Hostname: page-component-848d4c4894-4hhp2 Total loading time: 0 Render date: 2024-06-07T06:29:59.545Z Has data issue: false hasContentIssue false

A framework for the integrated analysis of the magnitude, selectivity, and biotic effects of extinction and origination

Published online by Cambridge University Press:  24 October 2019

Andrew M. Bush
Affiliation:
Department of Geosciences and Department of Ecology and Evolutionary Biology, University of Connecticut, 354 Mansfield Road, Unit 1045, Storrs, Connecticut06269. E-mail: andrew.bush@uconn.edu
Steve C. Wang
Affiliation:
Department of Mathematics and Statistics, Swarthmore College, Swarthmore, Pennsylvania19081. E-mail: scwang@swarthmore.edu
Jonathan L. Payne
Affiliation:
Department of Geological Sciences, Stanford University, Stanford, California94305. E-mail: jlpayne@stanford.edu
Noel A. Heim
Affiliation:
Department of Earth and Ocean Sciences, Tufts University, Lane Hall, Medford, Massachusetts02155. E-mail: noel.heim@tufts.edu

Abstract

The taxonomic and ecologic composition of Earth's biota has shifted dramatically through geologic time, with some clades going extinct while others diversified. Here, we derive a metric that quantifies the change in biotic composition due to extinction or origination and show that it equals the product of extinction/origination magnitude and selectivity (variation in magnitude among groups). We also define metrics that describe the extent to which a recovery (1) reinforced or reversed the effects of extinction on biotic composition and (2) changed composition in ways uncorrelated with the extinction. To demonstrate the approach, we analyzed an updated compilation of stratigraphic ranges of marine animal genera. We show that mass extinctions were not more selective than background intervals at the phylum level; rather, they tended to drive greater taxonomic change due to their higher magnitudes. Mass extinctions did not represent a separate class of events with respect to either strength of selectivity or effect. Similar observations apply to origination during recoveries from mass extinctions, and on average, extinction and origination were similarly selective and drove similar amounts of biotic change. Elevated origination during recoveries drove bursts of compositional change that varied considerably in effect. In some cases, origination partially reversed the effects of extinction, returning the biota toward the pre-extinction composition; in others, it reinforced the effects of the extinction, magnifying biotic change. Recoveries were as important as extinction events in shaping the marine biota, and their selectivity deserves systematic study alongside that of extinction.

Type
Featured Article
Copyright
Copyright © The Paleontological Society. All rights reserved 2019

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.)

Footnotes

Data available from the Dryad Digital Repository: https://doi.org/10.5061/dryad.mv97842

