachimski, M. M., Strauss, H., Veizer, J., and Schwark, L. 2004. Carbon, sulfur, oxygen and strontium isotope records, organic geochemistry and biostratigraphy across the Permian/Triassic boundary in Abadeh, Iran. International Journal of Earth Sciences, 93:565581.Google Scholar
Krug, A. Z., and Patzkowsky, M. E. 2007. Geographic variation in turnover and recovery from the Late Ordovician mass extinction. Paleobiology, 33:435454.Google Scholar
Kump, L. R. 1991. Interpreting carbon-isotope excursions: Strangelove Oceans. Geology, 19:299302.Google Scholar
Kump, L. R., and Arthur, M. A. 1999. Interpreting carbon-isotope excursions: carbonates and organic matter. Chemical Geology, 161:181198.Google Scholar
Kump, L. R., Arthur, M. A., Patzkowsky, M. E., Gibbs, M. T., Pinkus, D. S., and Sheehan, P. M. 1999. A weathering hypothesis for glaciation at high atmospheric pCO2 during the Late Ordovician. Palaeogeography, Palaeoclimatology, Palaeoecology, 152:173187.Google Scholar
LaFlamme, M., Darroch, S. A. F., Tweedt, S. M., Peterson, K. J., and Erwin, D. H. 2013. The end of the Ediacara biota: extinction, biotic replacement, or Cheshire Cat? Gondwana Research, 23:558573.Google Scholar
Leaky, R., and Lewin, R. 1992. The Sixth Extinction: Patterns of Life and the Future of Humankind. Anchor Press, New York.Google Scholar
Legendre, P., and Legendre, L. 1998. Numerical Ecology. Elsevier Science B.V., Amsterdam.Google Scholar
Lenton, T. M., Crouch, M., Johnson, M., Pires, N., and Dolan, L. 2012. First plants cooled the Ordovician. Nature Geoscience, 5:8689.Google Scholar
Leopold, A. 1939. A biotic view of the land. Journal of Forestry, 37:727730.Google Scholar
Levin, S. A. 1992. The problem of pattern and scale in ecology. Ecology, 73:19431967.Google Scholar
Lindeman, R. L. 1942. The trophic-dynamic aspect of ecology. Ecology, 23:399418.Google Scholar
Longrich, N. R., Bhullar, B. A. S., and Gauthier, J. A. 2012. Mass extinction of lizards and snakes at the Cretaceous–Paleogene boundary. Proceedings of the National Academy of Sciences of the United States of America, 109:2139621401.Google Scholar
Looy, C. V., Twitchett, R. J., Dilcher, D. L., Konijnenburg-Van Cittert, J. H. A., and Visscher, H. 2001. Life in the end-Permian dead zone. Proceedings of the National Academy of Sciences of the United States of America, 98:78797883.Google Scholar
Lopes Dos Santos, R. A., De Deckker, P., Hopmans, E. C., Magee, J. W., Mets, A., Sinnighe Damste, J. S., and Schouten, S. 2013. Abrupt vegetation change after the Late Quaternary megafaunal extinction in southeastern Australia. Nature Geoscience, 6:627631.Google Scholar
Loreau, M., Naeem, S., Inchausti, P., Bengtsson, J., Grime, J. P., Hector, A., Hooper, D. U., Huston, M. A., Raffaelli, D., Schmid, B., Tilman, D., and Wardle, D. A. 2001. Biodiversity and ecosystem functioning: Current knowledge and future challenges. Science, 294:804808.Google Scholar
Lotze, H. K., Coll, M., and Dunne, J. A. 2011. Historical changes in marine resources, food-web structure and ecosystem functioning in the Adriatic Sea, Mediterranean. Ecosystems, 14:198222.Google Scholar
MacLeod, K. G., Huber, B. T., and Fullagar, P. D. 2001. Evidence for a small (∼0.000030) but resolvable increase in seawater Sr87/Sr86 ratios across the Cretaceous–Tertiary boundary. Geology, 29:303306.Google Scholar
Magurran, A. E. 2004. Measuring Biological Diversity. Blackwell Publishing, Oxford.Google Scholar
Malkowski, K., and Racki, G. 2009. A global biogeochemical perturbation across the Silurian–Devonian boundary: Ocean-continent-biosphere feedbacks. Palaeogeography, Palaeoclimatology, Palaeoecology, 276:244254.Google Scholar
Mander, L., Kurschner, W. M., and McElwain, J. C. 2010. An explanation for conflicting records of Triassic–Jurassic plant diversity. Proceedings of the National Academy of Sciences of the United States of America, 107:1535115356.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, 165:319332.Google Scholar
Martin, E. E., and MacDougall, J. D. 1991. Seawater Sr Isotopes at the Cretaceous Tertiary Boundary. Earth and Planetary Science Letters, 104:166180.Google Scholar
Martindale, R. C., Berelson, W. M., Corsetti, F. A., Bottjer, D. J., and West, A. J. 2012. Constraining carbonate chemistry at a potential ocean acidification event (the Triassic–Jurassic boundary) using the presence of corals and coral reefs in the fossil record. Palaeogeography, Palaeoclimatology, Palaeoecology, 350:114123.Google Scholar
