Hostname: page-component-84b7d79bbc-fnpn6 Total loading time: 0 Render date: 2024-08-01T14:13:22.438Z Has data issue: false hasContentIssue false

Theoretical Ecospace for Ecosystem Paleobiology: Energy, Nutrients, Biominerals, and Macroevolution

Published online by Cambridge University Press:  21 July 2017

Andrew M. Bush
Affiliation:
Department of Ecology and Evolutionary Biology and Center for Integrative Geosciences, University of Connecticut, Storrs, CT 06269-3043
Sara B. Pruss
Affiliation:
Department of Geosciences, Smith College, Northampton, MA 01063
Get access

Abstract

Changes in nutrient cycles and energy fluxes (i.e., ecosystem dynamics) likely drove numerous trends and disruptions in the history of life. Advances in geochemistry offer great insights into paleoecosystem function, as does an understanding of the biogeochemical roles played by ancient organisms. A theoretical ecospace that describes the chemical exchanges between organisms and their environments is presented. Previous descriptions of ecospace principally described spatial and physical aspects of ecology; the new ecospace description broadens the concept to encompass a wider range of ecological processes that control abundance and diversity of fossil organisms. Organisms require materials from the environment for generating energy and building tissues, and these factors are broken down, ultimately specifying particular substances acquired from the environment. Different organisms require specific substances in different amounts depending on factors such as physiology, environmental conditions, etc.; thus, physiological ecospace describes an organism's sensitivity to ecosystem/earth system perturbations and trends. Several examples relating to organisms' requirements for skeletal minerals are reviewed, and a new analysis of extinction selectivity related to ocean acidification is presented. Selective extinction of heavily calcified metazoa is demonstrated to have occurred at least eight times during the Phanerozoic, including the early Cambrian, Frasnian (Late Devonian), and Aptian (Early Cretaceous). Multidimensional structure of ecospace occupation (e.g., correlations among ecological traits) strongly controls the effects of an extinction such that the same kill mechanism applied at different times will affect the ecological composition of the biosphere in a variety of ways.

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

Alroy, J., Aberhan, M., Bother, D. J., Foote, M., Fürsich, 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., Nürnberg, S., Powers, C. M., Sessa, J. A., Simpson, C., Tomašových, A., and Visaggi, C. C. 2008. Phanerozoic trends in the global diversity of marine invertebrates. Science, 321:97100.CrossRefGoogle ScholarPubMed
Alvarez, L. W., Alvarez, W., Asaro, F., and Michel, H. V. 1980. Extraterrestrial cause for the Cretaceous–Tertiary extinction. Science, 208:10951108.Google Scholar
Anbar, A. D., and Knoll, A. H. 2002. Proterozoic ocean chemistry and evolution: a bioorganic bridge? Science, 297:11371142.Google Scholar
Archer, D. 2005. Fate of fossil fuel CO2 in geologic time. Journal of Geophysical Research, 110:16.Google Scholar
Archer, D., Kheshgi, H., and Maier-Reimer, E. 1997. Multiple timescales for neutralization of fossil fuel CO2 . Geophysical Research Letters, 24:405408.CrossRefGoogle Scholar
Bambach, R. K. 1993. Seafood through time: changes in biomass, energetics, and productivity in the marine ecosystem. Paleobiology, 19:372397.CrossRefGoogle Scholar
Bambach, R. K. 1999. Energetics in the global marine fauna: a connection between terrestrial diversification and change in the marine biosphere. Geobios, 32:131144.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. Jr. 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
Beerling, D. J., and Royer, D. L. 2002. Reading a CO2 signal from fossil stomata. New Phytologist, 153:387397.Google Scholar