References

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 USA 109:728732.CrossRefGoogle Scholar
Alroy, J. 2000. New methods for quantifying macroevolutionary patterns and processes. Paleobiology 26:707733.2.0.CO;2>CrossRefGoogle Scholar
Alroy, J. 2004. Are Sepkoski's evolutionary faunas dynamically coherent? Evolutionary Ecology Research 6:132.Google Scholar
Alroy, J. 2008. Dynamics of origination and extinction in the fossil record. Proceedings of the National Academy of Sciences USA 105:1153611542.CrossRefGoogle Scholar
Alroy, J. 2014. Accurate and precise estimates of origination and extinction rates. Paleobiology 40:374397.CrossRefGoogle Scholar
Alroy, J. 2015. A more precise speciation and extinction rate estimator. Paleobiology 41:633639.CrossRefGoogle Scholar
Bambach, R. K. 2006. Phanerozoic biodiversity mass extinctions. Annual Review of Earth and Planetary Sciences 34:127–55.CrossRefGoogle Scholar
Bambach, R. K., Knoll, A. H., and Wang, S. C.. 2004. Origination, extinction, and mass depletions of marine diversity. Paleobiology 30:522542.2.0.CO;2>CrossRefGoogle Scholar
Becker, R. T., Kaiser, S. I., and Aretz, M.. 2016. Review of chrono-, litho- and biostratigraphy across the global Hangenberg Crisis and Devonian–Carboniferous boundary. In Becker, R. T., Königshof, P., and Brett, C. E., eds. Devonian climate, sea level and evolutionary events. Geological Society of London Special Publication 423:355386.Google Scholar
Benton, M. J. 1995. Diversification and extinction in the history of life. Science 268:5258.CrossRefGoogle ScholarPubMed
Bland, J. M., and Altman, D. G.. 2011. Correlation in restricted ranges of data. BMJ 342:d556.CrossRefGoogle Scholar
Bond, D. P., and Wignall, P. B.. 2008. The role of sea-level change and marine anoxia in the Frasnian–Famennian (Late Devonian) mass extinction. Palaeogeography, Palaeoclimatology, Palaeoecology 263:107118.CrossRefGoogle Scholar
Brayard, A., Escarguel, G., Bucher, H., Monnet, C., Brühwiler, T., Goudemand, N., Galfetti, T., and Guex, J.. 2009. Good genes and good luck: ammonoid diversity and the end-Permian mass extinction. Science 325:11181121.CrossRefGoogle Scholar
Bush, A. M., and Bambach, R. K.. 2011. Paleoecologic megatrends in marine Metazoa. Annual Review of Earth and Planetary Sciences 39:241269.CrossRefGoogle Scholar
Bush, A. M., and Brame, R. I.. 2010. Multiple paleoecological controls on the composition of marine fossil assemblages from the Frasnian (Late Devonian) of Virginia, with a comparison of ordination methods. Paleobiology 36:573591.CrossRefGoogle Scholar
Bush, A. M., and Pruss, S. B.. 2013. Theoretical ecospace for ecosystem paleobiology: energy, nutrients, biominerals, and macroevolution. In Bush, A. M., Pruss, S. B., and Payne, J. L., eds. Ecosystem paleobiology and geobiology. Paleontological Society Papers 19:120.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.CrossRefGoogle Scholar
Bush, A. M., Csonka, J. D., DiRenzo, G. V., Over, D. J., and Beard, J. A.. 2015. Revised correlation of the Frasnian-Famennian boundary and Kellwasser events (Upper Devonian) in shallow marine paleoenvironments of New York State. Palaeogeography, Palaeoclimatology, Palaeoecology 433:233246.CrossRefGoogle Scholar
Chen, Z.-Q., and Benton, M. J.. 2012. The timing and pattern of biotic recovery following the end-Permian mass extinction. Nature Geoscience 5:375.CrossRefGoogle Scholar
Christie, M., Holland, S. M., and Bush, A. M.. 2013. Contrasting the ecological and taxonomic consequences of extinction. Paleobiology 39:538559.CrossRefGoogle Scholar
Clapham, M. E. 2015. Ecological consequences of the Guadalupian extinction and its role in the brachiopod-mollusk transition. Paleobiology 41:266279.CrossRefGoogle Scholar
