Capitanian mass extinction event

From Wikipedia, the free encyclopedia
Extinction intensity.svgCambrianOrdovicianSilurianDevonianCarboniferousPermianTriassicJurassicCretaceousPaleogeneNeogene
Marine extinction intensity during the Phanerozoic
%
Millions of years ago
Extinction intensity.svgCambrianOrdovicianSilurianDevonianCarboniferousPermianTriassicJurassicCretaceousPaleogeneNeogene
Plot of extinction intensity (percentage of genera that are present in each interval of time but do not exist in the following interval) vs time in the past for marine genera.[1] Geological periods are annotated (by abbreviation and colour) above. The Capitanian extinction event occurred 260–259 million years ago, ~7 million years before the Permian–Triassic extinction event, with just over 35% (according to this source) failing to survive. (source and image info)

The Capitanian extinction event was an extinction event that occurred around 260 million years ago during a period of decreased species richness and increased extinction rates in the late Middle Permian during the Guadalupian epoch. It is also known as the end-Guadalupian extinction event because of its initial recognition between the Guadalupian and Lopingian series; however, more refined stratigraphic study suggests that extinction peaks in many taxonomic groups occurred within the Guadalupian, in the latter half of the Capitanian age.[2]

Magnitude[]

In the aftermath of Olson's Extinction, global diversity rose during the Capitanian. This was probably the result of disaster taxa replacing extinct guilds. The Capitanian mass extinction greatly reduced disparity (the range of different guilds); eight guilds were lost. It impacted the diversity within individual communities more severely than the Permian–Triassic extinction event.[3] Although faunas began rebuilding complex trophic structures and refilling guilds after the Capitanian extinction event,[3] diversity and disparity fell further until the Permian–Triassic boundary.[4]

Marine ecosystems[]

The impact of the Capitanian extinction event on marine ecosystems is still heavily debated by palaeontologists. Early estimates indicated a loss of marine invertebrate genera between 35–47%,[5][6] while an estimate published in 2016 suggested a loss of 33–35% of marine genera when corrected for background extinction, the Signor–Lipps effect and clustering of extinctions in certain taxa.[7] The loss of marine invertebrates during the Capitanian mass extinction was comparable in magnitude to the Cretaceous–Paleogene extinction event.[8] Some studies have considered it the third or fourth greatest mass extinction in terms of the proportion of marine invertebrate genera lost; a different study found the Capitanian extinction event to be only the ninth worst in terms of taxonomic severity (number of genera lost) but found it to be the fifth worst with regard to its ecological impact (i.e., the degree of taxonomic restructuring within ecosystems or the loss of ecological niches or even entire ecosystems themselves).[9]

Terrestrial ecosystems[]

Few published estimates for the impact on terrestrial ecosystems exist for the Capitanian mass extinction. Among vertebrates, Day and colleagues suggested a 74–80% loss of generic richness in tetrapods of the Karoo Basin in South Africa,[10] including the extinction of the dinocephalians.[11] In land plants, Stevens and colleagues found an extinction of 56% of plant species recorded in the mid-Upper Shihhotse Formation in North China,[12] which was approximately mid-Capitanian in age. 24% of plant species in South China went extinct.[13]

Timing[]

Although it is known that the Capitanian mass extinction occurred after Olson's Extinction and before the Permian–Triassic extinction event,[3] the exact age of the Capitanian mass extinction remains controversial. This is partly due to the somewhat circumstantial age of the Capitanian–Wuchiapingian boundary itself, which is currently estimated to be approximately 259.1 million years old,[10][11][14] but is subject to change by the Subcommission on Permian Stratigraphy of the International Commission on Stratigraphy. Additionally, there is a dispute regarding the severity of the extinction and whether the extinction in China happened at the same time as the extinction in Spitsbergen.[15]

The volcanics of the Emeishan Traps, which are interbedded with tropical carbonate platforms of the , are unique for preserving a mass extinction and the cause of that mass extinction.[13] Large phreatomagmatic eruptions occurred when the Emeishan Traps first started to erupt, leading to the extinction of fusulinacean foraminifera and calcareous algae.[16]

In the absence of radiometric ages directly constraining the extinction horizons themselves in the marine sections, most recent studies refrain from placing a number its age but based on extrapolations from the Permian timescale an age of approximately 260–262 Ma has been estimated;[10][17] this fits broadly with radiometric ages from the terrestrial realm, assuming the two events are contemporaneous. Plant losses occurred either at the same time as the marine extinction or after it.[13]

Marine realm[]

