Capitanian mass extinction event

Last updated
Extinction intensity.svg
Marine extinction intensity during Phanerozoic
%
Millions of years ago
(H)
Cap
Extinction intensity.svg
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 mass extinction event, also known as the end-Guadalupian extinction event, [2] the Guadalupian-Lopingian boundary mass extinction, [3] the pre-Lopingian crisis, [4] or the Middle Permian extinction, was an extinction event that predated the end-Permian extinction event. The mass extinction occurred during a period of decreased species richness and increased extinction rates near the end of the Middle Permian, also known as the Guadalupian epoch. It is often called 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. [5] The extinction event has been argued to have begun around 262 million years ago with the Late Guadalupian crisis, though its most intense pulse occurred 259 million years ago in what is known as the Guadalupian-Lopingian boundary event. [6]

Contents

Having historically been considered as part of the end-Permian extinction event, and only viewed as separate relatively recently,[ when? ] [7] this mass extinction is believed to be the third largest of the Phanerozoic in terms of the percentage of species lost, after the end-Permian and Late Ordovician mass extinctions, respectively, [8] while being the fifth worst in terms of ecological severity. [9] The global nature of the Capitanian mass extinction has been called into question by some palaeontologists as a result of some analyses finding it to have affected only low-latitude taxa in the Northern Hemisphere. [10]

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. [11] Although faunas began recovery immediately after the Capitanian extinction event, [12] [13] rebuilding complex trophic structures and refilling guilds, [11] diversity and disparity fell further until the Permian–Triassic boundary. [14]

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 and 47%, [15] [16] 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. [17] The loss of marine invertebrates during the Capitanian mass extinction was comparable in magnitude to the Cretaceous–Paleogene extinction event. [18] 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). [19]

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, [20] including the extinction of the dinocephalians. [21] In land plants, Stevens and colleagues found an extinction of 56% of plant species recorded in the mid-Upper Shihhotse Formation in North China, [22] which was approximately mid-Capitanian in age. 24% of plant species in South China went extinct. [23]

Timing

Although it is known that the Capitanian mass extinction occurred after Olson's Extinction and before the Permian–Triassic extinction event, [11] 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, [20] [21] [24] 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. [25] According to one study, the Capitanian mass extinction was not one discrete event but a continuous decline in diversity that began at the end of the Wordian. [26] Another study examining fossiliferous facies in Svalbard found no evidence for a sudden mass extinction, instead attributing local biotic changes during the Capitanian to the southward migration of many taxa through the Zechstein Sea. [27] Carbonate platform deposits in Hungary and Hydra show no sign of an extinction event at the end of the Capitanian; the extinction event there is recorded in the middle Capitanian. [28]

The volcanics of the Emeishan Traps, which are interbedded with tropical carbonate platforms of the Maokou Formation, are unique for preserving a mass extinction and the cause of that mass extinction. [23] Large phreatomagmatic eruptions occurred when the Emeishan Traps first started to erupt, leading to the extinction of fusulinacean foraminifera and calcareous algae. [29]

In the absence of radiometric ages directly constraining the extinction horizons themselves in the marine sections, most recent studies refrain from placing a number on its age, but based on extrapolations from the Permian timescale an age of approximately 260–262 Ma has been estimated; [20] [30] 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. [23]

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. [31] Brachiopod and coral losses occurred in the middle of the Capitanian stage. [32] The extinction suffered by the ammonoids may have occurred in the early Wuchiapingian. [32]

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 [33] 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. [34] [35] [36] [37] 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. [38] 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. [39]

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. [18] Though it was subsequently suggested that because the Russian Ischeevo fauna, 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, [38] this constraint applied to the type locality only. The recognition of a younger dinocephalian fauna in Russia (the Sundyr Tetrapod Assemblage) [40] 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 [20] [41] 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, foraminifera, bryozoans, and brachiopods. It was more severe in restricted marine basins than in the open oceans. [42] It appears to have been particularly selective against shallow-water taxa that relied on photosynthesis or a photosymbiotic relationship; [43] many species with poorly buffered respiratory physiologies also became extinct. [44] [45] The extinction event led to a collapse of the reef carbonate factory in the shallow seas surrounding South China. [46] [47]

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. [14] 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. [48] The Verbeekinidae, a family of large fusuline foraminifera, went extinct. [49]

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. [50] Approximately 70% of other species found at the Kapp Starostin Formation also vanished. [51] 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. [30] 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. [52] 87% of brachiopod species and 82% of fusulinacean foraminifer species in South China were lost. [30] Although severe for brachiopods, the Capitanian extinction's impact on their diversity was nowhere near as strong as that of the later end-Permian extinction. [53]

Biomarker evidence indicates red algae and photoautotrophic bacteria dominated marine microbial communities. Significant turnovers in microbial ecosystems occurred during the Capitanian mass extinction, though they were smaller in magnitude than those associated with the end-Permian extinction. [54]

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 Ocean. [55] Evidence from marine deposits in Japan and Primorye suggests that mid-latitude marine life became affected earlier by the extinction event than marine organisms of the tropics. [56]

Whether and to what degree latitude affected the likelihood of taxa to go extinct remains disputed amongst palaeontologists. Whereas some studies conclude that the extinction event was a regional one limited to tropical areas, [10] others suggest that there was little latitudinal variation in extinction patterns. [57] A study examining foraminiferal extinctions in particular found that the Central and Western Palaeotethys experienced taxonomic losses of a lower magnitude than the Northern and Eastern Palaeotethys, which had the highest extinction magnitude. The same study found that Panthalassa's overall extinction magnitude was similar to that of the Central and Western Palaeotethys, but that it had a high magnitude of extinction of endemic taxa. [58]

This mass extinction marked the beginning of the transition between the Palaeozoic and Modern evolutionary faunas. [2] The brachiopod-mollusc transition that characterised the broader shift from the Palaeozoic to Modern evolutionary faunas has been suggested to have had its roots in the Capitanian mass extinction event, although other research has concluded that this may be an illusion created by taphonomic bias in silicified fossil assemblages, with the transition beginning only in the aftermath of the more cataclysmic end-Permian extinction. [59] After the Capitanian mass extinction, disaster taxa such as Earlandia and Diplosphaerina became abundant in what is now South China. [5] The initial recovery of reefs consisted of non-metazoan reefs: algal bioherms and algal-sponge reef buildups. This initial recovery interval was followed by an interval of Tubiphytes-dominated reefs, which in turn was followed by a return of metazoan, sponge-dominated reefs. [60] Overall, reef recovery took approximately 2.5 million years. [3]

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. [18] The diversity of the anomodonts that lived during the late Guadalupian was cut in half by the Capitanian mass extinction. [61] Terrestrial survivors of the Capitanian extinction event were generally 20 kg (44 lb) to 50 kg (110 lb) and commonly found in burrows. [18]