Clapham, M. E. 2017. Organism activity levels predict marine invertebrate survival during ancient global change extinctions. Global Change Biology 23:14771485.CrossRefGoogle ScholarPubMed
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.CrossRefGoogle Scholar
Clarke, K. R., and Warwick, R. M.. 2001. Change in marine communities: an approach to statistical analysis and interpretation. Primer-E, Plymouth, U.K.Google Scholar
Clarkson, M., Kasemann, S., Wood, R., Lenton, T., Daines, S., Richoz, S., Ohnemueller, F., Meixner, A., Poulton, S., and Tipper, E.. 2015. Ocean acidification and the Permo-Triassic mass extinction. Science 348:229232.CrossRefGoogle ScholarPubMed
Congreve, C. R., Krug, A. Z., and Patzkowsky, M. E.. 2018. Evolutionary and biogeographical shifts in response to the Late Ordovician mass extinction. Palaeontology 62:267285.CrossRefGoogle Scholar
Copper, P. 2002. Reef development at the Frasnian/Famennian mass extinction boundary. Palaeogeography, Palaeoclimatology, Palaeoecology 181:2765.CrossRefGoogle Scholar
Dineen, A. A., Fraiser, M. L., and Sheehan, P. M.. 2014. Quantifying functional diversity in pre-and post-extinction paleocommunities: a test of ecological restructuring after the end-Permian mass extinction. Earth-Science Reviews 136:339349.CrossRefGoogle Scholar
Droser, M. L., Bottjer, D. J., Sheehan, P. M., and McGhee, G. R. Jr. 2000. Decoupling of taxonomic and ecologic severity of Phanerozoic marine mass extinctions. Geology 28:675678.2.0.CO;2>CrossRefGoogle Scholar
Erwin, D. H., and Hua-Zhang, P.. 1996. Recoveries and radiations: gastropods after the Permo-Triassic mass extinction. In Hart, M. B., ed. Biotic recovery from mass extinction events. Geological Society of London Special Publication 102:223229.Google Scholar
Faith, D. P., Minchin, P. R., and Belbin, L.. 1987. Compositional dissimilarity as a robust measure of ecological distance. Vegetatio 69:5768.CrossRefGoogle 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 USA 109:68296834.CrossRefGoogle ScholarPubMed
Finnegan, S., Rasmussen, C. M. Ø., and Harper, D. A. T.. 2016. Biogeographic and bathymetric determinants of brachiopod extinction and survival during the Late Ordovician mass extinction. Proceedings of the Royal Society of London B 283:20160007.CrossRefGoogle ScholarPubMed
Foote, M. 1999. Morphological diversity in the evolutionary radiation of Paleozoic and post-Paleozoic crinoids. Paleobiology 25:1115.CrossRefGoogle Scholar
Foote, M. 2000. Origination and extinction components of taxonomic diversity: general problems. Paleobiology 26:74102.CrossRefGoogle Scholar
Foote, M. 2003. Origination and extinction through the Phanerozoic: a new approach. Journal of Geology 111:125148.CrossRefGoogle Scholar
Foote, M. 2005. Pulsed origination and extinction in the marine realm. Paleobiology 31:620.2.0.CO;2>CrossRefGoogle Scholar
Foote, M. 2007. Extinction and quiescence in marine animal genera. Paleobiology 33:261272.CrossRefGoogle Scholar
Foote, M. 2010. The geological history of biodiversity. Pp. 479510in Bell, M. A., Futuyma, D. J., Eanes, W. F., and Levinton, J. S., eds. Evolution since Darwin: the first 150 years. Sinauer, Sunderland, Mass.Google Scholar
Foster, W. J., and Twitchett, R. J.. 2014. Functional diversity of marine ecosystems after the Late Permian mass extinction event. Nature Geoscience 7:233238.CrossRefGoogle Scholar
Friedman, M. 2009. Ecomorphological selectivity among marine teleost fishes during the end-Cretaceous extinction. Proceedings of the National Academy of Sciences USA 106:52185223.CrossRefGoogle ScholarPubMed
Galfetti, T., Hochuli, P. A., Brayard, A., Bucher, H., Weissert, H., and Vigran, J. O.. 2007. Smithian–Spathian boundary event: evidence for global climatic change in the wake of the end-Permian biotic crisis. Geology 35:291294.CrossRefGoogle Scholar