Mata, S. A., and Bottjer, D. J. 2011. Origin of Lower Triassic microbialites in mixed carbonatesiliciclastic successions: Ichnology, applied stratigraphy, and the end-Permian mass extinction. Palaeogeography, Palaeoclimatology, Palaeoecology, 300:158178.Google Scholar
Mata, S. A., and Bottjer, D. J. 2012. Microbes and mass extinctions: paleoenvironmental distribution of microbialites during times of biotic crisis. Geobiology, 10:324.Google Scholar
McElwain, J. C., Popa, M. E., Hesselbo, S. P., Haworth, M., and Surlyk, F. 2007. Macroecological responses of terrestrial vegetation to climatic and atmospheric change across the Triassic/Jurassic boundary in East Greenland. Paleobiology, 33:547573.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
McGhee, G. R. Jr., 1996. The Late Devonian Mass Extinction: the Frasnian/Famennian Crisis. Columbia University Press, New York.Google Scholar
McGhee, G. R. Jr., Sheehan, P. M., Bottjer, D. J., and Droser, M. L. 2004. Ecological ranking of Phanerozoic biodiversity crises: ecological and taxonomic severities are decoupled. Palaeogeography, Palaeoclimatology, Palaeoecology, 211:289297.Google Scholar
McGhee, G. R., Sheehan, P. M., Bottjer, D. J., and Droser, M. L. 2012. Ecological ranking of Phanerozoic biodiversity crises: The Serpukhovian (early Carboniferous) crisis had a greater ecological impact than the end-Ordovician. Geology, 40:147150.Google Scholar
Meyer, K. M., Yu, M., Jost, A. B., Kelley, B. M., and Payne, J. L. 2011. δ13C evidence that high primary productivity delayed recovery from the end-Permian mass extinction. Earth and Planetary Science Letters, 302:378384.Google Scholar
Meyer, K. M., Yu, M., Lehrmann, D., van de Schootbrugge, B., and Payne, J. L. 2013. Constraints on Early Triassic carbon cycle dynamics from paired organic and inorganic carbon isotope records. Earth and Planetary Science Letters, 361:429435.Google Scholar
Meyers, S. R. 2007. Production and preservation of organic matter: The significance of iron. Paleoceanography, 22:PA4211, doi:10.1029/2006PA001332 Google Scholar
Meyers, S. R., and Peters, S. E. 2011. A 56 million year rhythm in North American sedimentation during the Phanerozoic. Earth and Planetary Science Letters, 303:174180.Google Scholar
Misra, S., and Froelich, P. N. 2012. Lithium isotope history of Cenozoic seawater: Changes in silicate weathering and reverse weathering. Science, 335:818823.Google Scholar
Mitchell, J. S., Roopnarine, P. D., and Angielczyk, K. D. 2012. Late Cretaceous restructuring of terrestrial communities facilitated the end-Cretaceous mass extinction in North America. Proceedings of the National Academy of Sciences of the United States of America, 109:1885718861.Google Scholar
Moore, T. C. Jr., Jarrard, R. D., Olivarez Lyle, A., and Lyle, M. 2008. Eocene biogenic silica accumulation rates at the Pacific equatorial divergence zone. Paleoceanography, 23:PA2202, doi:10.1029/2007PA001514 Google Scholar
Morgan, J., Warner, M., Brittan, J., Buffler, R., Camargo, A., Christeson, G., Denton, P., Hildebrand, A., Hobbs, R., MacIntyre, H., MacKenzie, G., Maguire, P., Marin, L., Nakamura, Y., Pilkington, M., Sharpton, V., Snyder, D., Suarez, G., and Trejo, A. 1997. Size and morphology of the Chicxulub impact crater. Nature, 390:472476.Google Scholar
Morrow, J. R., and Hasiotis, S. T. 2007. Endobenthic response through mass extinction episodes: predictive models and observed patterns, p. 575611. In Miller, W. I. (ed.), Trace Fossils: Concepts, Problems, Prospects. Elsevier, Amsterdam.Google Scholar
Mukhopadhyay, S., Farley, K. A., and Montanari, A. 2001. A short duration of the Cretaceous–Tertiary boundary event: Evidence from extraterrestrial helium-3. Science, 291:19521955.Google Scholar
Mukhopadhyay, S., and Kreycik, P. 2008. Dust generation and drought patterns in Africa from helium-4 in a modern Cape Verde coral. Geophysical Research Letters, 35:L20820, doi:10.1029/2008GL035722 Google Scholar
Myers, N. 1990. Mass extinctions:what can the past tell us about the present and the future? Global and Planetary Change, 82:175185.Google Scholar
Newell, N. D. 1962. Paleontological gaps and geochronology. Journal of Paleontology, 36:592610.Google Scholar
Newell, N. D. 1967. Revolutions in the history of life, p. 6391 In Albritton, C. (ed.), Uniformity and Simplicity. Geological Society of America Special Paper, 89, Geological Society of America, Boulder, Colorado.Google Scholar
