Hostname: page-component-76fb5796d-r6qrq Total loading time: 0 Render date: 2024-04-27T18:54:24.475Z Has data issue: false hasContentIssue false

Altered Primary Production During Mass-Extinction Events

Published online by Cambridge University Press:  21 July 2017

Bas Van De Schootbrugge
Affiliation:
Institute of Earth Sciences, Utrecht University, Budapestlaan 4, 3584 CS Utrecht, The Netherlands
Sabine Gollner
Affiliation:
Senckenberg am Meer, German Center for Marine Biodiversity Research, Südstrand 44, 26382 Wilhelmshaven, Germany
Get access

Abstract

The Big Five mass-extinction events are characterized by dramatic changes in primary producers. Initial disturbance to primary producers is usually followed by a succession of pioneers that represent qualitative and quantitative changes in standing crops of land plants and/or phytoplankton. On land, a transient collapse of arborescent (tree-bearing) vegetation and the rapid spread of a pioneer vegetation dominated by ferns and fern allies characterizes the Permian/Triassic (P/T), Triassic/Jurassic (T/J), and Cretaceous/Paleogene (K/Pg) mass-extinction events. The availability of low-quality food, such as herbaceous low-growing plants, likely played a role in triggering secondary extinctions of herbivores (reptiles, insects). Furthermore, malformation of acritarchs, pollen, and spores during the end-Ordovician, end-Devonian, P/T and T/J extinctions also suggests primary producers were of lesser quality. More importantly, changes in vegetation drove important increases in weathering and erosion leading to elevated nutrient transfer from the continents to the oceans. In the marine realm, the end-Ordovician, end-Devonian, end-Permian, and end-Triassic extinction events are all followed by periods of high primary production, which is reflected in the widespread deposition of black shales. Due to their small size, low nutritional quality, and possible toxicity, the abundance of picoplankton, such as prasinophytes, acritarchs, as well as bacterioplankton (cyanobacteria and green sulfur bacteria) may have been additional factors in delaying ecosystem recovery.

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

Abramovich, S., and Keller, G. 2003. Planktonic foraminiferal response to the latest Maastrichtian abrupt warm event: a case study from South Atlantic DSDP site 525A. Marine Micropaleonotology, 48:225249.CrossRefGoogle 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 Science, 109:728732.CrossRefGoogle Scholar
Algeo, T. J., Chen, Z. Q., Fraiser, M. L., and Twitchett, R. J. 2011. Terrestrial-marine teleconnections in the collapse and rebuilding of Early Triassic marine ecosystems. Palaeogeography, Palaeoclimatology, Palaeoecology, 308:111.CrossRefGoogle Scholar
Algeo, T. J., Hinnov, L., Moser, J., Maynard, J. B., Elswick, E., Kuwahara, K., and Sano, H. 2010. Changes in productivity and redox conditions in the Panthalassic Ocean during the latest Permian. Geology, 38:187190.CrossRefGoogle 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 of London Series B, 353:113130.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.CrossRefGoogle Scholar
Arouri, K., Greenwood, P. F., and Walter, M. R. 1999. A possible Chlorophycean affinity of some Neoproterozoic acritarchs. Organic Geochemistry, 30:13231337.CrossRefGoogle Scholar
Bambach, R. K. 2006. Phanerozoic biodiversity mass extinctions. Annual Review of Earth and Planetary Sciences, 34:127155.CrossRefGoogle Scholar
Bateman, R. M., Crane, P. R., DiMichele, W. A., Kenrick, P. R., Rowe, N. P., Speck, T., and Stein, W. E. 1998. Early evolution of land plants: phylogeny, physiology, and ecology of the primary terrestrial radiation. Annual Review of Ecology and Systematics, 29:263292.CrossRefGoogle Scholar
Baud, A., Richoz, S., and Marcoux, J. 2005. Calcimicrobial cap rocks from the basal Triassic units: western Taurus occurrences (SW Turkey). Comptes Rendus Palevol, 4:569582.CrossRefGoogle 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.CrossRefGoogle Scholar
Becker, C. 2004. Living with constraints—food quality effects on zooplankton. Unpublished PhD dissertation, Kiel, Christian-Albrechts-Universität, 113 p.Google Scholar
Behrenfeld, M. J., and Falkowski, P. G. 1997. Photosynthetic rates derived from satellite-based chlorophyll concentration. Limnology and Oceanography, 42:120.CrossRefGoogle Scholar
Benton, M. J. 1986. More than one event in the Late Triassic mass extinction. Nature, 321:857861.CrossRefGoogle 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.CrossRefGoogle ScholarPubMed