Blättler, C. L., Jenkyns, H. C., Reynard, L. M., and Henderson, G. M. 2011. Significant increases in global weathering during Oceanic Anoxic Events la and 2 indicated by calcium isotopes. Earth and Planetary Sciences Letters, 309:7788.Google Scholar
Bond, D., Wignall, P. B., and Racki, G. 2004. Extent and duration of marine anoxia during the Frasnian–Famennian (Late Devonian) mass extinction in Poland, Germany, Austria and France. Geological Magazine, 141:173193.Google Scholar
Boyce, C. K., and Lee, J.-E. 2010. An exceptional role for flowering plant physiology in the expansion of tropical rainforests and biodiversity. Proceedings of the Royal Society B, 277:34373443.CrossRefGoogle ScholarPubMed
Boyce, C. K., and Lee, J.-E. 2011. Could land plant evolution have fed the marine revolution? Paleontological Research, 15:100105.Google Scholar
Boyer, D. L., and Droser, M. L. 2009. Palaeoecological patterns within the dysaerobic biofacies: examples from Devonian black shales of New York State. Palaeogeography, Palaeoclimatology, Palaeoecology, 276:206216.CrossRefGoogle Scholar
Brasier, M., Green, O., and Shields, G. 1997. Ediacaran sponge spicule clusters from southwestern Mongolia and the origins of the Cambrian fauna. Geology, 25:303306.Google Scholar
Braun, A., Chen, J., Waloszek, D., and Maas, A. 2007. First Early Cambrian radiolaria, p. 143149. In Vickers-Rich, P. and Komarower, P. (eds.), The Rise and Fall of the Ediacaran Biota. Geological Society Special Publications 286, Geological Society of London, London.Google 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., 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., Bambach, R. K., and Erwin, D. H. 2011. Ecospace utilization during the Ediacaran radiation and the Cambrian eco-explosion. p. 111134. 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, New York.Google Scholar
Bush, A. M., and Novack-Gottshall, P. M. 2012. Modelling the ecological–functional diversification of marine Metazoa on geological time scales. Biology Letters, 8:151155.Google Scholar
Caplan, M. L., and Bustin, R. M. 1999. Devonian–Carboniferous Hangenberg mass extinction event, widespread organic-rich mudrock and anoxia: causes and consequences. Palaeogeography, Palaeoclimatology, Palaeoecology, 148:187207.Google Scholar
Chen, D., Qing, H., and Li, R. 2005. The Late Devonian Frasnian–Famennian (F/F) biotic crisis: insights from δ13CCarb, δ13Corg and 87Sr/86Sr isotopic systematics. Earth and Planetary Science Letters, 235:151166.Google Scholar
Chen, D., Tucker, M. E., Shen, Y., Yans, J., and Prea, A. 2002. Carbon isotope excursions and sea-level change: implications for the Frasnian–Famennian biotic crisis. Journal of the Geological Society of London, 159:623626.Google Scholar
Clapham, M. E., and Payne, J. L. 2011. Acidification, anoxia, and extinction: a multiple regression analysis of extinction selectivity during the middle and late Permian. Geology, 39:10591062.Google Scholar
Coggon, R. M., Teagle, D. A. H., Smith-Duque, C. E., Alt, J. C., and Cooper, M. J. 2010. Reconstructing past seawater Mg/Ca and Sr/Ca from mid-ocean ridge flank calcium carbonate veins. Science, 327:11141117.Google Scholar
Colosimo, A., Bralower, T. J., and Zachos, J. C. 2006. Evidence for lysocline shoaling at the Paleocene/Eocene Thermal Maximum on Shatsky Rise, Northwest Pacific. Proceedings of the Ocean Drilling Program, Scientific Results, 198:136.Google Scholar
Copper, P. 2002. Reef development at the Frasnian/ Famennian mass extinction boundary. Palaeogeography, Palaeoclimatology, Palaeoecology, 181:2765.Google Scholar
Courtillot, V., Kravchinsky, V. A., Quidelleur, X., Renne, P. R., and Gladkochub, D. P. 2010. Preliminary dating of the Viluy traps (Eastern Siberia): eruption at the time of Late Devonian extinction events? Earth and Planetary Science Letters, 300:239245.Google Scholar