Norris, R. D. 1996. Symbiosis as an evolutionary innovation in the radiation of Paleocene planktic foraminifera. Paleobiology, 22:461480.Google Scholar
Novack-Gottshall, P. M. 2007. Using a theoretical ecospace to quantify the ecological diversity of Paleozoic and modern marine biotas. Paleobiology, 33:273294.Google Scholar
Olszewski, T. D., and Erwin, D. H. 2004. Dynamic response of Permian brachiopod communities to long-term environmental change. Nature, 428:738741.Google Scholar
Ourisson, G., Rohmer, M., and Poralla, K. 1987. Prokaryotic hopanoids and other polyterpenoid sterol surrogates. Annual Review of Microbiology, 41:301333.Google Scholar
Pandolfi, J. M., Bradbury, R. H., Sala, E., Hughes, T. P., Bjorndal, K. A., Cooke, R. G., McArdle, D., McClenachan, L., Newman, M. J. H., Paredes, G., Warner, R. R., and Jackson, J. B. C. 2003. Global trajectories of the long-term decline of coral reef ecosystems. Science, 301:955958.Google Scholar
Payne, J. L. 2005. Evolutionary dynamics of gastropod size across the end-Permian extinction and through the Triassic recovery interval. Paleobiology, 31:269290.Google Scholar
Payne, J. L., and Clapham, M. E. 2012. End-Permian mass extinction in the oceans: An ancient analog for the Twenty-first Century? Annual Review of Earth and Planetary Sciences, 40:89111.Google Scholar
Payne, J. L., and Kump, L. R. 2007. Evidence for recurrent Early Triassic massive volcanism from quantitative interpretation of carbon isotope fluctuations. Earth and Planetary Science Letters, 256:264277.Google Scholar
Payne, J. L., Lehrmann, D. J., Wei, J. Y., Orchard, M. J., Schrag, D. P., and Knoll, A. H. 2004. Large perturbations of the carbon cycle during recovery from the end-Permian extinction. Science, 305:506509.Google Scholar
Payne, J. L., Summers, M., Rego, B. L., Altiner, D., Wei, J. Y., Yu, M. Y., and Lehrmann, D. J. 2011. Early and Middle Triassic trends in diversity, evenness, and size of foraminifers on a carbonate platform in south China: implications for tempo and mode of biotic recovery from the end-Permian mass extinction. Paleobiology, 37:409425.Google Scholar
Peters, S. E. 2005. Geologic constraints on the macroevolutionary history of marine animals. Proceedings of the National Academy of Sciences of the United States of America, 102:1232612331.Google Scholar
Peters, S. E. 2008. Environmental determinants of extinction selectivity in the fossil record. Nature, 454:626629.Google Scholar
Peters, S. E., and Foote, M. 2002. Determinants of extinction in the fossil record. Nature, 416:420424.Google Scholar
Peters, S. E., and Heim, N. A. 2011. Macrostratigraphy and macroevolution in marine environments: testing the common-cause hypothesis, p. 95104. In McGowan, A. J. and Smith, A. B. (eds.), Comparing the Geological and Fossil Records: Implications for Biodiversity Studies. Geological Society Special Publication 358, The Geological Society of London.Google Scholar
Powell, M. G., and Kowalewski, M. 2002. Increase in evenness and sampled alpha diversity through the Phanerozoic: Comparison of early Paleozoic and Cenozoic marine fossil assemblages. Geology, 30:331334.Google Scholar
Pruss, S. B., and Bottjer, D. J. 2004. Early Triassic trace fossils of the western United States and their implications for prolonged environmental stress from the end-Permian mass extinction. PALAIOS, 19:551564.Google Scholar
Pruss, S., Fraiser, M., and Bottjer, D. J. 2004. Proliferation of Early Triassic wrinkle structures: Implications for environmental stress following the end-Permian mass extinction. Geology, 32:461464.Google Scholar
Rae, J. W. B. 2011. Boron isotopes and B/Ca in benthic foraminifera: proxies for the deep ocean carbonate system. Earth and Planetary Science Letters, 302:403413.Google Scholar
Rasmussen, C. M. O., and Harper, D. A. T. 2011. Did the amalgamation of continents drive the end Ordovician mass extinctions? Palaeogeography, Palaeoclimatology, Palaeoecology, 311:4862.Google Scholar
Raup, D. M., and Sepkoski, J. J. 1982. Mass extinctions in the marine fossil record. Science, 215:15011503.Google Scholar
Ravizza, G., and Peucker-Ehrenbrink, B. 2003. Chemostratigraphic evidence of Deccan volcanism from the marine osmium isotope record. Science, 302:13921395.Google Scholar