Bond, D. P. G., 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.CrossRefGoogle Scholar
Bonis, N. R., Kürschner, W. M., and Krystyn, L. 2009. A detailed palynological study of the Triassic–Jurassic transition in key sections of the Eiberg Basin (Northern Calcareous Alps, Austria). Review of Palaeobotany and Palynology, 156:376400.CrossRefGoogle Scholar
Bonis, N. R., Ruhl, M., and Kürschner, W. M. 2010. Climate change driven black shale deposition during the end-Triassic in the western Tethys. Palaeogeography, Palaeoclimatology, Palaeoecology, 290:151159.CrossRefGoogle Scholar
Bown, P. R. 2005. Selective calcareous nannoplankton survivorship at the Cretaceous–Tertiary boundary. Geology, 33:653656.CrossRefGoogle Scholar
Boyce, C. K., and Lee, J. E. 2011. Could land plant evolution have fed the marine revolution? Paleontological Research, 15:100105.CrossRefGoogle Scholar
Boyce, D. G., Lewis, M. R., and Worm, B. 2010. Global phytoplankton decline over the past century. Nature, 466:591596.CrossRefGoogle ScholarPubMed
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.CrossRefGoogle Scholar
Brinkhuis, H., Bujak, J. P., Smit, J., Versteegh, G. J. M., and Visscher, H. 1998. Dinoflagellate-based sea surface temperature reconstructions across the Cretaceous–Tertiary boundary. Palaeogeography, Palaeoclimatology, Palaeoecology, 141:6783.CrossRefGoogle Scholar
Brinkhuis, H., and Schioler, P. 1996. Palynology of the Geulhemmerberg Cretaceous/Tertiary boundary section (Limburg, SE Netherlands). Geologie en Mijnbouw, 75:193213.Google Scholar
Brown, M. R., Jeffrey, S. W., Volkman, J. K., and Dunstan, G. A. 1997. Nutritional properties of microalgae for mariculture. Aquaculture, 151:315331.CrossRefGoogle Scholar
Bukovinszky, T., van Veen, F. J. F., Jongema, Y., and Dicke, M. 2008. Direct and indirect effects of resource quality on food web structure. Science, 319:804807.CrossRefGoogle ScholarPubMed
Cao, C. Q., Love, G. D., Hays, L. E., Wang, W., and Shen, S. Z. 2009. Biogeochemical evidence for euxinic oceans and ecological disturbance presaging the end-Permian mass-extinction event. Earth and Planetary Science Letters, 281:188201.CrossRefGoogle Scholar
Castle, J. W., and Rodgers, J. H. 2009. Hypothesis for the role of toxin-producing algae in Phanerozoic mass extinctions based on evidence from the geologic record and modern environments. Environmental Geoscience, 16:123.CrossRefGoogle Scholar
Chen, D., Tucker, M. E., Shen, Y., Yans, J., and Preat, 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.CrossRefGoogle Scholar
Clemence, M.-E., Gardin, S., Bartolini, A., Paris, G., Beaumont, V., and Guex, J. 2010. Benthoplanktonic evidence from the Austrian Alps for a decline in sea-surface carbonate production at the end of the Triassic. Swiss Journal of Geoscience, 103:293315.CrossRefGoogle Scholar
Colbath, G. K. 1986. Abrupt terminal Ordovician extinction in phytoplankton associations, southern Appalachians. Geology, 14:943946.2.0.CO;2>CrossRefGoogle Scholar
Colbath, G. K., and Grenfell, H. R. 1995. Review of biological affinities of Paleozoic acid-resistant, organic-walled eukaryotic algal microfossils (including “acritarchs”). Review of Palaeobotany and Palynology, 86:287314.CrossRefGoogle Scholar
D'Hondt, S., Donaghay, P., Zachos, J. C., Luttenberg, D., and Lindlinger, M. 1998. Organic carbon fluxes and ecological recovery from the Cretaceous–Tertiary mass extinction. Science, 282:276279.CrossRefGoogle ScholarPubMed
de la Rue, S. R., Rowe, H. D., and Rimmer, S. M. 2007. Palynological and bulk geochemical constraints on the paleoceanographic conditions across the Frasnian–Famennian boundary, New Albany Shale, Indiana: International Journal of Coal Geology, 71:7284.CrossRefGoogle Scholar
Delabroye, A., Munnecke, A., Servais, T., Vandenbroucke, T. R. A., and Vecoli, M. 2012. Abnormal forms of acritarchs (phytoplankton) in the upper Hirnantian (Upper Ordovician) of Anticosti Island, Canada. Review of Palaeobotany and Palynology, 173:4656.CrossRefGoogle Scholar
Delabroye, A., Munnecke, A., Vecoli, M., Copper, P., Tribovillard, N., Joachimski, M. M., Desrochers, A., and Servais, T. 2011. Phytoplankton dynamics across the Ordovician/Silurian boundary at low palaeolatitudes: Correlations with the carbon isotopic and glacial events. Palaeogeography, Palaeoclimatology, Palaeoecology, 312:7997.CrossRefGoogle Scholar
Erwin, D. H. 2001. Lessons from the past: biotic recoveries from mass extinctions. Proceedings of the National Academy of Science, 98:53995403.CrossRefGoogle ScholarPubMed
Eshet, Y., Rampino, M. R., and Visscher, H. 1995. Fungal event and palynological record of ecological crisis and recovery across the Permian–Triassic boundary. Geology, 23:967970.2.3.CO;2>CrossRefGoogle Scholar