Causes

Emeishan Traps

Volcanic emissions

It is believed that the extinction, which coincided with the beginning of a major negative δ13C excursion signifying a severe disturbance of the carbon cycle, [62] [23] was triggered by eruptions of the Emeishan Traps large igneous province, [63] [64] [22] basalt piles from which currently cover an area of 250,000 to 500,000 km2, although the original volume of the basalts may have been anywhere from 500,000 km3 to over 1,000,000 km3. [45] The age of the extinction event and the deposition of the Emeishan basalts are in good alignment. [65] [66] Reefs and other marine sediments interbedded among basalt piles indicate Emeishan volcanism initially developed underwater; terrestrial outflows of lava occurred only later in the large igneous province's period of activity. [67] These eruptions would have released high doses of toxic mercury; [68] [69] increased mercury concentrations are coincident with the negative carbon isotope excursion, indicating a common volcanic cause. [70] Coronene enrichment at the Guadalupian-Lopingian boundary further confirms the existence of massive volcanic activity; coronene can only form at extremely high temperatures created either by extraterrestrial impacts or massive volcanism, with the former being ruled out because of an absence of iridium anomalies coeval with mercury and coronene anomalies. [71] A large amount of carbon dioxide and sulphur dioxide is believed to have been discharged 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. [29] The Emeishan Traps discharged between 130 and 188 teratonnes of carbon dioxide in total, doing so at a rate of between 0.08 to 0.25 gigatonnes of carbon dioxide per year, making them responsible for an increase in atmospheric carbon dioxide that was both one of the largest and one of the most precipitous in the entire geological history of the Earth. [72] The rate of carbon dioxide emissions during the Capitanian mass extinction, though extremely abrupt, was nonetheless significantly slower than that during the end-Permian extinction, during which carbon dioxide levels rose five times faster according to one study. [73] Significant quantities of methane released by dikes and sills intruding into coal-rich deposits has been implicated as an additional driver of warming, [74] though this idea has been challenged by studies that instead conclude that the extinction was precipitated directly by the Emeishan Traps or by their interaction with platform carbonates. [75] [76] [77] The emissions of the Emeishan Traps may also have contributed to the downfall of the ozone shield, exposing the Earth's surface to a vastly increased flux of high-frequency solar radiation. [78]

Anoxia and euxinia

Global warming resulting from the large igneous province's activity has been implicated as a cause of marine anoxia. [79] Two anoxic events, the middle Capitanian OAE-C1 and the end-Capitanian OAE-C2, occurred thanks to Emeishan volcanic activity. [80] Volcanic greenhouse gas release and global warming increased continental weathering and mineral erosion, which in turn has been propounded as a factor enhancing oceanic euxinia. [81] Euxinia may have been exacerbated even further by the increasing sluggishness of ocean circulation resulting from volcanically driven warming. [82] The initial hydrothermal nature of the Emeishan Traps meant that local marine life around South China would have been especially jeopardised by anoxia due to hyaloclastite development in restricted, fault-bounded basins. [67] Expansion of oceanic anoxia has been posited to have occurred slightly before the Capitanian extinction event itself by some studies, though it is probable that upwelling of anoxic waters prior to the mass extinction was a local phenomenon specific to South China. [83]

Hypercapnia and acidification

Because the ocean acts as a carbon sink absorbing atmospheric carbon dioxide, it is likely that the excessive volcanic emissions of carbon dioxide resulted in marine hypercapnia, which would have acted in conjunction with other killing mechanisms to further increase the severity of the biotic crisis. [84] The dissolution of volcanically emitted carbon dioxide in the oceans triggered ocean acidification, [30] [25] [50] which probably contributed to the demise of various calcareous marine organisms, particularly giant alatoconchid bivalves. [85] By virtue of the greater solubility of carbon dioxide in colder waters, ocean acidification was especially lethal in high latitude waters. [79] Furthermore, acid rain would have arisen as yet another biocidal consequence of the intense sulphur emissions produced by Emeishan Traps volcanism. [29] This resulted in soil acidification and a decline of terrestrial infaunal invertebrates. [86] Some researchers have cast doubt on whether significant acidification took place globally, concluding that the carbon cycle perturbation was too small to have caused a major worldwide drop in pH. [87]

Criticism of the volcanic cause hypothesis

Not all studies, however, have supported the volcanic warming hypothesis; analysis of δ13C and δ18O values from the tooth apatite of Diictodon feliceps specimens from the Karoo Supergroup shows a positive δ13C excursion and concludes 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. [88] 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. [89] A 2015 study called into question whether the Capitanian mass extinction event was global in nature at all or merely a regional biotic crisis limited to South China and a few other areas, finding no evidence for terrestrial or marine extinctions in eastern Australia linked to the Emeishan Traps or to any proposed extinction triggers invoked to explain the biodiversity drop in low-latitudes of the Northern Hemisphere. [10]

Sea level fall

The Capitanian mass extinction has been attributed to sea level fall, [90] with the widespread demise of reefs in particular being linked to this marine regression. [84] The Guadalupian-Lopingian boundary coincided with one of the most prominent first-order marine regressions of the Phanerozoic. [5] Evidence for abrupt sea level fall at the terminus of the Guadalupian comes from evaporites and terrestrial facies overlying marine carbonate deposits across the Guadalupian-Lopingian transition. [84] Additionally, a tremendous unconformity is associated with the Guadalupian-Lopingian boundary in many strata across the world. [91] The closure of the Sino-Mongolian Seaway at the end of the Capitanian has been invoked as a potential driver of Palaeotethyan biodiversity loss. [92]

Other hypotheses

Global drying, plate tectonics, and biological competition may have also played a role in the extinction. [5] [22] [48] [88] Potential drivers of extinction proposed as causes of end-Guadalupian reef decline include fluctuations in salinity and tectonic collisions of microcontinents. [84]

Related Research Articles

<span class="mw-page-title-main">Extinction event</span> Widespread and rapid decrease in the biodiversity on Earth

An extinction event is a widespread and rapid decrease in the biodiversity on Earth. Such an event is identified by a sharp fall in the diversity and abundance of multicellular organisms. It occurs when the rate of extinction increases with respect to the background extinction rate and the rate of speciation. Estimates of the number of major mass extinctions in the last 540 million years range from as few as five to more than twenty. These differences stem from disagreement as to what constitutes a "major" extinction event, and the data chosen to measure past diversity.