Rego, B. L., Wang, S. C., Altiner, D., and Payne, J. L. 2012. Within- and among-genus components of size evolution during mass extinction, recovery, and background intervals: a case study of Late Permian through Late Triassic foraminifera. Paleobiology, 38:627643.Google Scholar
Renne, P. R., Deino, A. L., Hilgen, F. J., Kuiper, K. F., Mark, D. F., Mitchell, W. S., Morgan, L. E., Mundil, R., and Smit, J. 2013. Time scales of critical events around the Cretaceous–Paleogene boundary. Science, 339:684687.Google Scholar
Retallack, G. J. 1997. Early forest soils and their role in Devonian global change. Science, 276:583585.Google Scholar
Richoz, S., Van De Schootbrugge, B., Pross, J., Puttmann, W., Quan, T. M., Lindstrom, S., Heunisch, C., Fiebig, J., Maquil, R., Schouten, S., Hauzenberger, C. A., and Wignall, P. B. 2012. Hydrogen sulphide poisoning of shallow seas following the end-Triassic extinction. Nature Geoscience, 5:662667.Google Scholar
Ridgwell, A. 2005. A Mid Mesozoic Revolution in the regulation of ocean chemistry. Marine Geology, 217:339357.Google Scholar
Ridgwell, A. 2007. Application of sediment core modelling to interpreting the glacial-interglacial record of Southern Ocean silica cycling. Climate of the Past, 3:387396.Google Scholar
Ridgwell, A., and Schmidt, D. N. 2010. Past constraints on the vulnerability of marine calcifiers to massive carbon dioxide release. Nature Geoscience, 3:196200.Google Scholar
Riding, R., and Liang, L. Y. 2005. Geobiology of microbial carbonates: metazoan and seawater saturation state influences on secular trends during the Phanerozoic. Palaeogeography, Palaeoclimatology, Palaeoecology, 219:101115.Google Scholar
Riede, J. O., Binzer, A., Brose, U., De Castro, F., Curtsdotter, A., Rall, B. C., and Eklof, A. 2011. Size-based food web characteristics govern the response to species extinctions. Basic and Applied Ecology, 12:581589.Google Scholar
Robinson, N., Ravizza, G., Coccioni, R., Peucker-Ehrenbrink, B., and Norris, R. 2009. A high-resolution marine Os187/Os188 record for the late Maastrichtian: Distinguishing the chemical fingerprints of Deccan volcanism and the K–P impact event. Earth and Planetary Science Letters, 281:159168.Google Scholar
Rohling, E. J., and Cooke, S. 1999. Stable oxygen and carbon isotopes in foraminiferal carbonate shells, p. 239258. In Sen Gupta, B. K. (ed.), Modern Foraminifera. Kluwer Academic Publishers, Great Britain.Google Scholar
Romano, C., Goudemand, N., Vennemann, T. W., Ware, D., Schneebeli-Hermann, E., Hochuli, P. A., Bruhwiler, T., Brinkmann, W., and Bucher, H. 2013. Climatic and biotic upheavals following the end-Permian mass extinction. Nature Geoscience, 6:5760.Google Scholar
Roopnarine, P. D. 2006. Extinction cascades and catastrophe in ancient food webs. Paleobiology, 32:119.Google Scholar
Roopnarine, P. D. 2009. Ecological modeling of paleocommunity food webs, p. 195220. In Dietl, G. P. and Flessa, K. W. (eds.), Conservation Paleobiology: Using the Past to Manage for the Future. The Paleontology Society Papers 15, Yale Press, New Haven.Google Scholar
Roopnarine, P. D. 2010. Networks, extinction, and paleocommunity food webs, p. 143161. In Alroy, J. and Hunt, G. (eds.), Quantitative Methods in Paleobiology. Paleontological Society Short Course. The Paleontology Society Papers 16, Yale Press, New Haven.Google Scholar
Roy, K., Jablonski, D., and Valentine, J. W. 2004. Beyond species richness: Biogeographic patterns and biodiversity dynamics using other metrics of diversity, p. 151170. In Lomolino, M. V. and Heaney, L. R. (eds.), Frontiers of Biogeography: New Directions in the Geography of Nature. Sinauer, Sunderland, MA.Google Scholar
Roy, K., Valentine, J. W., Jablonski, D., and Kidwell, S. M. 1996. Scales of climatic variability and time averaging in Pleistocene biotas: Implications for ecology and evolution. Trends in Ecology & Evolution, 11:458463.Google Scholar
Royer, D. L., Berner, R. A., and Beerling, D. J. 2001. Phanerozoic atmospheric CO2 change: evaluating geochemical and paleobiological approaches. Earth-Science Reviews, 54:349392.Google Scholar
Sageman, B. B., and Hollander, D. J. 1999. Cross correlation of paleoecological and geochemical proxies: a holistic approach to the study of past global change, p. 365384. In Barrera, E. and Johnson, C. C. (eds.), Evolution of the Cretaceous Ocean-Climate System. Geological Society of America Special Papers 332, Boulder, Colorado.Google Scholar
