fbpx
Wikipedia

Permian–Triassic extinction event

The Permian–Triassic (P–T, P–Tr)[3][4] extinction event,[5] also known as the Latest Permian extinction event, the End-Permian Extinction[6] and colloquially as the Great Dying,[7] formed the boundary between the Permian and Triassic geologic periods, and with them the Paleozoic and Mesozoic eras respectively, approximately 251.9 million years ago.[8] It is the Earth's most severe known extinction event,[9][10] with the extinction of 57% of biological families, 83% of genera, 81% of marine species[11][12][13] and 70% of terrestrial vertebrate species.[14] It was the largest known mass extinction of insects.

CambrianOrdovicianSilurianDevonianCarboniferousPermianTriassicJurassicCretaceousPaleogeneNeogene
Marine extinction intensity during the Phanerozoic
%
Millions of years ago
CambrianOrdovicianSilurianDevonianCarboniferousPermianTriassicJurassicCretaceousPaleogeneNeogene
Plot of extinction intensity (percentage of marine genera that are present in each interval of time but do not exist in the following interval) vs time in the past.[1] Geological periods are annotated (by abbreviation and colour) above. The Permian–Triassic extinction event is the most significant event for marine genera, with just over 50% (according to this source) perishing. (source and image info)
Permian–Triassic boundary at Frazer Beach in New South Wales, with the End Permian extinction event located just above the coal layer.[2]

There is evidence for one to three distinct pulses, or phases, of extinction.[15][14][16][17][18]

The scientific consensus is that the main cause of extinction was the large amount of carbon dioxide emitted by the volcanic eruptions that created the Siberian Traps, which elevated global temperatures, and in the oceans led to widespread anoxia and acidification.[19] Proposed contributing factors include: the emission of much additional carbon dioxide from the thermal decomposition of hydrocarbon deposits, including oil and coal, triggered by the eruptions; and emissions of methane by novel methanogenic microorganisms, perhaps nourished by minerals dispersed in the eruptions.[20][21]

The speed of recovery from the extinction is disputed. Some scientists estimate that it took 10 million years (until the Middle Triassic), due both to the severity of the extinction and because grim conditions returned periodically over the course of the Early Triassic,[22][23][24] causing further extinction events, such as the Smithian-Spathian boundary extinction.[10][25] However, studies in Bear Lake County, near Paris, Idaho,[26] and nearby sites in Idaho and Nevada[27] showed a relatively quick rebound in a localized Early Triassic marine ecosystem, taking around 3 million years to recover, while an unusually diverse and complex ichnobiota is known from Italy less than a million years after the end-Permian extinction,[28] suggesting that the impact of the extinction may have been felt less severely in some areas than in others. Differential environmental stress and instability following the extinction event has been theorised as an explanation for the disparity in the pace of biotic recovery between different regions and environments.[29]

Dating

Previously, it was thought that rock sequences spanning the Permian–Triassic boundary were too few and contained too many gaps for scientists to reliably determine its details.[33] However, it is now possible to date the extinction with millennial precision. U–Pb zircon dates from five volcanic ash beds from the Global Stratotype Section and Point for the Permian–Triassic boundary at Meishan, China, establish a high-resolution age model for the extinction – allowing exploration of the links between global environmental perturbation, carbon cycle disruption, mass extinction, and recovery at millennial timescales.

The extinction occurred between 251.941 ± 0.037 and 251.880 ± 0.031 million years ago, a duration of 60 ± 48 thousand years.[34] A large (approximately 0.9%), abrupt global decrease in the ratio of the stable isotope carbon-13 to that of carbon-12 coincides with this extinction,[31][35][36][37][38] and is sometimes used to identify the Permian–Triassic boundary in rocks that are unsuitable for radiometric dating.[37]

Further evidence for environmental change around the P–Tr boundary suggests an 8 °C (14 °F) rise in temperature,[31] and an increase in CO
2
levels by 2,000 ppm (for comparison, the concentration immediately before the Industrial Revolution was 280 ppm,[31] and the amount today is about 415 ppm[39]). There is also evidence of increased ultraviolet radiation reaching the earth, causing the mutation of plant spores.[31][40]

It has been suggested that the Permian–Triassic boundary is associated with a sharp increase in the abundance of marine and terrestrial fungi, caused by the sharp increase in the amount of dead plants and animals fed upon by the fungi.[41] For a while this "fungal spike" was used by some paleontologists to identify the Permian–Triassic boundary in rocks that are unsuitable for radiometric dating or have a lack of suitable index fossils, but even the proposers of the fungal spike hypothesis pointed out that "fungal spikes" may have been a repeating phenomenon created by the post-extinction ecosystem during the earliest Triassic.[41] The very idea of a fungal spike has been criticized on several grounds, including: Reduviasporonites, the most common supposed fungal spore, may be a fossilized alga;[31][42] the spike did not appear worldwide;[43][44][45] and in many places it did not fall on the Permian–Triassic boundary.[46] The reduviasporonites may even represent a transition to a lake-dominated Triassic world rather than an earliest Triassic zone of death and decay in some terrestrial fossil beds.[47] Newer chemical evidence agrees better with a fungal origin for Reduviasporonites, diluting these critiques.[48][49]