Dahl, T. W., Hammarlund, E. U., Anbar, A. D., Bond, D. P. G., Gill, B. C., Gordon, G. W., Knoll, A. H., Nielsen, A. T., Schovsbo, N. H., and Canfield, D. E. 2010. Devonian rise in atmospheric oxygen correlated to the radiations of terrestrial plants and large predatory fish. Proceedings of the National Academy of Sciences, 107:1791117915.Google Scholar
Doney, S. C., Fabry, V. J., Feely, R. A., and Kleypas, J. A. 2009. Ocean acidification: the other CO2 problem. Annual Review of Marine Science, 1:169192.Google Scholar
Dornbos, S. Q. 2006. Evolutionary palaeoecology of early epifaunal echinoderms: response to increasing bioturbation levels during the Cambrian radiation. Palaeogeography, Palaeoclimatology, Palaeoecology, 237:225239.CrossRefGoogle 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:e102.Google Scholar
Erba, E., Bottini, C., Weissert, H. J., and Keller, C. E. 2010. Calcareous nannoplankton response to surface-water acidification around Oceanic Anoxic Event la. Science, 329:428432.Google Scholar
Falkowski, P. G., Katz, M. E., Knoll, A. H., Quigg, A., Raven, J. A., Schofield, O., and Taylor, F. J. R. 2004. The evolution of modern eukaryotic phytoplankton. Science, 305:354360.CrossRefGoogle ScholarPubMed
Fennel, K., Follows, M., and Falkowski, P.G. 2005. The coevolution of the nitrogen, carbon and oxygen cycles in the Proterozoic ocean. American Journal of Science, 305:526545.Google Scholar
Fine, M., and Tchernov, D. 2007. Scleractinian coral species survive and recover from decalcification. Science, 315:1811.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 109:68296834.CrossRefGoogle ScholarPubMed
Foote, M. 2000. Origination and extinction components of taxonomic diversity: general problems. Paleobiology, 26:74102.Google Scholar
Gehling, J. G., and Rigby, J. K. 1996. Long expected sponges from the Neoproterozoic Ediacara fauna of South Australia. Journal of Paleontology, 70:185195.Google Scholar
Gharaie, M. H. M., Matsumoto, R., Racki, G., and Kakuwa, Y. 2007. Chemostratigraphy of Frasnian–Famennian transition: possibility of methane hydrate dissociation leading to mass extinction. Geological Society of America Special Paper, 424:109125.Google Scholar
Gillooly, J. F., Brown, J. H., West, G. B., Savage, V. M., and Charnov, E. L. 2001. Effects of size and temperature on metabolic rate. Science, 293:22482251.Google Scholar
Glazier, D. S. 2005. Beyond the ‘3/4-power law’: variation in the intra–and interspecific scaling of metabolic rate in animals. Biological Reviews, 80:611662.Google Scholar
Grasby, S. E., Sanei, H., and Beauchamp, B. 2011. Catastrophic dispersion of coal fly ash into oceans during the latest Permian extinction. Nature Geosciences, 4:104107.Google Scholar
Greene, S. E., Martindale, R. C., Ritterbush, K. A., Bottjer, D. J., Corsetti, F. A., and Berelson, W. M. 2012. Recognising ocean acidification in deep time: an evaluation of the evidence for acidification across the Triassic–Jurassic boundary. Earth-Science Reviews, 113:7293.Google Scholar
Grice, K., Cao, C., Love, G. D., Böttcher, M. E., Twitchett, R. J., Grosjean, E., Summons, R. E., Turgeon, S. C., Dunning, W., and Jin, Y. 2005. Photic zone euxinia during the Permian–Triassic superanoxic event. Science, 307:706709.Google Scholar
Harper, H. E. J., and Knoll, A. H. 1975. Silica, diatoms, and Cenozoic radiolarian evolution. Geology, 3:175177.Google Scholar
Hautmann, M. 2006. Shell mineralogical trends in epifaunal Mesozoic bivalves and their relationship to seawater chemistry and atmospheric carbon dioxide concentration. Facies, 52:417433.CrossRefGoogle Scholar
Hendy, A. J. W. 2009. The influence of lithification on Cenozoic marine biodiversity trends. Paleobiology, 35:5162.Google Scholar
Hesselbo, S. P., Gröcke, D. R., Jenkyns, H. C., Bjerrum, C. J., Farrimond, P., Morgans Bell, H. S., and Green, O. R. 2000. Massive dissociation of gas hydrate during a Jurassic oceanic anoxic event. Nature, 406:392395.Google Scholar