<span class="mw-page-title-main">Permian</span> Sixth and last period of the Paleozoic Era 299–252 million years ago

The Permian is a geologic period and stratigraphic system which spans 47 million years from the end of the Carboniferous Period 298.9 million years ago (Mya), to the beginning of the Triassic Period 251.902 Mya. It is the last period of the Paleozoic Era; the following Triassic Period belongs to the Mesozoic Era. The concept of the Permian was introduced in 1841 by geologist Sir Roderick Murchison, who named it after the region of Perm in Russia.

<span class="mw-page-title-main">Permian–Triassic extinction event</span> Earths most severe extinction event

Approximately 251.9 million years ago, the Permian–Triassicextinction event forms the boundary between the Permian and Triassic geologic periods, and with them the Paleozoic and Mesozoic eras. It is the Earth's most severe known extinction event, with the extinction of 57% of biological families, 83% of genera, 81% of marine species and 70% of terrestrial vertebrate species. It is also the greatest known mass extinction of insects. It is the greatest of the "Big Five" mass extinctions of the Phanerozoic. There is evidence for one to three distinct pulses, or phases, of extinction.

<span class="mw-page-title-main">Late Ordovician mass extinction</span> Extinction event around 444 million years ago

The Late Ordovician mass extinction (LOME), sometimes known as the end-Ordovician mass extinction or the Ordovician-Silurian extinction, is the first of the "big five" major mass extinction events in Earth's history, occurring roughly 445 million years ago (Ma). It is often considered to be the second-largest known extinction event, in terms of the percentage of genera that became extinct. Extinction was global during this interval, eliminating 49–60% of marine genera and nearly 85% of marine species. Under most tabulations, only the Permian-Triassic mass extinction exceeds the Late Ordovician mass extinction in biodiversity loss. The extinction event abruptly affected all major taxonomic groups and caused the disappearance of one third of all brachiopod and bryozoan families, as well as numerous groups of conodonts, trilobites, echinoderms, corals, bivalves, and graptolites. Despite its taxonomic severity, the Late Ordovician mass extinction did not produce major changes to ecosystem structures compared to other mass extinctions, nor did it lead to any particular morphological innovations. Diversity gradually recovered to pre-extinction levels over the first 5 million years of the Silurian period.

<span class="mw-page-title-main">Late Devonian extinction</span> One of the five most severe extinction events in the history of the Earths biota

The Late Devonian extinction consisted of several extinction events in the Late Devonian Epoch, which collectively represent one of the five largest mass extinction events in the history of life on Earth. The term primarily refers to a major extinction, the Kellwasser event, also known as the Frasnian-Famennian extinction, which occurred around 372 million years ago, at the boundary between the Frasnian stage and the Famennian stage, the last stage in the Devonian Period. Overall, 19% of all families and 50% of all genera became extinct. A second mass extinction called the Hangenberg event, also known as the end-Devonian extinction, occurred 359 million years ago, bringing an end to the Famennian and Devonian, as the world transitioned into the Carboniferous Period.

The Guadalupian is the second and middle series/epoch of the Permian. The Guadalupian was preceded by the Cisuralian and followed by the Lopingian. It is named after the Guadalupe Mountains of New Mexico and Texas, and dates between 272.95 ± 0.5 – 259.1 ± 0.4 Mya. The series saw the rise of the therapsids, a minor extinction event called Olson's Extinction and a significant mass extinction called the end-Capitanian extinction event. The Guadalupian was previously known as the Middle Permian.

The Lopingian is the uppermost series/last epoch of the Permian. It is the last epoch of the Paleozoic. The Lopingian was preceded by the Guadalupian and followed by the Early Triassic.

In the geologic timescale, the Capitanian is an age or stage of the Permian. It is also the uppermost or latest of three subdivisions of the Guadalupian Epoch or Series. The Capitanian lasted between 264.28 and 259.51 million years ago. It was preceded by the Wordian and followed by the Wuchiapingian.

<span class="mw-page-title-main">Early Triassic</span> First of three epochs of the Triassic Period

The Early Triassic is the first of three epochs of the Triassic Period of the geologic timescale. It spans the time between 251.9 Ma and 247.2 Ma. Rocks from this epoch are collectively known as the Lower Triassic Series, which is a unit in chronostratigraphy. The Early Triassic is the oldest epoch of the Mesozoic Era. It is preceded by the Lopingian Epoch and followed by the Middle Triassic Epoch. The Early Triassic is divided into the Induan and Olenekian ages. The Induan is subdivided into the Griesbachian and Dienerian subages and the Olenekian is subdivided into the Smithian and Spathian subages.

<span class="mw-page-title-main">Olenekian</span> Age in the Early Triassic epoch

In the geologic timescale, the Olenekian is an age in the Early Triassic epoch; in chronostratigraphy, it is a stage in the Lower Triassic series. It spans the time between 251.2 Ma and 247.2 Ma. The Olenekian is sometimes divided into the Smithian and the Spathian subages or substages. The Olenekian follows the Induan and is followed by the Anisian.

<span class="mw-page-title-main">Beaufort Group</span> Third of the main subdivisions of the Karoo Supergroup in South Africa

The Beaufort Group is the third of the main subdivisions of the Karoo Supergroup in South Africa. It is composed of a lower Adelaide Subgroup and an upper Tarkastad Subgroup. It follows conformably after the Ecca Group and unconformably underlies the Stormberg Group. Based on stratigraphic position, lithostratigraphic and biostratigraphic correlations, palynological analyses, and other means of geological dating, the Beaufort Group rocks are considered to range between Middle Permian (Wordian) to Early Triassic (Anisian) in age.

<i>Anteosaurus</i> Extinct genus of anteosaurid synapsid from the Permian

Anteosaurus is an extinct genus of large carnivorous dinocephalian synapsid. It lived at the end of the Guadalupian during the Capitanian stage, about 265 to 260 million years ago in what is now South Africa. It is mainly known by cranial remains and few postcranial bones. Measuring 5–6 m (16–20 ft) long and weighing about 600 kg (1,300 lb), Anteosaurus was the largest known carnivorous non-mammalian synapsid and the largest terrestrial predator of the Permian period. Occupying the top of the food chain in the Middle Permian, its skull, jaws and teeth show adaptations to capture large prey like the giants titanosuchids and tapinocephalids dinocephalians and large pareiasaurs.

<i>Daptocephalus</i> Assemblage Zone

The Daptocephalus Assemblage Zone is a tetrapod assemblage zone or biozone found in the Adelaide Subgroup of the Beaufort Group, a majorly fossiliferous and geologically important geological Group of the Karoo Supergroup in South Africa. This biozone has outcrops located in the upper Teekloof Formation west of 24°E, the majority of the Balfour Formation east of 24°E, and the Normandien Formation in the north. It has numerous localities which are spread out from Colesberg in the Northern Cape, Graaff-Reniet to Mthatha in the Eastern Cape, and from Bloemfontein to Harrismith in the Free State. The Daptocephalus Assemblage Zone is one of eight biozones found in the Beaufort Group and is considered Late Permian (Lopingian) in age. Its contact with the overlying Lystrosaurus Assemblage Zone marks the Permian-Triassic boundary.