The extinction of fusulinacean foraminifera in Southwest China was originally dated to the end of the Guadalupian, but studies published in 2009 and 2010 dated the extinction of these fusulinaceans to the mid-Capitanian.[18] Brachiopod and coral losses occurred in the middle of the Capitanian stage.[19] The extinction suffered by the ammonoids may have occurred in the early Wuchiapingian.[19]

Terrestrial realm[]

The existence of change in tetrapod faunas in the mid-Permian has long been known in South Africa and Russia. In Russia, it corresponded to the boundary between what became known as the Titanophoneus Superzone and the Scutosaurus Superzone[20] and later the Dinocephalian Superassemblage and the Theriodontian Superassemblage, respectively. In South Africa, this corresponded to the boundary between the variously named Pareiasaurus, Dinocephalian or Tapinocephalus Assemblage Zone and the overlying assemblages.[21][22][23][24] In both Russia and South Africa, this transition was associated with the extinction of the previously dominant group of therapsid amniotes, the dinocephalians, which led to its later designation as the dinocephalian extinction.[25] Post-extinction origination rates remained low through the Pristerognathus Assemblage Zone for at least 1 million years, which suggests that there was a delayed recovery of Karoo Basin ecosystems.[26]

After the recognition of a separate marine mass extinction at the end of the Guadalupian, the dinocephalian extinction was seen to represent its terrestrial correlate.[8] Though it was subsequently suggested that because the Russian , which was considered the youngest dinocephalian fauna in that region, was constrained to below the Illawarra magnetic reversal and therefore had to have occurred in the Wordian stage, well before the end of the Guadalupian,[25] this constraint applied to the type locality only. The recognition of a younger dinocephalian fauna in Russia (the Sundyr Tetrapod Assemblage)[27] and the retrieval of biostratigraphically well-constrained radiometric ages via uranium–lead dating of a tuff from the Tapinocephalus Assemblage Zone of the Karoo Basin[10][28] demonstrated that the dinocephalian extinction did occur in the late Capitanian, around 260 million years ago.

Effects on life[]

Marine life[]

In the oceans, the Capitanian extinction event led to high extinction rates among ammonoids, corals and calcareous algal reef-building organisms, foraminiferans, bryozoans and brachiopods. It appears to have been particularly selective against shallow-water taxa that relied on photosynthesis or a photosymbiotic relationship; many species with poor physiological buffering also became extinct.[29]

The ammonoids, which had been in a long-term decline for a 30 million year period since the Roadian, suffered a selective extinction pulse at the end of the Capitanian.[4] 75.6% of coral families, 77.8% of coral genera and 82.2% of coral species that were in Permian China were lost during the Capitanian mass extinction.[30] The Verbeekinidae, a family of large fusuline foraminifera, went extinct.[31]

87% of brachiopod species found at the Kapp Starostin Formation on Spitsbergen disappeared over a period of tens of thousands of years; though new brachiopod and bivalve species emerged after the extinction, the dominant position of the brachiopods was taken over by the bivalves.[32] Approximately 70% of other species found at the Kapp Starostin Formation also vanished.[33] The fossil record of East Greenland is similar to that of Spitsbergen; the faunal losses in Canada's Sverdrup Basin are comparable to the extinctions in Spitsbergen and East Greenland, but the post-extinction recovery that happened in Spitsbergen and East Greenland did not occur in the Sverdrup Basin.[17] Whereas rhynchonelliform brachiopods made up 99.1% of the individuals found in tropical carbonates in the Western United States, South China and Greece prior to the extinction, molluscs made up 61.2% of the individuals found in similar environments after the extinction.[34] 87% of brachiopod species and 82% of fusulinacean foraminifer species in South China were lost.[17]

Most of the marine victims of the extinction were either endemic species of epicontinental seas around Pangaea that died when the seas closed, or were dominant species of the Paleotethys.[35]

After the Capitanian mass extinction, disaster taxa such as Earlandia and Diplosphaerina became abundant in what is now South China.[2]

Terrestrial life[]

Among terrestrial vertebrates, the main victims were dinocephalian therapsids, which were one of the most common elements of tetrapod fauna of the Guadalupian; only one dinocephalian genus survived the Capitanian extinction event.[8] The diversity of the anomodonts that lived during the late Guadalupian was cut in half by the Capitanian mass extinction.[36] Terrestrial survivors of the Capitanian extinction event were generally 20 kg (44 lb) to 50 kg (110 lb) and commonly found in burrows.[8]

Causes[]