Falkowski, P. G. 2012. The power of the plankton. Nature, 483:1720.CrossRefGoogle ScholarPubMed
Falkowski, P. G., Barber, R. T., and Smetacek, V. 1998. Biogeochemical controls and feedbacks on ocean primary production. Science, 281:200206.CrossRefGoogle ScholarPubMed
Falkowski, P. G., and Raven, J. A. 2007. Aquatic Photosynthesis. Princeton University Press, Princeton.CrossRefGoogle Scholar
Feist-Burkhardt, S., Götz, A. E., Ruckwied, K., and Russell, J. W. 2008. Palynofacies patterns, acritarch diversity and stable isotope signatures in the Lower Muschelkalk (Middle Triassic) of N Switzerland. Evidence of third-order cyclicity. Swiss Journal of Geoscience, 101:115.CrossRefGoogle Scholar
Filipiak, P. 2009. Lower Famennian phytoplankton from the Holy Cross Mountains, Central Poland. Review of Palaeobotany and Palynology. 157:326338.CrossRefGoogle Scholar
Filipiak, P., and Racki, G. 2010. Proliferation of abnormal palynoflora during the end-Devonian biotic crisis. Geological Quarterly, 54:114.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.CrossRefGoogle ScholarPubMed
Finnie, J. W., Windsor, P. A., and Kessell, A. E. 2011. Neurological diseases of ruminant livestock in Australia. II: toxic disorders and nutritional deficiencies. Australian Veterinary Journal, 89:247253.CrossRefGoogle Scholar
Gardin, S., Krystyn, L., Richoz, S., Bartolini, A., and Galbrun, B. 2012. Where and when the earliest coccolithophores? Lethaia, 45:507523.CrossRefGoogle Scholar
Gastaldo, R. A., Neveling, J., Kittinger Clark, C., and Newbury, S. S. 2009. The terrestrial Permian–Triassic boundary event bed is a nonevent. Geology, 37:199202.CrossRefGoogle Scholar
Gonzalez, F. 2009. Reappraisal of the organic-walled microphytoplankton genus Maranhites: Morphology, excystment, and speciation. Review of Palaeobotany and Palynology, 154:621.CrossRefGoogle Scholar
Götz, A. E., and Feist-Burkhardt, S. 2012. Phytoplankton associations of the Anisian Peri-Tethys Basin (Central Europe): Evidence of basin evolution and palaeoenvironmental change. Palaeogeography Palaeoclimatology Palaeoecology, 337–338:151158.CrossRefGoogle Scholar
Götz, A. E., Ruckwied, K., Palfy, J., and Haas, J. 2009. Palynological evidence for synchronous changes within the terrestrial and marine realm at the Triassic–Jurassic boundary (Csövàr section, Hungary). Review of Palaeobotany and Palynology, 156:401409.CrossRefGoogle Scholar
Grauvogel-Stamm, L. and Ash, S. R. 2005. Recovery of the Triassic land flora from the end-Permian life crisis. Comptes Rendus Palevol, 4:593608.CrossRefGoogle Scholar
Grice, K., Twitchett, R. J., Alexander, R., Foster, C. B., and Looy, C. V. 2005. A potential biomarker for the Permian–Triassic ecological crisis. Earth and Planetary Science Letters, 236:315321.CrossRefGoogle Scholar
Habib, D., Moshkovitz, S., and Kramer, C. 1992. Dinoflagellate and calcareous nannofossil response to sea-level change in Cretaceous–Tertiary boundary sections. Geology, 20:6265.2.3.CO;2>CrossRefGoogle Scholar
Habib, D., and Saeedi, F. 2007. The Manumiella seelandica global spike: Cooling during regression at the close of the Maastrichtian. Palaeogeography, Palaeoclimatology, Palaeoecology, 255:8797.CrossRefGoogle Scholar
Hammarlund, E. U., Dahl, T. W., Harper, D. A. T., Bond, D. P. G., Nielsen, A. T., Bjerrum, C. J., Schovsbo, N. H., Schönlaub, H. P., Zalasiewicz, J. A., and Canfield, D. E. 2012. A sulfidic driver for the end-Ordovician mass extinction. Earth and Planetary Science Letters, 331–332:128139.CrossRefGoogle Scholar
Häntzschel, W., and Reineck, H.-E. 1968. Fazies-Untersuchungen im Hettangium von Helmstedt (Niedersachsen). Mitteilungen aus dem Geologischen Staatsinstitut in Hamburg, 37:539.Google Scholar
Harper, D. A. T., Hammarlund, E. U., and Rasmussen, C. M. O. 2013. End Ordovician extinctions: A coincidence of causes. Gondwana Research, in press.CrossRefGoogle Scholar
Higgs, K. T., Dreesen, R., Dusar, M., and Streel, M. 1992. Palynostratigraphy of the Tournaisian (Hastarian) rocks in the Namur Synclinorium, West Flanders, Belgium. Review of Palaeobotany and Palynology, 72:149158.CrossRefGoogle Scholar
Hints, O., Delabroye, A., Nolvak, J., Servais, T., Uutela, A., and Wallin, A. 2010. Biodiversity patterns of Ordovician marine microphytoplankton from Baltica: Comparison with other fossil groups and sea-level changes. Palaeogeography, Palaeoclimatology, Palaeoecology, 294:161173.CrossRefGoogle Scholar