The Emeishan Traps constitute a flood basalt volcanic province, or large igneous province, in south-western China, centred in Sichuan province. It is sometimes referred to as the Permian Emeishan Large Igneous Province or Emeishan Flood Basalts. Like other volcanic provinces or "traps", the Emeishan Traps are multiple layers of igneous rock laid down by large mantle plume volcanic eruptions. The Emeishan Traps eruptions were serious enough to have global ecological and paleontological impact.

<span class="mw-page-title-main">Abrahamskraal Formation</span> Geological formation of the Beaufort Group in South Africa

The Abrahamskraal Formation is a geological formation and is found in numerous localities in the Northern Cape, Western Cape, and the Eastern Cape of South Africa. It is the lowermost formation of the Adelaide Subgroup of the Beaufort Group, a major geological group that forms part of the greater Karoo Supergroup. It represents the first fully terrestrial geological deposits of the Karoo Basin. Outcrops of the Abrahamskraal Formation are found from the small town Middelpos in its westernmost localities, then around Sutherland, the Moordenaarskaroo north of Laingsburg, Williston, Fraserburg, Leeu-Gamka, Loxton, and Victoria West in the Western Cape and Northern Cape. In the Eastern Cape outcrops are known from Rietbron, north of Klipplaat and Grahamstown, and also southwest of East London.

<i>Pampaphoneus</i> Extinct genus of therapsids

Pampaphoneus is an extinct genus of carnivorous dinocephalian therapsid belonging to the family Anteosauridae. It lived 268 to 265 million years ago during the Wordian age of the Guadalupian period in what is now Brazil. Pampaphoneus is known by an almost complete skull with the lower jaw still articulated, discovered on the lands of the Boqueirão Farm, near the city of São Gabriel, in the state of Rio Grande do Sul. A second specimen from the same locality was reported in 2019 and 2020 but has not yet been described. It is composed of a skull associated with postcranial remains. It is the first South American species of dinocephalian to have been described. The group was previously known in South America only by a few isolated teeth and a jaw fragment reported in 2000 in the same region of Brazil. Phylogenetic analysis conducted by Cisneros and colleagues reveals that Pampaphoneus is closely related to anteosaurs from European Russia, indicating a closer faunal relationship between South America and Eastern Europe than previously thought, thus promoting a Pangea B continental reconstruction.

<span class="mw-page-title-main">Middleton Formation</span> Late middle Permian geological formation in the Eastern Cape

The Middleton Formation is a geological formation that extends through the Northern Cape, Western Cape, and Eastern Cape provinces of South Africa. It overlies the lower Abrahamskraal Formation, and is the eastern correlate, East of 24ºE, of the Teekloof Formation. Outcrops and exposures of the Middleton Formation range from Graaff-Reinet in the Eastern Cape onwards. The Middleton Formation's type locality lies near the small hamlet, Middleton, approximately 25 km south of Cookhouse. Other exposures lie in hillsides along the Great Fish River in the Eastern Cape. The Middleton Formation forms part of the Adelaide Subgroup of the Beaufort Group, which itself forms part of the Karoo Supergroup.

<span class="mw-page-title-main">Alatoconchidae</span> Extinct family of molluscs

Alatoconchidae is an extinct family of prehistoric bivalves that lived in the early to middle Permian period. Genera belonging to Alatoconchidae are characterized by their shell that is strongly compressed in the dorsoventral direction. Some species reached large sizes of as much as 1 metre (3.3 ft) long. It is hypothesized that some species in this family got energy from chemosynthetic bacteria.

The Toarcian extinction event, also called the Pliensbachian-Toarcian extinction event, the Early Toarcian mass extinction, the Early Toarcian palaeoenvironmental crisis, or the Jenkyns Event, was an extinction event that occurred during the early part of the Toarcian age, approximately 183 million years ago, during the Early Jurassic. The extinction event had two main pulses, the first being the Pliensbachian-Toarcian boundary event (PTo-E). The second, larger pulse, the Toarcian Oceanic Anoxic Event (TOAE), was a global oceanic anoxic event, representing possibly the most extreme case of widespread ocean deoxygenation in the entire Phanerozoic eon. In addition to the PTo-E and TOAE, there were multiple other, smaller extinction pulses within this span of time.