"A combination of a second-order global drop of sea level and regional constraints, e.g., salinity fluctuations (Weidlich and Bernecker, in press) or collisions of micro-continents account for the destruction of many reef sites."[37]

It is believed that the extinction was triggered by one or more eruptions of the Emeishan Traps, which released a large amount of carbon dioxide and sulfur dioxide into the stratosphere of the Northern and Southern Hemispheres due to the equatorial location of the Emeishan Traps, leading to sudden global cooling and long-term global warming; the eruptions also triggered ocean acidification, a depletion of seafloor oxygen and a severe disturbance of the carbon cycle.[12][15][16][17][29][32] The eruptions would have released high doses of toxic mercury.[38] Basalt piles from the Emeishan Traps currently cover an area of 250,000 to 500,000 km2, but the original volume of the basalts may have been anywhere from 500,000 km3 to over 1,000,000 km3.[29] Acid rain, global drying, plate tectonics, marine regression and biological competition may have also played a role in the extinction.[2][12][30] The extinction coincided with the beginning of a major negative δ13C excursion.[13] Analysis of vertebrate extinction rates in the Karoo Basin, specifically the upper Abrahamskraal Formation and lower Teekloof Formation, show that the large scale decrease in terrestrial vertebrate diversity coincided with volcanism in the Emeishan Traps, although robust evidence for a causal relationship between these two events remains elusive.[39]

Analysis of tooth apatite δ13C and δ18O values from the tooth apatite of Diictodon feliceps specimens from the Karoo Supergroup reveals that the end of the Capitanian was marked by massive aridification in the region, although the temperature remained largely the same, suggesting that global climate change did not account for the extinction event.[40]

References[]