Sahney, S., and Benton, M. J. 2008. Recovery from the most profound mass extinction of all time. Proceedings of the Royal Society B-Biological Sciences, 275:759765.Google Scholar
Sahney, S., Benton, M. J., and Ferry, P. A. 2010. Links between global taxonomic diversity, ecological diversity and the expansion of vertebrates on land. Biology Letters, 6:544547.Google Scholar
Sallan, L. C., Kammer, T. W., Ausich, W. I., and Cook, L. A. 2011. Persistent predator-prey dynamics revealed by mass extinction. Proceedings of the National Academy of Sciences of the United States of America, 108:83358338.Google Scholar
Sano, H., and Nakashima, K. 1997. Lowermost Triassic (Griesbachian) microbial bindstonecementstone facies, southwest Japan. Facies, 36:124.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.Google Scholar
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, 323:2739.Google Scholar
Schindewolf, O. H. 1954. Über die möglichen Ursachen der grossen erdgeschichtlichen Faunenschnitte. Neues Jahrbuch fur Geologie und Palaontologie Monatshefte: 457465.Google Scholar
Schmidt, D. N., Lazarus, D., Young, J. R., and Kucera, M. 2006. Biogeography and evolution of body size in marine plankton. Earth-Science Reviews, 78:239266.Google Scholar
Schmitz, O. J., and Booth, G. 1997. Modelling food web complexity: The consequences of individual-based, spatially explicit behavioural ecology on trophic interactions. Evolutionary Ecology, 11:379398.Google Scholar
Schmitz, O. J., Grabowski, J. H., Peckarsky, B. L., Preisser, E. L., Trussell, G. C., and Vonesh, J. R. 2008. From individuals to ecosystem function: Toward an integration of evolutionary and ecosystem ecology. Ecology, 89:24362445.Google Scholar
Schmitz, O. J., Hawlena, D., and Trussell, G. C. 2010. Predator control of ecosystem nutrient dynamics. Ecology Letters, 13:11991209.Google Scholar
Schneebeli-Hermann, E., Hochuli, P. A., Bucher, H., Goudemand, N., Bruhwiler, T., and Galfetti, T. 2012. Palynology of the Lower Triassic succession of Tulong, South Tibet—Evidence for early recovery of gymnosperms. Palaeogeography, Palaeoclimatology, Palaeoecology, 339:1224.Google Scholar
Schubert, J. K., and Bottjer, D. J. 1992. Early Triassic stromatolites as post mass extinction disaster forms. Geology, 20:883886.Google Scholar
Schulte, P., Alegret, L., Arenillas, I., Arz, J. A., Barton, P. J., Bown, P. R., Bralower, T. J., Christeson, G. L., Claeys, P., Cockell, C. S., Collins, G. S., Deutsch, A., Goldin, T. J., Goto, K., Grajales-Nishimura, J. M., Grieve, R. A. F., Gulick, S. P. S., Johnson, K. R., Kiessling, W., Koeberl, C., Kring, D. A., MacLeod, K. G., Matsui, T., Melosh, J., Montanari, A., Morgan, J. V., Neal, C. R., Nichols, D. J., Norris, R. D., Pierazzo, E., Ravizza, G., Rebolledo-Vieyra, M., Reimold, W. U., Robin, E., Salge, T., Speijer, R. P., Sweet, A. R., Urrutia-Fucugauchi, J., Vajda, V., Whalen, M. T., and Willumsen, P. S. 2010. The Chicxulub asteroid impact and mass extinction at the Cretaceous–Paleogene Boundary. Science, 327:12141218.Google Scholar
Sephton, M. A., Looy, C. V., Brinkhuis, H., Wignall, P. B., de Leeuw, J. W., and Visscher, H. 2005. Catastrophic soil erosion during the end-Permian biotic crisis. Geology, 33:941944.Google Scholar
Sepkoski, J. J. Jr. 1981. A factor analytic description of the Phanerozoic marine fossil record. Paleobiology, 7:3653.Google Scholar
Sepkoski, J. J. Jr. 1986. Phanerozoic overview of mass extinction, p. 277295. In Raup, D. M. and Jablonski, D. (eds.), Patterns and Processes in the History of Life. Springer-Verlag, Berlin.Google Scholar
Sepkoski, J. J. Jr. 1989. Periodicity in extinction and the problem of catastrophism in the history of life. Journal of the Geological Society of London, 146:719.Google Scholar
Sepulveda, J., Wendler, J. E., Summons, R. E., and Hinrichs, K. U. 2009. Rapid resurgence of marine productivity after the Cretaceous–Paleogene mass extinction. Science, 326:129132.Google Scholar
Sessa, J. A., Patzkowsky, M. E., and Bralower, T. 2009. The impact of lithification on the diversity, size distribution, and recovery dynamics of marine invertebrate assemblages. Geology, 37:115118.Google Scholar
Sexton, P. F., Wilson, P. A., and Pearson, P. N. 2006. Microstructural and geochemical perspectives on planktic foraminiferal preservation: “Glassy” versus “Frosty.” Geochemistry, Geophysics, Geosystems, 7:Q12P19, doi: 10.1029/2006GC001291 Google Scholar