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. 1 2 De la Horra, R.; Galán-Abellán, A. B.; López-Gómez, José; Sheldon, Nathan D.; Barrenechea, J. F.; Luque, F. J.; Arche, A.; Benito, M. I. (August–September 2012). "Paleoecological and paleoenvironmental changes during the continental Middle–Late Permian transition at the SE Iberian Ranges, Spain". Global and Planetary Change . 94–95: 46–61. Bibcode:2012GPC....94...46D. doi:10.1016/j.gloplacha.2012.06.008 . Retrieved 15 December 2022.
  3. 1 2 Huang, Yuangeng; Chen, Zhong-Qiang; Zhao, Laishi; Stanley Jr., George D.; Yan, Jiaxin; Pei, Yu; Yang, Wanrong; Huang, Junhua (1 April 2019). "Restoration of reef ecosystems following the Guadalupian–Lopingian boundary mass extinction: Evidence from the Laibin area, South China". Palaeogeography, Palaeoclimatology, Palaeoecology . 519: 8–22. Bibcode:2019PPP...519....8H. doi:10.1016/j.palaeo.2017.08.027. S2CID   134096639 . Retrieved 8 January 2023.
  4. Shen, Shu-Zhong; Shi, G. R. (September 2009). "Latest Guadalupian brachiopods from the Guadalupian/Lopingian boundary GSSP section at Penglaitan in Laibin, Guangxi, South China and implications for the timing of the pre-Lopingian crisis". Palaeoworld . 18 (2–3): 152–161. doi:10.1016/j.palwor.2009.04.010 . Retrieved 21 December 2022.
  5. 1 2 3 4 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
  6. Kaiho, Kunio (22 July 2022). "Relationship between extinction magnitude and climate change during major marine and terrestrial animal crises". Biogeosciences . 19 (14): 3369–3380. Bibcode:2022BGeo...19.3369K. doi: 10.5194/bg-19-3369-2022 . Retrieved 18 March 2023.
  7. Rampino, Michael R.; Shen, Shu-Zhong (5 September 2019). "The end-Guadalupian (259.8 Ma) biodiversity crisis: the sixth major mass extinction?". Historical Biology . 33 (5): 716–722. doi:10.1080/08912963.2019.1658096. S2CID   202858078 . Retrieved 26 December 2022.
  8. Isozaki, Yukio; Servais, Thomas (8 December 2017). "The Hirnantian (Late Ordovician) and end-Guadalupian (Middle Permian) mass-extinction events compared". Lethaia . 51 (2): 173–186. doi:10.1111/let.12252 . Retrieved 23 October 2022.
  9. McGhee Jr., George R.; Clapham, Matthew E.; Sheehan, Peter M.; Bottjer, David J.; Droser, Mary L. (15 January 2013). "A new ecological-severity ranking of major Phanerozoic biodiversity crises". Palaeogeography, Palaeoclimatology, Palaeoecology . 370: 260–270. Bibcode:2013PPP...370..260M. doi:10.1016/j.palaeo.2012.12.019 . Retrieved 2 December 2022.
  10. 1 2 3 Metcalfe, I.; Crowley, J. L.; Nicholl, R. S.; Schmitz, M. (August 2015). "High-precision U-Pb CA-TIMS calibration of Middle Permian to Lower Triassic sequences, mass extinction and extreme climate-change in eastern Australian Gondwana". Gondwana Research . 28 (1): 61–81. Bibcode:2015GondR..28...61M. doi:10.1016/j.gr.2014.09.002 . Retrieved 22 December 2022.
  11. 1 2 3 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.
  12. Zazzali, Sindbad; Crasquin, Sylvie; Deconinck, Jean-François; Feng, Qinglai (25 September 2015). "Biodiversity across the Guadalupian-Lopingian Boundary: first results on the ostracod (Crustacea) fauna, Chaotian section (Sichuan Province, South China)". Geodiversitas . 37 (3): 283–313. doi:10.5252/g2015n3a1. S2CID   128473981 . Retrieved 2 December 2022.
  13. Marchetti, Lorenzo; Logghe, Antoine; Mujal, Eudald; Barrier, Pascal; Montenat, Christian; Nel, André; Pouillon, Jean-Marc; Garrouste, Romain; Steyer, J. Sébastien (1 August 2022). "Vertebrate tracks from the Permian of Gonfaron (Provence, Southern France) and their implications for the late Capitanian terrestrial extinction event". Palaeogeography, Palaeoclimatology, Palaeoecology . 599: 111043. Bibcode:2022PPP...59911043M. doi:10.1016/j.palaeo.2022.111043. S2CID   248597280 . Retrieved 2 December 2022.
  14. 1 2 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.
  15. 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
  16. 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. S2CID   17279135.
  17. 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.
  18. 1 2 3 4 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 Bulletin118 (11-12): 1398-1411.
  19. 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, Palaeoecology211 (3-4), pp. 289–297. https://doi.org/10.1016/j.palaeo.2004.05.010
  20. 1 2 3 4 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 B282 (1811). https://doi.org/10.1098/rspb.2015.0834
  21. 1 2 "South Africa's Great Karoo reveals mass extinction". ScienceDaily. July 7, 2015.
  22. 1 2 3 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 Society168, pp. 607–619.
  23. 1 2 3 4 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.
  24. 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 Archived 2019-06-24 at the Wayback Machine . Lithos204, pp. 14-19.
  25. 1 2 Hand, Eric (April 16, 2015). "Sixth extinction, rivaling that of the dinosaurs, should join the big five, scientists say". Science.
  26. Clapham, Matthew E.; Shen, Shuzhong; Bottjer, David J. (8 April 2016). "The double mass extinction revisited: reassessing the severity, selectivity, and causes of the end-Guadalupian biotic crisis (Late Permian)". Paleobiology . 35 (1): 32–50. doi:10.1666/08033.1. S2CID   26571574 . Retrieved 22 March 2023.
  27. Lee, Sangmin; Shi, Guang R.; Nakrem, Hans A.; Woo, Jusun; Tazawa, Jun-Ichi (4 February 2022). "Mass extinction or extirpation: Permian biotic turnovers in the northwestern margin of Pangea". Geological Society of America Bulletin . 134 (9–10): 2399–2414. Bibcode:2022GSAB..134.2399L. doi:10.1130/B36227.1. hdl:10852/101313. S2CID   245242389 . Retrieved 2 April 2023.
  28. Wignall, Paul B.; Bond, David P. G.; Haas, János; Wang, Wei; Jiang, Haishui; Lai, Xulong; Altiner, Demir; Védrine, Stéphanie; Hips, Kinga; Zajzon, Norbert; Sun, Yadong; Newton, Robert J. (1 February 2012). "CAPITANIAN (MIDDLE PERMIAN) MASS EXTINCTION AND RECOVERY IN WESTERN TETHYS: A FOSSIL, FACIES, AND δ13C STUDY FROM HUNGARY AND HYDRA ISLAND (GREECE)". PALAIOS . 27 (2): 78–89. Bibcode:2012Palai..27...78W. doi:10.2110/palo.2011.p11-058r. S2CID   129448153 . Retrieved 6 April 2023.