  1. ^ Rohde, R.A. & Muller, R.A. (2005). "Cycles in fossil diversity". Nature. 434 (7030): 209–210. Bibcode:2005Natur.434..208R. doi:10.1038/nature03339. PMID 15758998. S2CID 32520208.
  2. ^ Jump up to: a b c Bond, D. P. G., Wignall, P. B., Wang, W., Izon, G., Jiang, H. S., Lai, X. L., Sund, Y.-D., Newtona, R.J., Shaoe, L.-Y., Védrinea, S. & Cope, H. (2010). "The mid-Capitanian (Middle Permian) mass extinction and carbon isotope record of South China". Palaeogeography, Palaeoclimatology, Palaeoecology, 292 (1-2), pp. 282-294. https://dx.doi.org/10.1016/j.palaeo.2010.03.056
  3. ^ Jump up to: a b c Sahney, S.; Benton, M.J. (2008). "Recovery from the most profound mass extinction of all time". Proceedings of the Royal Society B. 275 (1636): 759–765. doi:10.1098/rspb.2007.1370. PMC 2596898. PMID 18198148.
  4. ^ Jump up to: a b Villier, L.; Korn, D. (October 2004). "Morphological Disparity of Ammonoids and the Mark of Permian Mass Extinctions". Science. 306 (5694): 264–266. Bibcode:2004Sci...306..264V. doi:10.1126/science.1102127. ISSN 0036-8075. PMID 15472073. S2CID 17304091.
  5. ^ Sepkoski Jr., J. J. (1996). "Patterns of Phanerozoic Extinction: a Perspective from Global Data Bases". In: Walliser, O. H. (Ed.), Global Events and Event Stratigraphy in the Phanerozoic. Springer-Verlag, Berlin, pp. 35–51. https://doi.org/10.1007%2F978-3-642-79634-0_4
  6. ^ Bambach, R. K.; Knoll, A. H. & Wang, S. C. (2004). "Origination, extinction, and mass depletions of marine diversity". Paleobiology. 30 (4): 522–542. doi:10.1666/0094-8373(2004)030<0522:OEAMDO>2.0.CO;2.
  7. ^ Stanley, S. M., 2016. "Estimates of the magnitudes of major marine mass extinctions in earth history". Proceedings of the National Academy of Sciences of the United States of America, 113 (42), E6325–E6334.
  8. ^ Jump up to: a b c d Retallack, G. J., Metzger, C. A., Greaver, T., Jahren, A. H., Smith, R. M. H. & Sheldon, N. D. (2006). "Middle-Late Permian mass extinction on land". Geological Society of America Bulletin 118 (11-12): 1398-1411.
  9. ^ McGhee, G.R., Sheehan, P.M., Bottjer, D.J. & Droser, M.L., 2004. "Ecological ranking of Phanerozoic biodiversity crises: Ecological and taxonomic severities are decoupled". Palaeogeography, Palaeoclimatology, Palaeoecology 211 (3-4), pp. 289–297. https://doi.org/10.1016/j.palaeo.2004.05.010
  10. ^ Jump up to: a b c d Day, M.O., Ramezani, J., Bowring, S.A., Sadler, P.M., Erwin, D.H., Abdala, F. and Rubidge, B.S., July 2015. "When and how did the terrestrial mid-Permian mass extinction occur? Evidence from the tetrapod record of the Karoo Basin, South Africa". Proceedings of the Royal Society B 282 (1811). https://doi.org/10.1098/rspb.2015.0834
  11. ^ Jump up to: a b "South Africa's Great Karoo reveals mass extinction". ScienceDaily. July 7, 2015.
  12. ^ Jump up to: a b c Stevens, L.G., Hilton, J., Bond, D.P.G., Glasspool, I.J. & Jardine, P.E., 2011. "Radiation and extinction patterns in Permian floras from North China as indicators for environmental and climate change". Journal of the Geological Society 168, pp. 607–619.
  13. ^ Jump up to: a b c d Keller, Gerta; Kerr, Andrew C., eds. (2014). Volcanism, Impacts, and Mass Extinctions: Causes and Effects. The Geological Society of America. p. 37. ISBN 978-0-8137-2505-5.
  14. ^ Zhong, Y.-T., He, B., Mundil, R., and Xu, Y.-G. (2014). CA-TIMS zircon U–Pb dating of felsic ignimbrite from the Binchuan section: Implications for the termination age of Emeishan large igneous province. Lithos 204, pp. 14-19.
  15. ^ Jump up to: a b Hand, Eric (April 16, 2015). "Sixth extinction, rivaling that of the dinosaurs, should join the big five, scientists say". Science.
  16. ^ Jump up to: a b P. B. Wignall, Y. Sun, D. P. G. Bond, G. Izon, R. J. Newton, S. Védrine, M. Widdowson, J. R. Ali, X. Lai, H. Jiang, H. Cope & S. H. Bottrell (2009). Volcanism, Mass Extinction, and Carbon Isotope Fluctuations in the Middle Permian of China. Science 324 (5931), pp. 1179-1182.
  17. ^ Jump up to: a b c d Bond, D.P.G., Wignall, P.B., Joachimski, M.M., Sun, Y., Savov, I., Grasby, S.E., Beauchamp, B. and Blomeier, D.P. 2015. An abrupt extinction in the Middle Permian (Capitanian) of the Boreal Realm (Spitsbergen) and its link to anoxia and acidification. Geological Society of America Bulletin, 127 (9-10): 1411-1421.
  18. ^ Keller, Gerta; Kerr, Andrew C., eds. (2014). Volcanism, Impacts, and Mass Extinctions: Causes and Effects. The Geological Society of America. pp. 36–37. ISBN 978-0-8137-2505-5.
  19. ^ Jump up to: a b Bond, D.P.G.; Hilton, J.; Wignall, P.B.; Ali, J.R.; Steven, L.G.; Sun, Y.; Lai, X. September 2010. The Middle Permian (Capitanian) mass extinction on land and in the oceans. Earth-Science Reviews 102 (1-2): 100-116.