Holser, W. T., and Magaritz, M. 1992. Cretaceous/ Tertiary and Permian/Triassic boundary events compared. Geochimica et Cosmochimica Acta, 56:32973309.CrossRefGoogle 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
Hummel, J., Gee, C. T., Südekam, K.-H., Sander, P. M., Nogge, G., and Clauss, M. 2008. In vitro digestibility of fern and gymnosperm foliage: implications for sauropod feeding ecology and diet selection. Proceedings of the Royal Society of London Series B, 275:10151021.Google ScholarPubMed
Isozaki, Y. 1997. Permo–Triassic boundary superanoxia and stratified superocean: Records from lost deep sea. Science. 276:235238.CrossRefGoogle ScholarPubMed
Irigoien, X., Verheye, H. M., Harris, R. P., and Harbour, D. 2005. Effect of food composition on egg production and hatching success rate of two copepod species (Calanoides carinatus and Rhincalanus nasutus) in the Benguela upwelling system. Journal of Plankton Research, 27:735742.CrossRefGoogle Scholar
Jiang, H. X., Wu, Y., and Cai, C. F. 2008. Filamentous cyanobacteria fossils and their significance in the Permian–Triassic boundary section at Laolongdong, Chongqing. Chinese Science Bulletin, 53:18711879.CrossRefGoogle Scholar
Joachimski, M. M., and Buggisch, W. 1993. Anoxic events in the late Frasnian—causes of the Frasnian–Famennian faunal crisis. Geology, 21:675678.2.3.CO;2>CrossRefGoogle Scholar
Joachimski, M. M., Ostertag-Hennig, C., Pancost, R. D., Strauss, H., Freeman, K. H., Littke, R., Sinninghe Damste, J. S., and Racki, G. 2001. Water column anoxia, enhanced productivity and concomitant changes in δ13C and δ34S across the Frasnian–Famennian boundary (Kowala-Holy Cross Mountains/Poland). Chemical Geology, 175:109131.CrossRefGoogle Scholar
Joachimski, M. M., Pancost, R. D., Freeman, K. H., Ostertag-Hennig, C., and Buggisch, W. 2002. Carbon isotope geochemistry of the Frasnian–Famennian transition: Palaeogeography, Palaeoclimatology, Palaeoecology, 181:91109.CrossRefGoogle Scholar
Johnson, K. R., and Hickey, L. J. 1990. Megafloral change across the Cretaceous–Tertiary boundary in the northern Great Plains and Rocky Mountains, USA, p. 433444. In Sharpton, V. L. and Ward, P. D. (eds.), Global Catastrophes in Earth History: An Interdisciplinary Conference on Impacts, Volcanism, and Mass Mortality. Geological Society of America Special Paper 247, Geological Society of America, Boulder, Colorado.CrossRefGoogle Scholar
Kaiser, S. I., Steuber, T., Becker, R. T., and Joachimski, M. M. 2006. Geochemical evidence for major environmental change at the Devonian–Carboniferous boundary in the Carnic Alps and the Rhenish Massif. Palaeogeography, Palaeoclimatology, Palaeoecology, 240:146160.CrossRefGoogle Scholar
Kazmierczak, J., and Kremer, B. 2009. Spore-like bodies in some early Paleozoic acritarchs: Clues to chlorococcalean affinities. Acta Palaeontologica Polonica, 54:541551.CrossRefGoogle Scholar
Keller, G., Abramovich, S., Berner, Z., and Adatte, T. 2009. Biotic effects of the Chicxulub impact, K–T catastrophe and sea level change in Texas. Palaeogeography, Palaeoclimatology, Palaeoecology, 271:5268.CrossRefGoogle Scholar
Kershaw, S., Zhang, T., and Lan, G. 1999. A? microbialite carbonate crust at the Permian–Triassic boundary in South China, and its palaeoenvironmental significance. Palaeogeography, Palaeoclimatology, Palaeoecology, 146:118.CrossRefGoogle 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.CrossRefGoogle ScholarPubMed
Kiessling, W., and Danelian, T. 2011. Trajectories of Late Permian–Jurassic radiolarian extinction rates: no evidence for an end-Triassic mass extinction. Fossil Record, 14:95101.CrossRefGoogle Scholar
Kump, L. R. 1991. Interpreting carbon-isotope excursions: Strangelove oceans. Geology. 19:299302.2.3.CO;2>CrossRefGoogle Scholar
Kump, L. R., and Arthur, M. A. 1999. Interpreting carbon-isotope excursions: carbonates and organic matter. Chemical Geology, 161:181198.CrossRefGoogle Scholar
Kürschner, W., Bonis, N., and Krysytn, L. 2007. Carbon isotope stratigraphy and palynostratigraphy of the Triassic–Jurassic transition in the Tiefengraben section-Northern Calcareous Alps (Austria): Palaeogeography, Palaeoclimatology, Palaeoecology. 244:257280.CrossRefGoogle Scholar
Labandeira, C. C., Johnson, K. R., and Wilf, P. 2002. Impact of the terminal Cretaceous event on plant–insect associations. Proceedings of the National Academy of Science, 99:20612066.CrossRefGoogle ScholarPubMed
Lenton, T. M., Crouch, M., Johnson, M., Pires, N., and Dolan, L. 2012. First plants cooled the Ordovician. Nature Geoscience, 5:8689.CrossRefGoogle Scholar