Sheehan, P. M. 2001. The Late Ordovician mass extinction. Annual Review of Earth and Planetary Sciences, 29:331364.Google Scholar
Sheehan, P. M., and Fastovsky, D. E. 1992. Major extinctions of land-dwelling vertebrates at the Cretaceous–Tertiary boundary, Eastern Montana. Geology, 20:556560.Google Scholar
Sheehan, P. M., and Hansen, T. A. 1986. Detritus feeding as a buffer to extinction at the end of the Cretaceous. Geology, 14:868870.Google Scholar
Shen, S. Z., Crowley, J. L., Wang, Y., Bowring, S. A., Erwin, D. H., Sadler, P. M., Cao, C. Q., Rothman, D. H., Henderson, C. M., Ramezani, J., Zhang, H., Shen, Y. N., Wang, X. D., Wang, W., Mu, L., Li, W. Z., Tang, Y. G., Liu, X. L., Liu, L. J., Zeng, Y., Jiang, Y. F., and Jin, Y. G. 2011. Calibrating the end-Permian mass extinction. Science, 334:13671372.Google Scholar
Smit, J. 1999. The global stratigraphy of the Cretaceous–Tertiary boundary impact ejecta. Annual Review of Earth and Planetary Sciences, 27:75113.Google Scholar
Smit, J., and Hertogen, J. 1980. An extraterrestrial event at the Cretaceous–Tertiary boundary. Nature, 285:198200.Google Scholar
Smith, A. B., Gale, A. S., and Monks, N. E. A. 2001. Sea-level change and rock-record bias in the Cretaceous: a problem for extinction and biodiversity studies. Paleobiology, 27:241253.Google Scholar
Sobolev, S. V., Sobolev, A. V., Kuzmin, D. V., Krivolutskaya, N. A., Petrunin, A. G., Arndt, N. T, Radko, V. A., and Vasiliev, Y. R. 2011. Linking mantle plumes, large igneous provinces and environmental catastrophes. Nature, 477:312–U380.Google Scholar
Sole, R. V., Montoya, J. M., and Erwin, D. H. 2002. Recovery after mass extinction: evolutionary assembly in large-scale biosphere dynamics. Philosophical Transactions of the Royal Society of London Series B-Biological Sciences, 357:697707.Google Scholar
Sole, R. V., Saldana, J., Montoya, J. M., and Erwin, D. H. 2010. Simple model of recovery dynamics after mass extinction. Journal of Theoretical Biology, 267:193200.Google Scholar
Sperling, E. A. 2013. Tackling the 99%: Can we begin to understand the paleoecology of the small and soft-bodied animal majority? p. 7786. In Bush, A. M., Pruss, S. B., and Payne, J. L. (eds.), Ecosystem Paleobiology and Geobiology, The Paleontological Society Papers 19. Yale Press, New Haven.Google Scholar
Spero, H. J. 1998. Life history and stable isotope geochemistry of planktonic foraminifera, p. 736. In Norris, R. D. and Corfield, R. M. (eds.), Isotope Paleobiology and Paleoecology. Paleontological Society Papers 4, The Paleontological Society, Pittsburgh, Pennsylvania.Google Scholar
Srinivasan, U. T., Dunne, J. A., Harte, J., and Martinez, N. D. 2007. Response of complex food webs to realistic extinction sequences. Ecology, 88:671682.Google Scholar
Summons, R. E., and Lincoln, S. A. 2012. Biomarkers: informative molecules for studies in geobiology, p. 269296. In Knoll, A. H., Canfield, D. E., and Konhauser, K. O. (eds.), Fundamentals of Geobiology. Wiley-Blackwell Publishing Ltd., Oxford.Google Scholar
Tansley, A. G. 1935. The use and abuse of vegetational concepts and terms. Ecology, 16:284307.Google Scholar
Thomas, E. 2007. Cenozoic mass extinctions in the deep sea: What perturbs the largest habitat on Earth?, p. 123. In Monechi, S., Coccioni, R., and Rampino, M. R. (eds.), Large Ecosystem Perturbations: Causes and Consequences: Geological Society of America Special Paper 424, The Geological Society, Boulder, Colorado.Google Scholar
Tobin, K. J., and Bergstrom, S. M. 2002. Implications of Ordovician (∼460 Myr) marine cement for constraining seawater temperature and atmospheric pCO2 . Palaeogeography, Palaeoclimatology, Palaeoecology, 181:399417.Google Scholar
Tobin, K. J., Bergstrom, S. M., and De La Garza, P. 2005. A mid-Caradocian (453 Ma) drawdown in atmospheric pCO2 without ice sheet development? Palaeogeography, Palaeoclimatology, Palaeoecology, 226:187204.Google Scholar
Tomasovych, A., and Kidwell, S. M. 2010a. The effects of temporal resolution on species turnover and on testing metacommunity models. American Naturalist, 175:587606.Google Scholar
Tomasovych, A., and Kidwell, S. M. 2010b. Predicting the effects of increasing temporal scale on species composition, diversity, and rank-abundance distributions. Paleobiology, 36:672695.Google Scholar
Tomasovych, A., and Kidwell, S. M. 2011. Accounting for the effects of biological variability and temporal autocorrelation in assessing the preservation of species abundance. Paleobiology, 37:332354.Google Scholar