Gilinksy, N. L. 1994. Volatility and the Phanerozoic decline of background extinction intensity. Paleobiology 20:445458.Google Scholar
Gould, S. J. 1984. Smooth curve of evolutionary rate: a psychological and mathematical artifact. Science 226:994995.CrossRefGoogle ScholarPubMed
Harnik, P. G. 2011. Direct and indirect effects of biological factors on extinction risk in fossil bivalves. Proceedings of the National Academy of Sciences USA 108:1359413599.CrossRefGoogle ScholarPubMed
Harnik, P. G., Simpson, C., and Payne, J. L.. 2012. Long-term differences in extinction risk among the seven forms of rarity. Proceedings of the Royal Society of London B 279:49694976.CrossRefGoogle ScholarPubMed
Heim, N. A., and Peters, S. E.. 2011. Regional environmental breadth predicts geographic range and longevity in fossil marine genera. PLoS ONE 6:e18946.CrossRefGoogle ScholarPubMed
Heim, N. A., Knope, M. L., Schaal, E. K., Wang, S. C., and Payne, J. L.. 2015. Cope's rule in the evolution of marine animals. Science 347:867870.CrossRefGoogle ScholarPubMed
Henehan, M. J., Hull, P. M., Penman, D. E., Rae, J. W., and Schmidt, D. N.. 2016. Biogeochemical significance of pelagic ecosystem function: an end-Cretaceous case study. Philosophical Transactions of the Royal Society of London B 371:20150510.CrossRefGoogle ScholarPubMed
Hull, P. 2015. Life in the aftermath of mass extinctions. Current Biology 25:R941R952.CrossRefGoogle ScholarPubMed
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.CrossRefGoogle Scholar
Hull, P. M., Darroch, S. A., and Erwin, D. H.. 2015. Rarity in mass extinctions and the future of ecosystems. Nature 528:345351.CrossRefGoogle ScholarPubMed
Jablonski, D. 1996. Body size and macroevolution. Pp. 256289in Jablonski, D., Erwin, D. H., and Lipps, J. H., eds. Evolutionary paleobiology. University of Chicago Press, Chicago.Google ScholarPubMed
Jablonski, D., and Raup, D. M.. 1995. Selectivity of end-Cretaceous marine bivalve extinctions. Science 268:389391.CrossRefGoogle ScholarPubMed
Janevski, G. A., and Baumiller, T. K.. 2009. Evidence for extinction selectivity throughout the marine invertebrate fossil record. Paleobiology 35:553564.CrossRefGoogle Scholar
Joachimski, M. M., and Buggisch, W.. 2002. Conodont apatite δ18O signatures indicate climatic cooling as a trigger of the Late Devonian mass extinction. Geology 30:711714.2.0.CO;2>CrossRefGoogle Scholar
Kiessling, W., and Aberhan, M.. 2007. Geographical distribution and extinction risk: lessons from Triassic–Jurassic marine benthic organisms. Journal of Biogeography 34:14731489.CrossRefGoogle 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.CrossRefGoogle Scholar
Kitchell, J. A., Clark, D. L., and Gombos, A. M. Jr. 1986. Biological selectivity of extinction: a link between background and mass extinction. Palaios 1:504511.CrossRefGoogle 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.CrossRefGoogle ScholarPubMed
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.CrossRefGoogle Scholar
Krug, A. Z., and Patzkowsky, M. E.. 2015. Phylogenetic clustering of origination and extinction across the Late Ordovician mass extinction. PLoS ONE 10:e0144354.CrossRefGoogle ScholarPubMed
Legendre, P., and Legendre, L.. 2012. Numerical ecology, 3rd ed. Elsevier, Amsterdam.Google Scholar
Leighton, L. R., and Schneider, C. L.. 2008. Taxon characteristics that promote survivorship through the Permian–Triassic interval: transition from the Paleozoic to the Mesozoic brachiopod fauna. Paleobiology 34:6578.CrossRefGoogle Scholar
Lockwood, R. 2003. Abundance not linked to survival across the end-Cretaceous mass extinction: patterns in North American bivalves. Proceedings of the National Academy of Sciences USA 100:24782482.CrossRefGoogle Scholar