Lindström, S., Van De Schootbrugge, B., Dybkjaer, K., Pedersen, G. K., Fiebig, J., Nielsen, L. H., and Richoz, S. 2012. No causal link between terrestrial ecosystem change and methane release during the end-Triassic mass extinction. Geology, 40:531534.CrossRefGoogle Scholar
Looy, C. V., Brugman, W. A., Dilcher, D. L., and Visscher, H. 1999. The delayed resurgence of equatorial forests after the Permian–Triassic ecologic crisis. Proceedings of the National Academy of Science, 96:1385713862.CrossRefGoogle ScholarPubMed
Looy, C. V., Twitchett, R. J., Dilcher, D. L., van Konijnenburg-Van Cittert, J. H. A., and Visscher, H. 2001. Life in the end-Permian dead zone. Proceedings of the National Academy of Science, 98:78797883.CrossRefGoogle ScholarPubMed
Marynowski, L., and Filipiak, P. 2007. Water column euxinia and wildfire evidence during deposition of the Upper Famennian Hangenberg event horizon from the Holy Cross Mountains (central Poland). Geological Magazine, 144:569595.CrossRefGoogle Scholar
Marynowski, L., Filipiak, P., and Zaton, M. 2010. Geochemical and palynological study of the Upper Famennian Dasberg event horizon from the Holy Cross Mountains (central Poland). Geological Magazine, 147:527550.CrossRefGoogle Scholar
Marynowski, L., Narkiewicz, M., and Grelowski, C. 2000. Biomarkers as environmental indicators in a carbonate complex, example from the Middle to Upper Devonian, Holy Cross Mountains, Poland. Sedimentary Geology, 137:187212.CrossRefGoogle Scholar
Marynowski, L., Rakocinski, M., Borcuch, E., Kremer, B., Schubert, B. A., and Jahren, A. H. 2011. Molecular and petrography indicators of redox conditions and bacterial communities after the F/F mass extinction (Kowala, Holy Cross Mountains, Poland). Palaeogeography, Palaeoclimatology, Palaeoecology, 306:114.CrossRefGoogle Scholar
Marynowski, L., Zaton, M., Rakocinski, 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.CrossRefGoogle 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.CrossRefGoogle ScholarPubMed
Maziane, N., Higgs, K. T., and Streel, M. 2002. Biometry and paleoenvironment of Retispora lepidophyta (Kedo) Playford 1976 and associated miospores in the latest Famennian nearshore marine facies, eastern Ardenne (Belgium). Review of Paleobotany and Palynology, 118:211226.CrossRefGoogle Scholar
McElwain, J. C., Beerling, D. J., and Woodward, F. I. 1999. Fossil plants and global warming at the Triassic–Jurassic boundary. Science, 285:13861390.CrossRefGoogle ScholarPubMed
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.CrossRefGoogle Scholar
McElwain, J. C., and Punyasena, S. W. 2007. Mass extinction events and the plant fossil record. Trends in Ecology and Evolution, 22:548557.CrossRefGoogle ScholarPubMed
McGhee, G. R. Jr. 1988. The late Devonian extinction event: evidence for abrupt ecosystem collapse. Paleobiology, 14:250257.CrossRefGoogle 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.CrossRefGoogle 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 Science, 109:1185718861.CrossRefGoogle ScholarPubMed
Nabbefeld, B., Grice, K., Twitchett, R. J., Summons, R. E., Hays, L., Böttcher, M. E., and Asif, M. 2010. An integrated biomarker, isotopic and palaeoenvironmental study through the late Permian event at Lusitaniadalen, Spitsbergen. Earth and Planetary Science Letters, 291:8496.CrossRefGoogle Scholar
Nichols, D. J. 2007. Selected plant microfossil records of the terminal Cretaceous event in terrestrial rocks, western North America. Palaeogeography, Palaeoclimatology, Palaeoecology, 255:2234.CrossRefGoogle Scholar
Nichols, D. J., and Fleming, R. F. 1990. Plant microfossil record of the terminal Cretaceous event in the western United States and Canada, p. 445455. In Sharpton, V. L. and Ward, P. D. (eds.), Global Catastrophes in Earth History: An Interdisciplinary Conference on Impacts, Volcanism, and Mass Mortality. Geological Society of America Special Paper 247, Geological Society of America, Boulder, Colorado.CrossRefGoogle Scholar
Olsen, P. E., Kent, D. V., Sues, H.-D., Koeberl, C., Huber, H., Montanari, A., Rainforth, E. C., Fowell, S. J., Szajina, M. J., and Hartline, B. 2002. Ascent of dinosaurs linked to an iridium anomaly at the Triassic–Jurassic boundary. Science, 296:13051307.CrossRefGoogle Scholar
Palliani, R. B., and Buratti, N. 2006. High diversity dinoflagellate cyst assemblages from the Late Triassic of southern England: new information on early dinoflagellate evolution and palaeogeography. Lethaia, 39:305312.CrossRefGoogle Scholar
Paris, F., Girard, C., Feist, R., and Winchester-Seeto, T. 1996. Chitinozoan bio-event in the Frasnian–Famennian boundary beds at La Serre (Montagne Noire, Southern France). Palaeogeography, Palaeoclimatology, Palaeoecology, 121:131145.CrossRefGoogle 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.CrossRefGoogle ScholarPubMed