  29. 1 2 3 Wignall, Paul B.; Sun, Yadong; Bond, David P. G.; Izon, Gareth; Newton, Robert J.; Védrine, Stéphanie; Widdowson, Mike; Ali, Jason R.; Lai, Xulong; Jiang, Haishui; Cope, Helen; Bottrell, Simon H. (2009). "Volcanism, Mass Extinction, and Carbon Isotope Fluctuations in the Middle Permian of China". Science . 324 (5931): 1179–1182. Bibcode:2009Sci...324.1179W. doi:10.1126/science.1171956. PMID   19478179. S2CID   206519019.
  30. 1 2 3 4 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.
  31. 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.
  32. 1 2 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 Reviews102 (1-2): 100-116.
  33. 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).
  34. 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.
  35. 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 Africa12, pp. 1-36.
  36. Broom, R. 1906. "On the Permian and Triassic Faunas of South Africa." Geological Magazine3 (1), pp. 29-30. https://doi.org/10.1017/S001675680012271X
  37. Watson, D. M. S. May 1914. "The Zones of the Beaufort Beds of the Karoo System in South Africa." Geological Magazine1 (5), pp. 203-208. https://doi.org/10.1017/S001675680019675X
  38. 1 2 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.
  39. 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.
  40. Golubev, V. K. (2015). "Dinocephalian stage in the history of the Permian tetrapod fauna of Eastern Europe". Paleontological Journal . 49 (12): 1346–1352. Bibcode:2015PalJ...49.1346G. doi:10.1134/S0031030115120059. S2CID   130694755.
  41. 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". Geology41 (3): 363-366.
  42. Smith, Benjamin P.; Larson, Toti; Martindale, Rowan C.; Kerans, Charles (15 January 2020). "Impacts of basin restriction on geochemistry and extinction patterns: A case from the Guadalupian Delaware Basin, USA". Earth and Planetary Science Letters . 530: 115876. doi:10.1016/j.epsl.2019.115876 . Retrieved 13 March 2024 via Elsevier Science Direct.
  43. Kofukuda, Daisuke; Isozaki, Yukio; Igo, Hisayoshi (15 March 2014). "A remarkable sea-level drop and relevant biotic responses across the Guadalupian–Lopingian (Permian) boundary in low-latitude mid-Panthalassa: Irreversible changes recorded in accreted paleo-atoll limestones in Akasaka and Ishiyama, Japan". Journal of Asian Earth Sciences . 82: 47–65. Bibcode:2014JAESc..82...47K. doi:10.1016/j.jseaes.2013.12.010 . Retrieved 25 December 2022.
  44. Wei, Hengye; Tang, Zhanwen; Yan, Detian; Wang, Jianguo; Roberts, Andrew P. (30 December 2019). "Guadalupian (Middle Permian) ocean redox evolution in South China and its implications for mass extinction". Chemical Geology . 530: 119318. Bibcode:2019ChGeo.53019318W. doi:10.1016/j.chemgeo.2019.119318. S2CID   210302271 . Retrieved 25 December 2022.
  45. 1 2 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.
  46. Lai, Xulong; Wang, Wei; Wignall, Paul B.; Bond, David P. G.; Jiang, Haishui; Ali, J. R.; John, E. H.; Sun, Yadong (4 November 2008). "Palaeoenvironmental change during the end-Guadalupian (Permian) mass extinction in Sichuan, China". Palaeogeography, Palaeoclimatology, Palaeoecology . 269 (1–2): 78–93. Bibcode:2008PPP...269...78L. doi:10.1016/j.palaeo.2008.08.005 . Retrieved 21 December 2022.
  47. Meng, Qi; Xue, Wuqiang; Chen, Fayao; Yan, Jiaxin; Cai, Jiahua; Sun, Yadong; Wignall, Paul B.; Liu, Ke; Liu, Zhichen; Chen, Deng (May 2022). "Stratigraphy of the Guadalupian (Permian) siliceous deposits from central Guizhou of South China: Regional correlations with implications for carbonate productivity during the Middle Permian biocrisis". Earth-Science Reviews . 228: 104011. Bibcode:2022ESRv..22804011M. doi:10.1016/j.earscirev.2022.104011. S2CID   247754192. Archived from the original on 22 December 2022. Retrieved 21 December 2022.{{cite journal}}: CS1 maint: bot: original URL status unknown (link)
  48. 1 2 Wang, X.-D. & Sugiyama, T. (December 2000). "Diversity and extinction patterns of Permian coral faunas of China". Lethaia . 33 (4): 285–294. Bibcode:2000Letha..33..285W. doi:10.1080/002411600750053853.
  49. 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.
  50. 1 2 Berezow, Alex (April 21, 2015). "New mass extinction event identified by geologists". BBC.
  51. Kaplan, Sarah (April 23, 2015). "Tantalizing evidence of a mass extinction". The Washington Post.
  52. 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.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  53. Shen, Shu-Zhong; Zhang, Yi-Chun (14 July 2015). "Earliest Wuchiapingian (Lopingian, late Permian) brachiopods in southern Hunan, South China: implications for the pre-Lopingian crisis and onset of Lopingian recovery/radiation". Journal of Paleontology . 82 (5): 924–937. doi:10.1666/07-118.1. S2CID   140628267 . Retrieved 22 March 2023.
  54. Xie, Shucheng; Algeo, Thomas J.; Zhou, Wenfeng; Ruan, Xiaoyan; Luo, Genming; Huang, Junhua; Yan, Jiaxin (15 February 2017). "Contrasting microbial community changes during mass extinctions at the Middle/Late Permian and Permian/Triassic boundaries". Earth and Planetary Science Letters . 460: 180–191. Bibcode:2017E&PSL.460..180X. doi:10.1016/j.epsl.2016.12.015 . Retrieved 4 January 2023.
  55. Yugan, J.; Jing, Z. & Shang, Q. (1994). "Two Phases of the End-Permian Mass Extinction". Pangea: Global Environments and Resources (Memoir 17): 813–822.
  56. Kani, Tomomi; Isozaki, Yukio; Hayashi, Ryutaro; Zakharov, Yuri; Popov, Alexander (15 June 2018). "Middle Permian (Capitanian) seawater 87Sr/86Sr minimum coincided with disappearance of tropical biota and reef collapse in NE Japan and Primorye (Far East Russia)". Palaeogeography, Palaeoclimatology, Palaeoecology . 499: 13–21. Bibcode:2018PPP...499...13K. doi:10.1016/j.palaeo.2018.03.033. S2CID   133766806 . Retrieved 2 December 2022.