Toon, O. B., Pollack, J. B., Ackerman, T. P., Turco, R. P., McKay, C. P., and Liu, M. S. 1982. Evolution of an impact-generated dust cloud and its effects on the atmosphere, p. 187200. In Silver, L. T. and Schultz, P. H. (eds.), Geological Implications of Impacts of Large Asteroids and Comets on the Earth: Geological Society of America Special Paper 190, Geological Society of America, Boulder, CO.Google Scholar
Toon, O. B., Zahnle, K., Morrison, D., Turco, R. P., and Covey, C. 1997. Environmental perturbations caused by the impacts of asteroids and comets. Reviews of Geophysics, 35:4178.Google Scholar
Twitchett, R. J. 1999. Palaeoenvironments and faunal recovery after the end-Permian mass extinction. Palaeogeography, Palaeoclimatology, Palaeoecology, 154:2737.Google Scholar
Twitchett, R. J. 2006. The palaeoclimatology, palaeoecology and palaeoenvironmental analysis of mass extinction events. Palaeogeography, Palaeoclimatology, Palaeoecology, 232:190213.Google Scholar
Twitchett, R. J., and Barras, C. G. 2004. Trace fossils in the aftermath of mass extinction events, p. 397418. In McIlroy, D. (ed.), Application of Ichnology to Palaeoenvironmental and Stratigraphic Analysis. The Geological Society of London Special Publication 228, The Geological Society, London.Google Scholar
Twitchett, R. J., Krystyn, L., Baud, A., Wheeley, J. R., and Richoz, S. 2004. Rapid marine recovery after the end-Permian mass-extinction event in the absence of marine anoxia. Geology, 32:805808.Google Scholar
Twitchett, R. J., Looy, C. V., Morante, R., Visscher, H., and Wignall, P. B. 2001. Rapid and synchronous collapse of marine and terrestrial ecosystems during the end-Permian biotic crisis. Geology, 29:351354.Google Scholar
Twitchett, R. J., and Wignall, P. B. 1996. Trace fossils and the aftermath of the Permo–Triassic mass extinction: Evidence from northern Italy. Palaeogeography, Palaeoclimatology, Palaeoecology, 124:137151.Google Scholar
Uhen, M. D. 2010. The Origin(s) of whales. Annual Review of Earth and Planetary Sciences, 38:189219.Google Scholar
Valdovinos, F. S., Ramos-Jiliberto, R., Garay-Narvaez, L., Urbani, P., and Dunne, J. A. 2010. Consequences of adaptive behaviour for the structure and dynamics of food webs. Ecology Letters, 13:15461559.Google Scholar
Van De Schootbrugge, B., Payne, J. L., Tomasovych, A., Pross, J., Fiebig, J., Benbrahim, M., Follmi, K. B., and Quan, T. M. 2008. Carbon cycle perturbation and stabilization in the wake of the Triassic–Jurassic boundary mass-extinction event. Geochemistry Geophysics Geosystems, 9:Q04028, doi:10.1029/2007gc001914 Google Scholar
Van De Schootbrugge, B., Quan, T. M., Lindstrom, S., Puttmann, W., Heunisch, C., Pross, J., Fiebig, J., Petschick, R., Rohling, H. G., Richoz, S., Rosenthal, Y., and Falkowski, P. G. 2009. Floral changes across the Triassic/Jurassic boundary linked to flood basalt volcanism. Nature Geoscience, 2:589594.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
Vandenbroucke, T. R. A., Armstrong, H. A., Williams, M., Paris, F., Zalasiewicz, J. A., Sabbe, K., Nolvak, J., Challandsa, T. J., Verniers, J., and Servais, T. 2010. Polar front shift and atmospheric CO2 during the glacial maximum of the Early Paleozoic Icehouse. Proceedings of the National Academy of Sciences of the United States of America, 107:1498314986.Google Scholar
Vermeij, G. J. 2004. Ecological avalanches and the two kinds of extinction. Evolutionary Ecology Research, 6:315337.Google Scholar
Villeger, S., Novack-Gottshall, P. M., and Mouillot, D. 2011. The multidimensionality of the niche reveals functional diversity changes in benthic marine biotas across geological time. Ecology Letters, 14:561568.Google Scholar
Villier, L., and Korn, D. 2004. Morphological disparity of ammonoids and the mark of Permian mass extinctions. Science, 306:264266.Google Scholar
Wade, B. S., and Twitchett, R. J. 2009. Extinction, dwarfing and the Lilliput effect Preface. Palaeogeography, Palaeoclimatology, Palaeoecology, 284:13.Google Scholar
Wagner, P. J., Kosnik, M. A., and Lidgard, S. 2006. Abundance distributions imply elevated complexity of post-Paleozoic marine ecosystems. Science, 314:12891292.Google Scholar
Wang, S. C. 2003. On the continuity of background and mass extinction. Paleobiology, 29:455467.Google Scholar