Payne, J. L., and van de Schootbrugge, B. 2007. Life in Triassic oceans: Links between planktonic and benthic recovery and radiation, p. 165189. In Falkowski, P. G. and Knoll, A. H. (eds.), Evolution of Primary Producers in the Sea. Elsevier Academic Press, Amsterdam.CrossRefGoogle Scholar
Payne, , Summers, J. L. 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.CrossRefGoogle Scholar
Pospichal, J. J. 1994. Calcareous nannofossils at the K/T boundary, El Kef: No evidence for stepwise, gradual, or sequential extinctions. Geology, 22:99102.2.3.CO;2>CrossRefGoogle Scholar
Pujol, F., Berner, Z., and Stüben, D. 2006. Palaeoenvironmental changes at the Frasnian/ Famennian boundary in key European sections. Chemostratigraphic constraints: Palaeogeography, Palaeoclimatology, Palaeoecology, 240:120145.Google Scholar
Rabalais, N. N., Turner, R. E., Dortch, Q., Justic, D., Bierman, V. J., and Wiseman, W. J. 2002. Nutrient-enhanced productivity in the northern Gulf of Mexico: past, present and future. Hydrobiologia, 475/476:3963.CrossRefGoogle Scholar
Racki, G. 1999. Silica-secreting biota and mass extinctions: survival patterns and processes. Palaeogeography, Palaeoclimatology, Palaeoecology, 154:107132.CrossRefGoogle Scholar
Rampino, M. R., and Caldeira, K. 2005. Major perturbation of ocean chemistry and a “Strangelove Ocean” after the end-Permian mass extinction. Terra Nova, 17:554559.CrossRefGoogle Scholar
Retallack, G. J., Veevers, J. J., and Morante, R. 1996. Global coal gap between Permian–Triassic extinction and Middle Triassic recovery of peat-forming plants: GSA Bulletin, 108:195207.2.3.CO;2>CrossRefGoogle Scholar
Richoz, S., Van De Schootbrugge, B., Pross, J., Püttmann, W., Quan, T. M., Lindström, S., Heunisch, C., Fiebig, J., Maquil, R., Hauzenberger, C., Schouten, S., and Wignall, P. B. 2012. Hydrogen sulphide poisoning of shallow seas due to end-Triassic global warming. Nature Geoscience, 5:662666.CrossRefGoogle Scholar
Riding, J. B., and Ioannides, N. S. 1996. A review of Jurassic dinoflagellate cyst biostratigraphy and global provincialism. Bulletin Societé Géologique de France. 167:314.Google Scholar
Riegel, W. 1993. Die geologische Bedeutung der Prasinophyten im Paläozoikum: Göttinger Arbeiten in Geologie und Paläontologie, 58:3950.Google Scholar
Riegel, W. 2008. The Late Palaeozoic phytoplankton blackout—Artefact or evidence of global changes? Review of Palaeobotany and Palynology, 148:7390.CrossRefGoogle Scholar
Roopnarine, P. D. 2006. Extinction cascades and catastrophe in ancient food webs. Paleobiology, 32:119.CrossRefGoogle Scholar
Rothwell, G. W., Scheckler, S. E., and Gillespie, W. H. 1989. Elkinsia gen. nov., a Late Devonian gymnosperm with cupulate ovules. Botanical Gazette, 150:170189.CrossRefGoogle Scholar
Ruhl, M., Deenen, M. H. L., Abels, H. A., Bonis, N. R., Krijgsman, W., and Kürschner, W. M. 2010. Astronomical constraints on the duration of the early Jurassic Hettangian stage and recovery rates following the end-Triassic mass extinction (St Audrie's Bay/East Quantoxhead, UK). Earth and Planetary Science Letters, 295:262276.CrossRefGoogle Scholar
Ruhl, M., Bonis, N. R., Reichart, G.-J., Sinninghe Damste, J. S., and Kürschner, W. 2011. Atmospheric carbon injection linked to end-Triassic mass extinction. Science, 333:430434.CrossRefGoogle ScholarPubMed
Sahney, S., and Benton, M. J. 2008. Recovery from the most profound mass extinction of all time. Proceedings of the Royal Society of London Series B, 275:759765.Google ScholarPubMed
Sarjeant, W. A. S. 1970. Acritarchs and Tasmanitids from the Chhidru Formation, Uppermost Permian of West Pakistan, p. 277304. In Kummel, B. and Teichert, C. (eds.), Stratigraphic boundary problems: Permian and Triassic of West Pakistan. The University of Kansas Paleontological Contributions Special Publication 4, University of Kansas Press, Lawrence, KS.Google Scholar
Sarjeant, W. A. S. 1973. Acritarchs and Tasmanitids from the Mianwali and Tredian Formations (Triassic) of the Salt and Surghar Ranges, West Pakistan, p. 3573. In Logan, A. and Hilles, V. (eds.), The Permian and Triassic Systems and Their Mutual Boundary, Canadian Society of Petroleum Geologists Memoir 2, Calgary, Alberta.Google Scholar
Schmidt, K., and Jonasdottir, S. H. 1997. Nutritional quality of two cyanobacteria: How rich is “poor food”? Marine Ecology Progress Series, 151:110.CrossRefGoogle 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
Seawright, A. A. 1995. Directly toxic effects of plant chemicals which may occur in humans and animal foods. Natural Toxins, 3:227232.CrossRefGoogle ScholarPubMed