  57. Allen, Bethany J.; Clapham, Matthew E.; Saupe, Erin E.; Wignall, Paul B.; Hill, Daniel J.; Dunhill, Alexander M. (10 February 2023). "Estimating spatial variation in origination and extinction in deep time: a case study using the Permian–Triassic marine invertebrate fossil record". Paleobiology . 49 (3): 509–526. Bibcode:2023Pbio...49..509A. doi:10.1017/pab.2023.1. S2CID   256801383 . Retrieved 22 March 2023.
  58. Bond, David P. G.; Wignall, Paul B. (8 April 2016). "Latitudinal selectivity of foraminifer extinctions during the late Guadalupian crisis". Paleobiology . 35 (4): 465–483. doi:10.1666/0094-8373-35.4.465. S2CID   140713258 . Retrieved 23 March 2023.
  59. Clapham, Matthew E. (24 February 2015). "Ecological consequences of the Guadalupian extinction and its role in the brachiopod-mollusk transition". Paleobiology . 41 (2): 266–279. Bibcode:2015Pbio...41..266C. doi:10.1017/pab.2014.15. S2CID   33132983 . Retrieved 20 February 2023.
  60. Wang, X.; Foster, W. J.; Yan, J.; Li, A.; Mutti, M. (September 2019). "Delayed recovery of metazoan reefs on the Laibin-Heshan platform margin following the Middle Permian (Capitanian) mass extinction". Global and Planetary Change . 180: 1–15. Bibcode:2019GPC...180....1W. doi:10.1016/j.gloplacha.2019.05.005. S2CID   181887456 . Retrieved 6 January 2023.
  61. 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.
  62. Shen, Shu-zhong; Cao, Chang-qun; Zhang, Hua; Bowring, Samuel A.; Henderson, Charles M.; Payne, Jonathan L.; Davydov, Vladimir I.; Chen, Bo; Yuan, Dong-xun; Zhang, Yi-zhun; Wang, Wei; Zheng, Quan-feng (1 August 2013). "High-resolution δ13Ccarb chemostratigraphy from latest Guadalupian through earliest Triassic in South China and Iran". Earth and Planetary Science Letters . 375: 156–165. Bibcode:2013E&PSL.375..156S. doi:10.1016/j.epsl.2013.05.020 . Retrieved 13 January 2023.
  63. Ling, Kunyue; Wen, Hanjie; Grasby, Stephen E.; Zhao, Haonan; Deng, Changzhou; Yin, Runsheng (5 February 2023). "The Emeishan large igneous province eruption triggered coastal perturbations and the Capitanian mass extinction: Insights from mercury in Permian bauxite beds". Chemical Geology . 617: 121243. Bibcode:2023ChGeo.61721243L. doi:10.1016/j.chemgeo.2022.121243. S2CID   254298090 . Retrieved 14 March 2023.
  64. Arefifard, Sakineh; Payne, Jonathan L. (15 July 2020). "End-Guadalupian extinction of larger fusulinids in central Iran and implications for the global biotic crisis". Palaeogeography, Palaeoclimatology, Palaeoecology . 550: 109743. Bibcode:2020PPP...55009743A. doi:10.1016/j.palaeo.2020.109743. S2CID   216327821 . Retrieved 5 November 2022.
  65. He, Bin; Xu, Yi-Gang; Huang, Xiao-Long; Luo, Zhen-Yu; Shi, Yu-Ruo; Yun, Qi-Jun; Yu, Song-Yue (30 March 2007). "Age and duration of the Emeishan flood volcanism, SW China: Geochemistry and SHRIMP zircon U–Pb dating of silicic ignimbrites, post-volcanic Xuanwei Formation and clay tuff at the Chaotian section". Earth and Planetary Science Letters . 255 (3–4): 306–323. Bibcode:2007E&PSL.255..306H. doi:10.1016/j.epsl.2006.12.021 . Retrieved 15 August 2023.
  66. Huang, Hu; Cawood, Peter A.; Hou, Ming-Cai; Yang, Jiang-Hai; Ni, Shi-Jun; Du, Yuan-Sheng; Yan, Zhao-Kun; Wang, Jun (1 November 2016). "Silicic ash beds bracket Emeishan Large Igneous province to < 1 m.y. at ~ 260 Ma". Lithos . 264: 17–27. Bibcode:2016Litho.264...17H. doi:10.1016/j.lithos.2016.08.013. hdl:10023/11511 . Retrieved 15 August 2023.
  67. 1 2 Jerram, Dougal A.; Widdowson, Mike; Wignall, Paul B.; Sun, Yadong; Lai, Xulong; Bond, David P. G.; Torsvik, Trond H. (1 January 2016). "Submarine palaeoenvironments during Emeishan flood basalt volcanism, SW China: Implications for plume–lithosphere interaction during the Capitanian, Middle Permian ('end Guadalupian') extinction event". Palaeogeography, Palaeoclimatology, Palaeoecology . 441: 65–73. Bibcode:2016PPP...441...65J. doi:10.1016/j.palaeo.2015.06.009 . Retrieved 19 December 2022.
  68. Grasby, Stephen E.; Beauchamp, Benoit; Bond, David P. G.; Wignall, Paul B.; Sanei, Hamed (2016). "Mercury anomalies associated with three extinction events (Capitanian Crisis, Latest Permian Extinction and the Smithian/Spathian Extinction) in NW Pangea". Geological Magazine . 153 (2): 285–297. Bibcode:2016GeoM..153..285G. doi:10.1017/S0016756815000436. S2CID   85549730 . Retrieved 16 September 2022.
  69. Freydlin, Julie (16 July 2015). "Brachiopod die-off signaled mid-Permian mass extinction". Earth Magazine.
  70. Ling, Kunyue; Wen, Hanjie; Grasby, Stephen E.; Zhao, Haonan; Deng, Changzhou; Yin, Runsheng (5 February 2023). "The Emeishan large igneous province eruption triggered coastal perturbations and the Capitanian mass extinction: Insights from mercury in Permian bauxite beds". Chemical Geology . 617: 121243. Bibcode:2023ChGeo.61721243L. doi:10.1016/j.chemgeo.2022.121243. S2CID   254298090 . Retrieved 18 April 2023.
  71. Kaiho, Kunio; Grasby, Stephen E.; Chen, Zhong-Qiang (15 May 2023). "High-temperature combustion event spanning the Guadalupian−Lopingian boundary terminated by soil erosion". Palaeogeography, Palaeoclimatology, Palaeoecology . 618: 111518. Bibcode:2023PPP...61811518K. doi:10.1016/j.palaeo.2023.111518. S2CID   257675859 . Retrieved 20 April 2023.
  72. Jiang, Qiang; Jourdan, Fred; Olierook, Hugo K. H.; Merle, Renaud E.; Bourdet, Julien; Fougerouse, Denis; Godel, Belinda; Walker, Alex T. (25 July 2022). "Volume and rate of volcanic CO2 emissions governed the severity of past environmental crises". Proceedings of the National Academy of Sciences of the United States of America . 119 (31): e2202039119. Bibcode:2022PNAS..11902039J. doi: 10.1073/pnas.2202039119 . PMC   9351498 . PMID   35878029.
  73. Wang, Wen-qian; Zheng, Feifei; Zhang, Shuang; Cui, Ying; Zheng, Quan-feng; Zhang, Yi-chun; Chang, Dong-xun; Zhang, Hua; Xu, Yi-gang; Shen, Shu-zhong (15 January 2023). "Ecosystem responses of two Permian biocrises modulated by CO2 emission rates". Earth and Planetary Science Letters . 602: 117940. Bibcode:2023E&PSL.60217940W. doi:10.1016/j.epsl.2022.117940. S2CID   254660567 . Retrieved 30 January 2023.