Hesselbo, S. P., Robinson, S.A., Surlyk, F., and Piasecki, S. 2002. Terrestrial and marine extinction at the Triassic–Jurassic boundary synchronized with major carbon cycle perturbation: a link to initiation of massive volcanism. Geology, 30:251254.Google Scholar
Higgins, J. A., Fischer, W. W., and Schrag, D. P. 2009. Oxygenation of the ocean and sediments: consequences for the seafloor carbonate factory. Earth and Planetary Science Letters, 284:2533.Google Scholar
Higgins, J. A., and Schrag, D. P. 2006. Beyond methane: towards a theory for the Paleocene–Eocene Thermal Maximum. Earth and Planetary Science Letters, 245:523537.Google Scholar
Hinojosa, J. L., Brown, S. T., Chen, J., DePaolo, D. J., Paytan, A., Shen, S.-Z., and Payne, J. L. 2012. Evidence for end-Permian ocean acidification from calcium isotopes in biogenic apatite. Geology, 40:743746.Google Scholar
Hoegh-Guldberg, O., Mumby, P. J., Hooten, A. J., Steneck, R. S., Greenfield, P., Gomez, E., Harvell, C. D., Sale, P. F., Edwards, A. J., Caldeira, K., Knowlton, N., Eakin, C. M., Iglesias-Prieto, R., Muthiga, N., Bradbury, R. H., Dubi, A., Hatziolos, M. E. 2007. Coral reefs under rapid climate change and ocean acidification. Science, 318:17371742.Google Scholar
Hong, J., Cho, S.-H., Choh, S.-J., Woo, J., and Lee, D.-J. 2012. Middle Cambrian siliceous spongecalcimicrobe buildups (Daegi Formation, Korea): metazoan buildup constituents in the aftermath of the Early Cambrian extinction event. Sedimentary Geology, 253:4757.Google Scholar
Horita, J., Zimmermann, H., and Holland, H. D. 2002. Chemical evolution of seawater during the Phanerozoic: Implications from the record of marine evaporites. Geochimica et Cosmochimica Acta, 66:37333756.Google Scholar
Joachimski, M. M., and Buggisch, W. 1993. Anoxic events in the late Frasnian—causes of the Frasnian–Famennian faunal crisis? Geology, 21:675678.Google 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.Google Scholar
Kasemann, S. A., Prave, A. R., Fallick, A. E., Hawkesworth, C. J., and Hoffman, K. H. 2010. Neoproterozoic ice ages, boron isotopes, and ocean acidification: implications for a snowball Earth. Geology, 38:775778.Google Scholar
Katz, M. E., Pak, D. K., Dickens, G. R., and Miller, K. G. 1999. The source and fate of massive carbon input during the latest Paleocene Thermal Maximum. Science, 286:15311533.Google Scholar
Kidder, D. L., and Erwin, D. H. 2001. Secular distribution of biogenic silica through the Phanerozoic: comparison of silica-replaced fossils and bedded cherts at the series level. Journal of Geology, 109:509522.Google Scholar
Kiessling, W., and Simpson, C. 2010. On the potential for ocean acidification to be a general cause of ancient reef crises. Global Change Biology, 2010:112.Google Scholar
Knoll, A. H. 2003a. Life on a Young Planet: the First Three Billion Years of Evolution on Earth. Princeton University Press, Princeton.Google Scholar
Knoll, A. H. 2003b. Biomineralization and evolutionary history. Reviews in Mineralogy and Geochemistry, 54:329356.Google Scholar
Knoll, A. H. 2013. Systems paleobiology. Geological Society of America Bulletin, 125:313.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
Knoll, A. H., and Fischer, W. W. 2011. Skeletons and ocean chemistry: the long view, p. 6782. In Gattuso, J. P. and Hansson, L. (eds.), Ocean Acidification. Oxford University Press, Oxford.Google Scholar
Kroecker, K. J., Kordas, R. L., Crim, R. N., and Singh, G. G. 2010. Meta-analysis reveals negative yet variable effects of ocean acidification on marine organisms. Ecology Letters, 13:14191434.Google Scholar
Kroecker, K. J., Micheli, F., Gambi, M. C., and Martz, T. R. 2011. Divergent ecosystem responses within a benthic marine community to ocean acidification. Proceedings of the National Academy of Sciences of the United States of America, 108:1451514520.Google Scholar