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.CrossRefGoogle 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, 326:127132.CrossRefGoogle ScholarPubMed
Servais, T., Harper, D. A. T., Li, J., Munnecke, A., Owen, A. W., and Sheehan, P. M. 2009. Understanding the Great Ordovician Biodiversification Event (GOBE): influences of paleogeography, paleoclimate, or paleoecology? GSA Today, 19:410.CrossRefGoogle Scholar
Servais, T., Lehnert, O., Li, J., Mullins, G. L., Munnecke, A., Nützel, A., and Vecoli, M. 2008. The Ordovician biodiversification: revolution in the oceanic trophic chain. Lethaia, 41:99109.CrossRefGoogle Scholar
Sheehan, P. M. 2001. The Late Ordovician mass extinction. Annual Review of Earth and Planetary Sciences, 29:331364.CrossRefGoogle Scholar
Sheehan, P. M., Corough, P. J., and Fastovsky, D. E. 1996. Biotic selectivity during the K/ T and Late Ordovician extinction events, p. 477489. In Ryder, G., Fastovsky, D. E., and Gartner, S. (eds.), The Cretaceous–Tertiary Event and Other Catastrophes in Earth History, Geological Society of America Special Paper 307: Geological Society of America, Boulder, Colorado.Google Scholar
Sheehan, P. M., and Harris, M. T. 2004. Microbialite resurgence after the Late Ordovician extinction. Nature, 430:7578.CrossRefGoogle ScholarPubMed
Sogot, C. E., Harper, E. M., and Taylor, P. D. 2013. Biogeographical and ecological patterns in bryozoans across the Cretaceous–Paleogene boundary: Implications for the phytoplankton collapse hypothesis. Geology, 41:631634.CrossRefGoogle Scholar
Somvanshi, R., Lauren, D. R., Smith, B. L., Dawra, R. K., Sharma, O. P., Sharma, V. K., Singh, A. K., and Gangwar, N. K. 2006. Estimation of the fern toxin, ptaquiloside, in certain Indian ferns other than bracken. Current Science, 91:15471552.Google Scholar
Song, H. J., Tong, J. N., Xiong, Y. L., Sun, D. Y., Tian, L., and Song, H. Y. 2012. The large increase of δ13Ccarb-depth gradient and the end-Permian mass extinction. Science China-Earth Sciences, 55:11011109.CrossRefGoogle Scholar
Strother, P. K. 2008. A speculative review of factors controlling the evolution of phytoplankton during Palaeozoic time. Revue de Micropaleontologie, 51:921.CrossRefGoogle Scholar
Svensen, H., Planke, S., Polozov, A. G., Schmidbauer, N., Corfu, F., Podladchikov, Y. Y., and Jamtveit, B. 2009. Siberian gas venting and the end-Permian environmental crisis. Earth and Planetary Science Letters, 277:490500.CrossRefGoogle Scholar
Sweet, A. R., Braman, D. R., and Lerbekmo, J. F. 1990. Palynofloral response to K/T boundary events: a transitory interruption within a dynamic system, p. 457469. In Sharpton, V. L. and Ward, P. D. (eds.), Global Catastrophes in Earth History: An Interdisciplinary Conference on Impacts, Volcanism, and Mass Mortality. Geological Society of America Special Paper 247, Geological Society of America, Boulder, Colorado.CrossRefGoogle Scholar
Tribovillard, N., Averbuch, O., Devleeschouwer, X., Racki, G., and Riboulleau, A. 2004. Deep-water anoxia over the Frasnian–Famennian boundary (La Serre, France): a tectonically induced oceanic anoxic event. Terra Nova, 16:288295.CrossRefGoogle Scholar
Tschudy, R. H., Pillmore, C. L., Orth, C. J., Gilmore, J. S., and Knight, J. D. 1984. Disruption of the terrestrial plant ecosystem at the Cretaceous–Tertiary boundary, western North America. Science, 225:10301032.CrossRefGoogle Scholar
Turner, J. F., Ianora, A., Esposito, F., Carotenuto, Y., and Miralto, A. 2002. Zooplankton feeding ecology: does a diet of Phaeocystis support good copepod grazing, survival, egg production and egg hatching success? Journal of Plankton Research, 24:11851195.CrossRefGoogle Scholar
Vajda, V., and McLoughlin, S. 2006. Extinction and recovery patterns of the vegetation across the Cretaceous–Palaeogene boundary—a tool for unravelling the causes of the end-Permian mass extinction. Review of Palaeobotany and Palynology, 144:99112.CrossRefGoogle Scholar
Vajda, V., and Raine, I. J. 2003. Pollen and spores in marine Cretaceous/Tertiary boundary sediments at mid-Waipara River, North Canterbury, New Zealand. New Zealand Journal of Geology and Geophysics, 46:255273.CrossRefGoogle Scholar
Vajda, V., Raine, I. J., and Hollis, C. J. 2001. Indication of global deforestation at the Cretaceous–Tertiary boundary by New Zealand fern spike. Science, 294:17001702.CrossRefGoogle ScholarPubMed
van de Schootbrugge, B., Payne, J. L., Tomasovych, A., Pross, J., Fiebig, J., Benbrahim, M., Föllmi, 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.CrossRefGoogle Scholar