  74. Retallack, Gregory J.; Jahren, A. Hope (1 October 2007). "Methane Release from Igneous Intrusion of Coal during Late Permian Extinction Events". The Journal of Geology . 116 (1): 1–20. doi:10.1086/524120. S2CID   46914712 . Retrieved 30 September 2022.
  75. Sheldon, Nathan D.; Chakrabarti, Ramananda; Retallack, Gregory J.; Smith, Roger M. H. (20 February 2014). "Contrasting geochemical signatures on land from the Middle and Late Permian extinction events". Sedimentology. 61 (6): 1812–1829. doi:10.1111/sed.12117. hdl:2027.42/108696. S2CID   129862176 . Retrieved 23 December 2022.
  76. Zhu, Jiang; Zhang, Zhaochong; Santosh, M.; Tan, Shucheng; Deng, Yinan; Xie, Qiuhong (1 July 2021). "Recycled carbon degassed from the Emeishan plume as the potential driver for the major end-Guadalupian carbon cycle perturbations". Geoscience Frontiers. 12 (4): 101140. Bibcode:2021GeoFr..1201140Z. doi:10.1016/j.gsf.2021.101140. ISSN   1674-9871 . Retrieved 12 January 2024 via Elsevier Science Direct.
  77. Ganino, Clément; Arndt, Nicholas T. (1 April 2009). "Climate changes caused by degassing of sediments during the emplacement of large igneous provinces". Geology . 37 (4): 323–326. Bibcode:2009Geo....37..323G. doi:10.1130/G25325A.1. ISSN   1943-2682 . Retrieved 12 January 2024 via GeoScienceWorld.
  78. Wignall, Paul B. (29 September 2015). "Extinction in the Shadows". The Worst of Times: How Life on Earth Survived Eighty Million Years of Mass Extinctions. Princeton: Princeton University Press. p. 38. ISBN   978-0691142098.
  79. 1 2 Bond, David P. G.; Wignall, Paul B.; Grasby, Stephen E. (30 August 2019). "The Capitanian (Guadalupian, Middle Permian) mass extinction in NW Pangea (Borup Fiord, Arctic Canada): A global crisis driven by volcanism and anoxia". Geological Society of America Bulletin . 132 (5–6): 931–942. doi:10.1130/B35281.1. S2CID   199104686 . Retrieved 29 November 2022.
  80. Song, Huyue; Algeo, Thomas J.; Song, Haijun; Tong, Jinnan; Wignall, Paul B.; Bond, David P. G.; Zheng, Wang; Chen, Xinming; Romaniello, Stephen J.; Wei, Hengye; Anbar, Ariel D. (15 May 2023). "Global oceanic anoxia linked with the Capitanian (Middle Permian) marine mass extinction". Earth and Planetary Science Letters . 610: 118128. Bibcode:2023E&PSL.61018128S. doi:10.1016/j.epsl.2023.118128. S2CID   257873150. Archived from the original on 22 April 2023. Retrieved 21 April 2023.{{cite journal}}: CS1 maint: bot: original URL status unknown (link)
  81. Zhang, Bolin; Wignall, Paul B.; Yao, Suping; Hu, Wenxuan; Liu, Biao (January 2021). "Collapsed upwelling and intensified euxinia in response to climate warming during the Capitanian (Middle Permian) mass extinction". Gondwana Research . 89: 31–46. Bibcode:2021GondR..89...31Z. doi:10.1016/j.gr.2020.09.003. S2CID   224981591 . Retrieved 30 September 2022.
  82. Zhang, Bolin; Yao, Suping; Hu, Wenxuan; Ding, Hai; Liu, Bao; Ren, Yongle (1 October 2019). "Development of a high-productivity and anoxic-euxinic condition during the late Guadalupian in the Lower Yangtze region: Implications for the mid-Capitanian extinction event". Palaeogeography, Palaeoclimatology, Palaeoecology . 531: 108630. Bibcode:2019PPP...53108630Z. doi:10.1016/j.palaeo.2018.01.021. S2CID   133916878 . Retrieved 7 November 2022.
  83. Saitoh, Masafumi; Isozaki, Yukio; Ueno, Yuichiro; Yoshida, Naohiro; Yao, Jianxin; Ji, Zhansheng (May 2013). "Middle–Upper Permian carbon isotope stratigraphy at Chaotian, South China: Pre-extinction multiple upwelling of oxygen-depleted water onto continental shelf". Journal of Asian Earth Sciences . 67–68: 51–62. Bibcode:2013JAESc..67...51S. doi:10.1016/j.jseaes.2013.02.009 . Retrieved 14 January 2023.
  84. 1 2 3 4 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. Bibcode:2002Geobi..35..287W. doi:10.1016/S0016-6995(02)00066-9 . Retrieved 13 January 2023.
  85. Chen, Fayao; Xue, Wuqiang; Yan, Jiaxin; Meng, Qi (19 April 2021). "The implications of the giant bivalve family Alatoconchidae for the end-Guadalupian (Middle Permian) extinction event". Geological Journal . 56 (2): 6073–6087. Bibcode:2021GeolJ..56.6073C. doi:10.1002/gj.4151. S2CID   234815123 . Retrieved 7 November 2022.
  86. Buatois, Luis A.; Borruel-Abadía, Violeta; De la Horra, Raúl; Galán-Abellán, Ana Belén; López-Gómez, José; Barrenechea, José F.; Arche, Alfredo (25 March 2021). "Impact of Permian mass extinctions on continental invertebrate infauna". Terra Nova . 33 (5): 455–464. Bibcode:2021TeNov..33..455B. doi:10.1111/ter.12530. S2CID   233616369 . Retrieved 23 December 2022.
  87. Jost, Adam B.; Mundil, Roland; He, Bin; Brown, Shaun T.; Altiner, Demir; Sun, Yadong; DePaolo, Donald J.; Payne, Jonathan L. (15 June 2014). "Constraining the cause of the end-Guadalupian extinction with coupled records of carbon and calcium isotopes". Earth and Planetary Science Letters . 396: 201–212. Bibcode:2014E&PSL.396..201J. doi:10.1016/j.epsl.2014.04.014 . Retrieved 12 January 2023.
  88. 1 2 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. S2CID   135404039 . Retrieved 18 March 2023.
  89. Day, Michael O.; Rubidge, Bruce S. (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 .
  90. Hallam, A.; Wignall, Paul B. (December 1999). "Mass extinctions and sea-level changes". Earth-Science Reviews . 48 (4): 217–250. Bibcode:1999ESRv...48..217H. doi:10.1016/S0012-8252(99)00055-0 . Retrieved 13 January 2023.
  91. Saitoh, Masamufi; Isozaki, Yukio; Yao, Jianxin; Ji, Zhansheng; Ueno, Yuichiro; Yoshida, Naohiro (June 2015). "The appearance of an oxygen-depleted condition on the Capitanian disphotic slope/basin in South China: Middle–Upper Permian stratigraphy at Chaotian in northern Sichuan". Global and Planetary Change . 105: 180–192. doi:10.1016/j.gloplacha.2012.01.002 . Retrieved 13 January 2023.
  92. Shu-Zhong, Shen; Shi, G. R. (8 April 2016). "Paleobiogeographical extinction patterns of Permian brachiopods in the Asian–western Pacific region". Paleobiology . 28 (4): 449–463. doi:10.1666/0094-8373(2002)028<0449:PEPOPB>2.0.CO;2. JSTOR   3595495. S2CID   35611701.