Wang, Y. 2004. Some trace fossils after the Frasnian–Famennian extinction in Dushan area, southern Guizhou Province, China. Acta Palaeontologica Sinica, 43:132141.Google Scholar
Wappler, T., Currano, E. D., Wilf, P., Rust, J., and Labandeira, C. C. 2009. No post-Cretaceous ecosystem depression in European forests? Rich insect-feeding damage on diverse middle Palaeocene plants, Menat, France. Proceedings of the Royal Society of London B-Biological Sciences, 276:42714277.Google Scholar
Webb, A. E., and Leighton, L. R. 2011. Exploring the ecological dynamics of extinction, p. 185220. In LaFlamme, M., Schiffbauer, J. D., and Dornbos, S. Q. (eds.), Quantifying the Evolution of Early Life. Topics in Geology 36, Springer, Dordrecht.Google Scholar
Westbroek, P. 1991. Life as a Geological Force: Dynamics of the Earth. W.W. Norton & Company, New York.Google Scholar
Whiteside, J. H., Olsen, P. E., Eglinton, T., Brookfield, M. E., and Sambrotto, R. N. 2010. 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 of the United States of America, 107:67216725.Google Scholar
Whiteside, J. H., and Ward, P. D. 2011. Ammonoid diversity and disparity track episodes of chaotic carbon cycling during the early Mesozoic. Geology, 39:99102.Google Scholar
Wignall, P. B., and Twitchett, R. J. 1996. Oceanic anoxia and the end Permian mass extinction. Science, 272:11551158.Google Scholar
Wilf, P., and Johnson, K. R. 2004. Land plant extinction at the end of the Cretaceous: A quantitative analysis of the North Dakota megafloral record. Paleobiology, 30:347368.Google Scholar
Williford, K. H., Foriel, J., Ward, P. D., and Steig, E. J. 2009. Major perturbation in sulfur cycling at the Triassic–Jurassic boundary. Geology, 37:835838.Google Scholar
Wood, R. 1999. Reef Evolution. Oxford University Press, Oxford.Google Scholar
Wood, R. 2004. Palaeoecology of a post-extinction reef: Famennian (Late Devonian) of the Canning Basin, north-western Australia. Palaeontology, 47:415445.Google Scholar
Woods, A. D., Bottjer, D. J., and Corsetti, F. A. 2007. Calcium carbonate seafloor precipitates from the outer shelf to slope facies of the Lower Triassic (Smithian–Spathian) Union Wash Formation, California, USA: Sedimentology and palaeobiologic significance. Palaeogeography, Palaeoclimatology, Palaeoecology, 252:281290.Google Scholar
Wootton, J. T. 2004. Markov chain models predict the consequences of experimental extinctions. Ecology Letters, 7:653660.Google Scholar
Worm, B., Barbier, E. B., Beaumont, N., Duffy, J. E., Folke, C., Halpern, B. S., Jackson, J. B. C., Lotze, H. K., Micheli, F., Palumbi, S. R., Sala, E., Selkoe, K. A., Stachowicz, J. J., and Watson, R. 2006. Impacts of biodiversity loss on ocean ecosystem services. Science, 314:787790.Google Scholar
Yapp, C. J., and Poths, H. 1996. Carbon isotopes in continental weathering environments and variations in ancient atmospheric CO2 pressure. Earth and Planetary Science Letters, 137:7182.Google Scholar
Yedid, G., Ofria, C. A., and Lenski, R. E. 2009. Selective press extinctions, but not random pulse extinction cause delayed ecological recovery in communities of digital organisms. American Naturalist, 173:139154.Google Scholar
Young, S. A., Saltzman, M. R., Foland, K. A., Linder, J. S., and Kump, L. R. 2009. A major drop in seawater 87Sr/86Sr during the Middle Ordovician (Darriwilian): Links to volcanism and climate? Geology, 37:951954.Google Scholar
Zachos, J. C., Arthur, M. A., and Dean, W. E. 1989. Geochemical evidence for suppression of pelagic marine productivity at the Cretaceous/Tertiary boundary. Nature, 337:6164.Google Scholar
Zachos, J., Pagani, M., Sloan, L., Thomas, E., and Billups, K. 2001. Trends, rhythms, and aberrations in global climate 65 Ma to present. Science, 292:686693.Google Scholar
Zhang, T. G., Shen, Y. N., Zhan, R. B., Shen, S. Z., and Chen, X. 2009. Large perturbations of the carbon and sulfur cycle associated with the Late Ordovician mass extinction in South China. Geology, 37:299302.Google Scholar
Zhang, Y. G., Pagani, M., Liu, Z., Bohaty, S. M., and Deconto, R. in press. A 40-million year history of atmospheric CO2 . Philosophical Transactions of the Royal Society of London A-Mathematical Physical and Engineering Sciences.Google Scholar
Zimov, S. A., Chuprynin, V. I., Oreshko, A. P., Chapin, F. S., Reynolds, J. F., and Chapin, M. C. 1995. Steppe-Tundra transition: A herbivore-driven biome shift at the end of the Pleistocene. American Naturalist, 146:765794.Google Scholar