Leinfelder, R. R., Werner, W., Nose, M., Schmid, D. U., Krautter, M., Laternser, R., Takacs, M., and Hartmann, D. 1996. Paleoecology, growth parameters and dynamics of coral, sponge and microbolite reefs from the Late Jurassic, p. 227248. In Reitner, J., Neuweiler, F., and Gunkel, F. (eds.), Global and Regional Controls on Biogenic Sedimentation. I. Reef Evolution. Göttinger Arbeiten zur Geologie und Paläontologie, Sb2, Göttingen.Google Scholar
Maldonado, M., Carmona, M. C., Uriz, M. J., and Cruzado, A. 1999. Decline in Mesozoic reef-building sponges explained by silicon limitation. Nature, 401:785788.Google Scholar
Maliva, R. G., Knoll, A. H., and Siever, R. 1989. Secular change in chert distribution: a reflection of evolving biological participation in the silica cycle. PALAIOS, 4:519532.Google Scholar
Martin, R. E. 2003. The fossil record of biodiversity: nutrients, productivity, habitat area and differential preservation. Lethaia, 36:179193.Google Scholar
Marx, F. G., and Uhen, M. D. 2010. Climate, critters, and cetaceans: Cenozoic drivers of the evolution of modern whales. Science, 327:993–96.Google Scholar
Marynowski, , Zatoń, L. 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, 346–347:6686.Google Scholar
Marzoli, A., Bertrand, H., Knight, K. B., Cirilli, S., Buratti, N., Vérati, C., Nomade, S., Renne, Paul R., Youbi, Nasrrddine, Martini, R., Allenbach, K., Neuwerth, R., Rapaille, C., Zaninetti, L., and Bellieni, G. 2004. Synchrony of the Central Atlantic magmatic province and the Triassic–Jurassic boundary climatic and biotic crisis. Geology, 32:973976.Google Scholar
McAlester, A. L. 1970. Animal extinctions, oxygen consumption, and atmospheric history. Journal of Paleontology, 44:405409.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
McInerney, F. A., and Wing, S. L. 2011. The Paleocene–Eocene Thermal Maximum: a perturbation of carbon cycle, climate, and biosphere with implications for the future. Annual Review of Earth and Planetary Science, 39:489516.Google Scholar
Müller, M., Mentel, M., Van Hellemond, J. J., Henze, K., Woehle, C., Gould, S. B., Yu, R.-Y., Van Der Giezen, M., Tielens, A. G. M., and Martin, W. F. 2012. Biochemistry and evolution of anaerobic energy metabolism in eukaryotes. Microbiology and Molecular Biology Reviews, 76:444495.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
Orr, J. C., Fabry, V. J., Aumont, O., Bopp, L., Doney, S. C., Feely, R. A., Gnanadesikan, A., Gruber, N., Ishida, A., Joos, F., Key, R. M., Lindsay, K., Maier-Reimer, E., Matear, R., Monfray, P., Mouchet, A., Najjar, R. G., Plattner, G.-K., Rodgers, K. B., Sabine, C. L., Sarmiento, J. L., Schlitzer, R., Slater, R. D., Totterdell, I. J., Weirig, M.-F., Yamanaka, Y., and Yool, A. 2005. Anthropogenic ocean acidification over the twenty-first century and its impact on calcifying organisms. Nature, 437:681686.Google Scholar
Pálfy, J., and Smith, P. L. 2000. Synchrony between Early Jurassic extinction, oceanic anoxic event, and the Karoo-Ferrar flood basalt volcanism. Geology, 28:747750.Google Scholar
Payne, J. L., Boyer, A. G., Brown, J. H., Finnegan, S., Kowalewski, M., Krause, R. A. Jr., Lyons, S. K., McClain, C. R., McShea, D. W., Novack-Gottshall, P. M., Smith, F. A., Stempien, J. A., and Wang, S. C. 2009. Two-phase increase in the maximum size of life over 3.5 billion years reflects biological innovation and environmental opportunity. Proceedings of the National Academy of Sciences, 106:2427.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 Finnegan, S. 2006. Controls on marine animal biomass through geological time. Geobiology, 4:110.Google Scholar
Payne, J. L., Lehrmann, D. J., Follett, D., Seibel, M., Kump, L. R., Riccardi, A., Altiner, D., Sano, H., and Wei, J. 2007. Erosional truncation of uppermost Permian shallow-marine carbonates and implications for Permian–Triassic boundary events. Geological Society of America Bulletin, 119:771784.Google Scholar
Payne, J. L., Summers, M., Rego, B. L., Altiner, D., Wei, J., Yu, M., 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