Lockwood, R. 2004. The K/T event and infaunality: morphological and ecological patterns of extinction and recovery in veneroid bivalves. Paleobiology 30:507521.2.0.CO;2>CrossRefGoogle Scholar
Lockwood, R. 2005. Body size, extinction events, and the early Cenozoic record of veneroid bivalves: a new role for recoveries? Paleobiology 31:578590.CrossRefGoogle Scholar
Ma, X., Gong, Y., Chen, D., Racki, G., Chen, X., and Liao, W.. 2015. The Late Devonian Frasnian–Famennian Event in South China—patterns and causes of extinctions, sea level changes, and isotope variations. Palaeogeography, Palaeoclimatology, Palaeoecology 448:121.Google Scholar
Marynowski, L., Zatoń, M., Rakociński, M., Filipiak, P., Kurkiewicz, S., and Pearce, T. J.. 2012. Deciphering the upper Famennian Hangenberg black shale depositional environments based on multi-proxy record. Palaeogeography, Palaeoclimatology, Palaeoecology 346:6686.CrossRefGoogle 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.CrossRefGoogle Scholar
McGhee, G. R. Jr., 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.CrossRefGoogle Scholar
McKinney, M. L. 1987. Taxonomic selectivity and continuous variation in mass and background extinctions of marine taxa. Nature 325:143145.CrossRefGoogle Scholar
Newell, N. D. 1967. Revolutions in the history of life. In Albritton, J. C. C. Jr., ed. Uniformity and simplicity: a symposium on the principle of the uniformity of nature. Geological Society of America Special Paper 89:6391.CrossRefGoogle Scholar
Novack-Gottshall, P. M. 2007. Using a theoretical ecospace to quantify the ecological diversity of Paleozoic and modern marine biotas. Paleobiology 33:273294.CrossRefGoogle Scholar
Payne, J. L., and Finnegan, S.. 2007. The effect of geographic range on extinction risk during background and mass extinction. Proceedings of the National Academy of Sciences USA 104:1050610511.CrossRefGoogle ScholarPubMed
Payne, J. L., and van de Schootbrugge, B.. 2007. Life in Triassic oceans: links between planktonic and benthic recovery and radiation. Pp. 165189in Falkowski, P. G., and Knoll, A. H., eds. Evolution of primary producers in the sea. Academic Press, Burlington, Mass.CrossRefGoogle Scholar
Payne, J. L., Lehrmann, D. J., Wei, J., 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.CrossRefGoogle ScholarPubMed
Payne, J. L., Bush, A. M., Chang, E. T., Heim, N. A., Knope, M. L., and Pruss, S. B.. 2016a. Extinction intensity, selectivity, and their combined macroevolutionary influence in the fossil record. Biology Letters 12:20160202.CrossRefGoogle Scholar
Payne, J. L., Bush, A. M., Heim, N. A., Knope, M. L., and McCauley, D. J.. 2016b. Ecological selectivity of the emerging mass extinction in the oceans. Science 353:12841286.CrossRefGoogle Scholar
Penn, J. L., Deutsch, C., Payne, J. L., and Sperling, E. A.. 2018. Temperature-dependent hypoxia explains biogeography and severity of end-Permian marine mass extinction. Science 362:eaat1327.CrossRefGoogle ScholarPubMed
Petsios, E., Thompson, J. R., Pietsch, C., and Bottjer, D. J.. 2019. Biotic impacts of temperature before, during, and after the end-Permian extinction: a multi-metric and multi-scale approach to modeling extinction and recovery dynamics. Palaeogeography, Palaeoclimatology, Palaeoecology 513:8699.CrossRefGoogle Scholar
Pietsch, C., Ritterbush, K. A., Thompson, J. R., Petsios, E., and Bottjer, D. J.. 2019. Evolutionary models in the Early Triassic marine realm. Palaeogeography, Palaeoclimatology, Palaeoecology 513:6585.CrossRefGoogle Scholar
Powell, M. G. 2008. Timing and selectivity of the Late Mississippian mass extinction of brachiopod genera from the central Appalachian Basin. Palaios 23:525534.CrossRefGoogle Scholar