  20. ^ Ivakhnenko, M. F., Golubev, V. K., Gubin, Yu. M., Kalandadze, N. N., Novikov, I. V., Sennikov, A. G. & Rautian, A. S. 1997. "Permskiye I Triasovyye tetrapody vostochnoi Evropy (Permian and Triassic Tetrapods of Eastern Europe)". GEOS, Moscow (original in Russian).
  21. ^ Kitching, James W. (1977). The Distribution of the Karroo Vertebrate Fauna: With Special Reference to Certain Genera and the Bearing of this Distribution on the Zoning of the Beaufort Beds. Memoir No. 1. Johannesburg: Bernard Price Institute for Palaeontological Research, University of the Witwatersrand. pp. 1–131. ISBN 9780854944279.
  22. ^ Keyser, A. W. & Smith, R. H. M. 1979. "Vertebrate biozonation of the Beaufort Group with special reference to the Western Karoo Basin." Annals of the Geological Survey of South Africa 12, pp. 1-36.
  23. ^ Broom, R. 1906. "On the Permian and Triassic Faunas of South Africa." Geological Magazine 3 (1), pp. 29-30. https://doi.org/10.1017/S001675680012271X
  24. ^ Watson, D. M. S. May 1914. "The Zones of the Beaufort Beds of the Karoo System in South Africa." Geological Magazine 1 (5), pp. 203-208. https://doi.org/10.1017/S001675680019675X
  25. ^ Jump up to: a b Lucas, S. G. (2009). "Timing and magnitude of tetrapod extinctions across the Permo-Triassic boundary". Journal of Asian Earth Sciences. 36 (6): 491–502. Bibcode:2009JAESc..36..491L. doi:10.1016/j.jseaes.2008.11.016.
  26. ^ M. O. Day. The Pristerognathus AZ and the aftermath of the Capitanian extinction event in the main Karoo Basin. 35th International Geological Congress. Cape Town.
  27. ^ Golubev, V. K. (2015). "Dinocephalian stage in the history of the Permian tetrapod fauna of Eastern Europe". Paleontological Journal. 49 (12): 1346–1352. doi:10.1134/S0031030115120059. S2CID 130694755.
  28. ^ Rubidge, B. S., Erwin, D. H., Ramezani, J., Bowring, S. A. & De Klerk, W. J. (2013). "High-precision temporal calibration of Late Permian vertebrate biostratigraphy: U-Pb zircon constraints from the Karoo Supergroup, South Africa". Geology 41 (3): 363-366.
  29. ^ Jump up to: a b c Keller, Gerta; Kerr, Andrew C., eds. (2014). Volcanism, Impacts, and Mass Extinctions: Causes and Effects. The Geological Society of America. p. 38. ISBN 978-0-8137-2505-5.
  30. ^ Jump up to: a b Wang, X.-D. & Sugiyama, T. (December 2000). "Diversity and extinction patterns of Permian coral faunas of China". Lethaia. 33 (4): 285–294. doi:10.1080/002411600750053853.
  31. ^ Ota, A. & Isozaki, Y. (March 2006). "Fusuline biotic turnover across the Guadalupian–Lopingian (Middle–Upper Permian) boundary in mid-oceanic carbonate buildups: Biostratigraphy of accreted limestone in Japan". Journal of Asian Earth Sciences. 26 (3–4): 353–368. Bibcode:2006JAESc..26..353O. doi:10.1016/j.jseaes.2005.04.001.
  32. ^ Jump up to: a b Berezow, Alex (April 21, 2015). "New mass extinction event identified by geologists". BBC.
  33. ^ Kaplan, Sarah (April 23, 2015). "Tantalizing evidence of a mass extinction". The Washington Post.
  34. ^ Clapham, M.E., Bottjer, D.J. and Shen, S. (2006). "Decoupled diversity and ecology during the end-Guadalupian extinction (late Permian)". Geological Society of America Abstracts with Programs. 38 (7): 117. Archived from the original on 2015-12-08. Retrieved 2019-07-10.CS1 maint: multiple names: authors list (link)
  35. ^ Yugan, J.; Jing, Z. & Shang, Q. (1994). "Two Phases of the End-Permian Mass Extinction". Pangea: Global Environments and Resources (Memoir 17): 813–822.
  36. ^ J. Fröbisch (November 2008). "Global Taxonomic Diversity of Anomodonts (Tetrapoda, Therapsida) and the Terrestrial Rock Record Across the Permian-Triassic Boundary". PLOS ONE. 3 (11): e3733. Bibcode:2008PLoSO...3.3733F. doi:10.1371/journal.pone.0003733. PMC 2581439. PMID 19011684.
  37. ^ Weidlich, O. (2002). "Permian reefs re-examined: extrinsic control mechanisms of gradual and abrupt changes during 40 my of reef evolution". Geobios. 35 (1): 287–294. doi:10.1016/S0016-6995(02)00066-9.
  38. ^ Freydlin, Julie (July 16, 2015). "Brachiopod die-off signaled mid-Permian mass extinction". Earth Magazine.
  39. ^ Michael O. Day; Bruce S. Rubidge (18 February 2021). "The Late Capitanian Mass Extinction of Terrestrial Vertebrates in the Karoo Basin of South Africa". Frontiers in Earth Science. 9: 15. Bibcode:2021FrEaS...9...15D. doi:10.3389/feart.2021.631198.
  40. ^ Kévin Rey; Michael O. Day; Romain Amiot; Jean Goedert; Christophe Lécuyer; Judith Sealy; Bruce S. Rubidge (July 2018). "Stable isotope record implicates aridification without warming during the late Capitanian mass extinction". Gondwana Research. 59: 1–8. Bibcode:2018GondR..59....1R. doi:10.1016/j.gr.2018.02.017.
Retrieved from ""