Payne, J. L., Turchyn, A. V., Paytan, A., Depaolo, D. P., Lehrmann, D. J., Yu, Y., and Wei, J. 2010. Calcium isotope constraints on the end-Permian mass extinction. Proceedings of the National Academy of Sciences of the United States of America, 107:85438548.Google Scholar
Pearson, P. N., McMillan, I. K., Wade, B. S., Jones, T. D., Coxall, H. K., Bown, P. R., and Lear, C. H. 2008. Extinction and environmental change across the Eocene–Oligocene boundary in Tanzania. Geology, 36:179182.Google Scholar
Peters, S. E. 2008. Environmental determinants of extinction selectivity in the fossil record. Nature, 454:626629.Google Scholar
Peters, S. E. 2013. Sepkoski's Online Genus Database. Retrieved January 1, 2013. http://strata.geology.wisc.edu/jack/.Google Scholar
Plotnick, R. E., Dornbos, S. Q., and Chen, J. 2010. Information landscapes and sensory ecology of the Cambrian radiation. Paleobiology, 36:303317.Google Scholar
Porter, S. M. 2007. Seawater chemistry and early carbonate biomineralization. Science, 316:1302.Google Scholar
Porter, S. M. 2010. Calcite and aragonite seas and the de novo acquisition of carbonate skeletons. Geobiology, 8:256277.Google Scholar
Pörtner, H. O., Langenbuch, M., and Michaelidis, B. 2005. Synergistic effects of temperature extremes, hypoxia, and increases in CO2 on marine animals: from Earth history to global change. Journal of Geophysical Research, 110:C09S10.Google Scholar
Pörtner, H. O., Langenbuch, M., and Reipschläger, A. 2004. Biological impact of elevated ocean CO2 concentrations: lessons from animal physiology and Earth history. Journal of Oceanography, 60:705718.Google Scholar
Prasad, V., Strömberg, C. A. E., Alimohammadian, H., Sahni, A. 2005. Dinosaur coprolites and the early evolution of grasses and grazers. Science, 310:11771180.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. B., Finnegan, S., Fischer, W. W., Knoll, A. H. 2010. Carbonates in skeleton-poor seas: new insights from Cambrian and Ordovician strata of Laurentia. PALAIOS, 25:7384.Google Scholar
Quigg, A., Finkel, Z. V., Irwin, A. J., Rosenthal, Y., Ho, T. Y., Reinfelder, J. R., Schofield, O., Morel, F. M. M., and Falkowski, P. G. 2003. The evolutionary inheritance of elemental stoichiometry in marine phytoplankton. Nature, 425:291294.Google Scholar
Racki, G. 1999. Silica-secreting biota and mass extinctions: survival patterns and processes. Palaeogeography, Palaeoclimatology, Palaeoecology, 154:107132.Google Scholar
Ratti, S., Knoll, A. H., and Giordano, M. 2011. Did sulfate availability facilitate the evolutionary expansion of chlorophyll a+c phytoplankton in the oceans? Geobiology, 9:301312.Google Scholar
Ries, J. B., Stanley, S. M., and Hardie, L. A. 2006. Scleractinian corals produce calcite, and grow more slowly, in artificial Cretaceous seawater. Geology, 34:525528.Google Scholar
Roopnarine, P. 2006. Extinction cascades and catastrophe in ancient food webs. Paleobiology, 32:119.Google Scholar
Rowley, D. B. 2002. Rate of plate creation and destruction: 180 Ma to present. Geological Society of America Bulletin, 114:927933.Google Scholar
Schulte, P., Alegret, L., Arenillas, I., Arz, J. A., Barton, P. J., Bown, P. R., and Bralower, T. J. 2010. The Chicxulub asteroid impact and mass extinction at the Cretaceous–Paleogene boundary. Science, 327:12141218.Google Scholar
Sepkoski, J. J. Jr. 2002. A compendium of fossil marine animal genera. Bulletins of American Paleontology, 363:1560.Google Scholar
Sessa, J. A., Patzkowsky, M. E., and Bralower, T. J. 2009. The impact of lithification on the diversity, size distribution, and recovery dynamics of marine invertebrate assemblages. Geology, 37:115118.Google Scholar
Sheehan, P. M. 2001. The Late Ordovician mass extinction. Annual Review of Earth and Planetary Science, 29:331364.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, 477:312316.Google Scholar