Powell, M. G., and MacGregor, J.. 2011. A geographic test of species selection using planktonic foraminifera during the Cretaceous/Paleogene mass extinction. Paleobiology 37:426437.CrossRefGoogle Scholar
Raup, D. M., and Sepkoski, J. J. Jr. 1982. Mass extinctions in the marine fossil record. Science 215:15011503.CrossRefGoogle ScholarPubMed
Reddin, C. J., Kocsis, Á. T., and Kiessling, W.. 2019. Climate change and the latitudinal selectivity of ancient marine extinctions. Paleobiology 45:7084.CrossRefGoogle Scholar
Rivadeneira, M. M., and Marquet, P. A.. 2007. Selective extinction of late Neogene bivalves on the temperate Pacific coast of South America. Paleobiology 33:455468.CrossRefGoogle Scholar
Romano, C., Goudemand, N., Vennemann, T. W., Ware, D., Schneebeli-Hermann, E., Hochuli, P. A., Brühwiler, T., Brinkmann, W., and Bucher, H.. 2013. Climatic and biotic upheavals following the end-Permian mass extinction. Nature Geoscience 6:5760.CrossRefGoogle Scholar
Roopnarine, P. D., Angielczyk, K. D., Wang, S. C., and Hertog, R.. 2007. Trophic network models explain instability of Early Triassic terrestrial communities. Proceedings of the Royal Society of London B 274:20772086.CrossRefGoogle ScholarPubMed
Sallan, L. C., and Coates, M. I.. 2010. End-Devonian extinction and a bottleneck in the early evolution of modern jawed vertebrates. Proceedings of the National Academy of Sciences USA 107:1013110135.CrossRefGoogle 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 USA 108:83358338.CrossRefGoogle ScholarPubMed
Sepkoski, J. J. Jr. 2002. A compendium of fossil marine animal genera. Bulletins of American Paleontology 363:1560.Google Scholar
Silverman, B. W. 1981. Using kernel density estimates to investigate multimodality. Journal of the Royal Statistical Society, Series B 43:9799.Google Scholar
Smith, A. B., and Jeffery, C. H.. 1998. Selectivity of extinction among sea urchins at the end of the Cretaceous period. Nature 392:6971.CrossRefGoogle Scholar
Solé, 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 B 357:697707.CrossRefGoogle ScholarPubMed
Song, H., Wignall, P. B., Tong, J., and Yin, H.. 2013. Two pulses of extinction during the Permian–Triassic crisis. Nature Geoscience 6:52.CrossRefGoogle Scholar
Song, H., Wignall, P. B., Chu, D., Tong, J., Sun, Y., Song, H., He, W., and Tian, L.. 2014. Anoxia/high temperature double whammy during the Permian–Triassic marine crisis and its aftermath. Scientific Reports 4:4132.CrossRefGoogle ScholarPubMed
Stanley, S. M. 1979. Macroevolution: pattern and process. Freeman, San Francisco.Google Scholar
Stanley, S. M. 2007. An analysis of the history of marine animal diversity. Paleobiology Memoirs 4:155.Google Scholar
Stanley, S. M. 2016. Estimates of the magnitudes of major marine mass extinctions in earth history. Proceedings of the National Academy of Sciences USA 113:E6325E6334.CrossRefGoogle ScholarPubMed
Stigall, A. L. 2012. Speciation collapse and invasive species dynamics during the Late Devonian “mass extinction.” GSA Today 22:49.CrossRefGoogle Scholar
Vilhena, D. A., Harris, E. B., Bergstrom, C. T., Maliska, M. E., Ward, P. D., Sidor, C. A., Strömberg, C. A. E., and Wilson, G. P.. 2013. Bivalve network reveals latitudinal selectivity gradient at the end-Cretaceous mass extinction. Scientific Reports 3:1790.CrossRefGoogle Scholar
Wang, S. C. 2003. On the continuity of background and mass extinction. Paleobiology 29:455467.2.0.CO;2>CrossRefGoogle Scholar
Wang, S. C. 2010. Principles of statistical inference: likelihood and the Bayesian paradigm. In Alroy, J. and Hunt, G., eds. Quantitative methods in paleobiology. Paleontological Society Papers 16:118.Google Scholar
Wignall, P., and Benton, M.. 1999. Lazarus taxa and fossil abundance at times of biotic crisis. Journal of the Geological Society 156:453456.CrossRefGoogle Scholar