van de Schootbrugge, B., Quan, T. M., Lindström, S., Püttmann, W., Heunisch, C., Pross, J., Fiebig, J., Petschick, R., Röhling, 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, 2:489594.CrossRefGoogle Scholar
van de Schootbrugge, B., Tremolada, F., Bailey, T. R., Rosenthal, Y., 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.CrossRefGoogle Scholar
Visscher, H., Brinkhuis, H., Dilcher, D. L., Elsik, W. C., Eshet, Y., Looy, C. V., Rampino, M. R., and Traverse, A. 1996. The terminal Paleozoic fungal event: evidence of terrestrial ecosystem destabilization and collapse. Proceedings of the National Academy of Science, 93:21552158.CrossRefGoogle ScholarPubMed
Visscher, H., Looy, C. V., Collinson, M. E., Brinkhuis, H., van Konijnenburg-Cittert, J. H. A., Kürschner, W. M., and Sephton, M. A. 2004. Environmental mutagenesis during the end-Permian ecological crisis. Proceedings of the National Academy of Science, 101:1295212956.CrossRefGoogle ScholarPubMed
Visscher, H., Sephton, M. A., and Looy, C. V. 2011. Fungal virulence at the time of the end-Permian biosphere crisis? Geology, 39:883886.CrossRefGoogle Scholar
Ward, P. D., Haggart, J. W., Carter, E. S., Wilbur, D., Tipper, H. W., and Evans, T. 2001. Sudden productivity collapse associated with the Triassic–Jurassic boundary mass extinction. Science, 292:11481151.CrossRefGoogle ScholarPubMed
Ward, P. D., Montgomery, D. R., and Smith, R. 2000. Altered river morphology in South Africa related to the Permian–Triassic extinction. Science, 289:17401743.CrossRefGoogle Scholar
Weiss, M. B., Curran, P. B., Peterson, B. J., and Gobler, C. J. 2007. The influence of plankton composition and water quality on hard clam (Mercenaria mercenaria) populations across Long Island's south shore lagoon estuaries (New York, USA). Journal of Experimental Marine Biology and Ecology, 345:1225.CrossRefGoogle Scholar
Whalen, M. T., Day, J., Eberli, G. P., and Homewood, P. W. 2002. Microbial carbonates as indicators of environmental change and biotic crises in carbonate systems. Examples from the Late Devonian, Alberta basin, Canada. Palaeogeography, Palaeoclimatology, Palaeoecology, 181:127151.CrossRefGoogle Scholar
Wiggins, V. D. 1973. Upper Triassic dinoflagellates from arctic Alaska. Micropaleontology, 19:116.CrossRefGoogle Scholar
Wignall, P. B., and Bond, D. P. G. 2008. The end-Triassic and Early Jurassic mass extinction records in the British Isles. Proceedings of the Geologist's Association, 119:7384.CrossRefGoogle Scholar
Wignall, P. B., Bond, D. P. G., Kuwahara, K., Kakuwa, Y., Newton, R. J., and Poulton, S. W. 2010. An 80 million year oceanic redox history from the Permian to Jurassic pelagic sediments of the Mino-Tamba terrane, SW Japan, and the origin of four mass extinctions. Global and Planetary Change, 71:109123.CrossRefGoogle Scholar
Wignall, P. B., and Hallam, A. 1992. Anoxia as a cause of the Permian/Triassic mass extinction: facies evidence from northern Italy and the western United States. Palaeogeography, Palaeoclimatology, Palaeoecology, 93:2146.CrossRefGoogle Scholar
Wignall, P. B., Morante, R., and Newton, R. J. 1998. The Permo–Triassic transition in Spitsbergen: δ13C(org) chemostratigraphy, Fe and S geochemistry, facies, fauna and trace fossils: Geological Magazine, 135:4762.CrossRefGoogle Scholar
Wignall, P. B., and Twitchett, R. J. 2002. Extent, duration, and nature of the Permian–Triassic superanoxic event, p. 395413. In MacLeod, K. G. (ed.), Catastrophic Events and Mass Extinctions; Impacts and Beyond, Geological Society of America Special Papers 356. Geological Society of America, Boulder, Colorado.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.CrossRefGoogle Scholar
Williford, K. H., Ward, P. D., Garrison, G. H., and Buick, R. 2007. An extended organic carbon-isotope record across the Triassic–Jurassic boundary in the Queen Charlotte Islands, British Columbia, Canada. Palaeogeography, Palaeoclimatology, Palaeoecology, 244:290296.CrossRefGoogle Scholar
Xie, S. C., Pancost, R. D., Yin, H. F., Wang, H. M., and Evershed, R. P. 2005. Two episodes of microbial change coupled with Permo/Triassic faunal mass extinction. Nature, 434:494497.CrossRefGoogle ScholarPubMed
Zachos, J. C., Arthur, M. A., and Dean, W. A. 1989. Geochemical evidence for suppression of pelagic marine productivity at the Cretaceous/Tertiary boundary. Nature, 337:6164.CrossRefGoogle Scholar
Zhao, M., Heinsch, F. A., Nemani, R. R., and Running, S. W. 2005. Improvements of the MODIS terrestrial gross and net primary production global data set. Remote Sensing of Environment, 95:164176.CrossRefGoogle Scholar