Stanley, S. M., and Hardie, L. A. 1998. Secular oscillations in the carbonate mineralogy of reef-building and sediment-producing organisms driven by technically forced shifts in seawater chemistry. Palaeogeography, Palaeoclimatology, Palaeoecology, 144:319.Google Scholar
Stanley, S. M., and Powell, M. G. 2003. Depressed rates of origination and extinction during the late Paleozoic ice age: a new state for the global marine ecosystem. Geology, 31:877880.Google Scholar
Stanley, S. M., Ries, J. B., and Hardie, L. A. 2005. Seawater chemistry, coccolithophore population growth, and the origin of Cretaceous chalk. Geology, 33:593596.Google Scholar
Sterner, R. W., Elser, J. J., and Vitousek, P. 2002. Ecological Stoichiometry: the Biology of Elements from Molecules to the Biosphere. Princeton University Press, Princeton.Google Scholar
Streel, M., Caputo, M. V., Loboziak, S., and Melo, J. H. G. 2000. Late Frasnian–Famennian climates based on palynomorph analyses and the question of the Late Devonian glaciations. Earth-Science Reviews, 52:121173.Google Scholar
Svensen, H., Planke, S., Malthe-Sørenssen, A., Jamtveit, B., Myklebust, R., Eidem, T. R., and Rey, S. S. 2004. Release of methane from a volcanic basin as a mechanism for initial Eocene global warming. Nature, 429:542545.Google Scholar
Svensen, H., Planke, S., Polozov, A. G., Schimdbauer, N., Corfu, F., Podladchikov, Y. Y., and Jamtveit, B. 2009. Siberian gas venting and the end-Permian environmental crisis. Earth and Planetary Sciences Letters 277, 490500.Google Scholar
Talmage, S. C., and Gobler, C. J. 2010. Effects of past, present, and future ocean carbon dioxide concentrations on the growth and survival of larval shellfish. Proceedings of the National Academy of Sciences of the United States of America 107:1724617251.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
Valentine, J. W. 1973. Evolutionary Paleoecology of the Marine Biosphere. Prentice-Hall, Englewood Cliffs, New Jersey.Google Scholar
Vermeij, G. J. 1995. Economics, volcanoes, and Phanerozoic revolutions. Paleobiology, 21:125152.Google Scholar
Vermeij, G. J. 2004. Nature: An Economic History. Princeton University Press, Princeton.Google Scholar
Villéger, 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
Ward, P. 2006. Out of Thin Air: Dinosaurs, Birds, and Earth's Ancient Atmosphere. Joseph Henry Press, Washington, D.C. Google Scholar
Wignall, P. B., and Twitchett, R. J. 1996. Oceanic anoxia and the end Permian mass extinction. Science, 272:11551158.Google Scholar
Wilson, J. P. 2103. Modeling 400 million years of plant hydraulics, p. 175194. 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
Wilson, J. P., and Knoll, A. H. 2010. A physiologically explicit morphospace for tracheidbased water transport in modern and extinct seed plants. Paleobiology, 36:335355.Google Scholar
Wood, R., and Zhuravlev, A. Y. 2012. Escalation and ecological selectively of mineralogy in the Cambrian Radiation of skeletons. Earth-Science Reviews, 115:249261.Google Scholar
Zachos, J. C., Röhl, U., Schellenberg, S. A., Sluijs, A., Hodell, D. A., Kelly, D. C., Thomas, E., Nicolo, M., Raffi, I., Lourens, L. J., McCarren, H., and Kroon, D. 2005. Rapid acidification of the ocean during the Paleocene–Eocene Thermal Maximum. Science, 308:16111615.Google Scholar
Zambito, J. J. Iv, Brett, C. E., and Baird, G. C. 2012. The late Middle Devonian (Givetian) global Taghanic biocrisis in its type area (northern Appalachian Basin): geologically rapid faunal transitions driven by global and local environmental changes, p. 677703. In Talent, J. A. (ed.), Earth and Life. Springer, Dordrecht.Google Scholar
Zhang, X-G., and Pratt, B. R. 1994. New and extraordinary Early Cambrian sponge spicule assemblage from China. Geology, 22:4346.Google Scholar
Zhuravlev, A. Y., and Wood, R. A. 1996. Anoxia as the cause of the mid-Early Cambrian (Botomian) extinction event. Geology, 24:311314.Google Scholar
Zhuravlev, A. Y., and Wood, R. A. 2008. Eve of biomineralization: controls on skeletal mineralogy. Geology, 36:923926.Google Scholar