Uncertainty exists regarding the duration of the overall extinction and about the timing and duration of various groups' extinctions within the greater process. Some evidence suggests that there were multiple extinction pulses[14] or that the extinction was long and spread out over a few million years, with a sharp peak in the last million years of the Permian.[46][50] Statistical analyses of some highly fossiliferous strata in Meishan, Zhejiang Province in southeastern China, suggest that the main extinction was clustered around one peak.[16] Recent research shows that different groups became extinct at different times; for example, while difficult to date absolutely, ostracod and brachiopod extinctions were separated by around 670,000 to 1.17 million years.[51] In a well-preserved sequence in east Greenland, the decline of animal life is concentrated in a period approximately 10,000 to 60,000 years long, with plants taking an additional several hundred thousand years to show the full impact of the event.[52] Palaeoenvironmental analysis of Lopingian strata in the Bowen Basin of Queensland indicates numerous intermittent periods of environmental stress from the middle to late Lopingian leading up to the end-Permian extinction proper.[53]

An older theory, still supported in some recent papers,[14][54] is that there were two major extinction pulses 9.4 million years apart, separated by a period of extinctions well above the background level, and that the final extinction killed off only about 80% of marine species alive at that time while the other losses occurred during the first pulse or the interval between pulses. According to this theory one of these extinction pulses occurred at the end of the Guadalupian epoch of the Permian.[14][55] For example, all dinocephalian genera died out at the end of the Guadalupian,[54] as did the Verbeekinidae, a family of large-size fusuline foraminifera.[56] The impact of the end-Guadalupian extinction on marine organisms appears to have varied between locations and between taxonomic groups – brachiopods and corals had severe losses.[57][58]

Studies of the timing and causes of the Permian-Triassic extinction are complicated by the often-overlooked Capitanian extinction (also called the Guadalupian extinction), just one of perhaps two mass extinctions in the late Permian that closely preceded the Permian-Triassic event. In short, when the Permian-Triassic starts it is difficult to know whether the end-Capitanian had finished, depending on the factor considered.[1][59] Some of the extinctions dated to the Permian-Triassic boundary have recently been redated to the end-Capitanian. Further, it is unclear whether some species who survived the prior extinction(s) had recovered well enough for their final demise in the Permian-Triassic event to be considered separate from Capitanian event. A minority point of view considers the sequence of environmental disasters to have effectively constituted a single, prolonged extinction event, perhaps depending on which species is considered.

Extinction patterns

Marine extinctions Genera extinct Notes
Arthropoda
Eurypterids 100% May have become extinct shortly before the P–Tr boundary
Ostracods 59%  
Trilobites 100% In decline since the Devonian; only 2 genera living before the extinction
Brachiopoda
Brachiopods 96% Orthids and productids died out
Bryozoa
Bryozoans 79% Fenestrates, trepostomes, and cryptostomes died out
Chordata
Acanthodians 100% In decline since the Devonian, with only one living family
Cnidaria
Anthozoans 96% Tabulate and rugose corals died out
Echinodermata
Blastoids 100% May have become extinct shortly before the P–Tr boundary
Crinoids 98% Inadunates and camerates died out
Mollusca
Ammonites 97% Goniatites died out
Bivalves 59%  
Gastropods 98%  
Retaria
Foraminiferans 97% Fusulinids died out, but were almost extinct before the catastrophe
Radiolarians 99%[60]

Marine organisms

Marine invertebrates suffered the greatest losses during the P–Tr extinction. Evidence of this was found in samples from south China sections at the P–Tr boundary. Here, 286 out of 329 marine invertebrate genera disappear within the final two sedimentary zones containing conodonts from the Permian.[16] The decrease in diversity was probably caused by a sharp increase in extinctions, rather than a decrease in speciation.[61]

The extinction primarily affected organisms with calcium carbonate skeletons, especially those reliant on stable CO2 levels to produce their skeletons.[62] These organisms were susceptible to the effects of the ocean acidification that resulted from increased atmospheric CO2.

Among benthic organisms the extinction event multiplied background extinction rates, and therefore caused maximum species loss to taxa that had a high background extinction rate (by implication, taxa with a high turnover).[63][64] The extinction rate of marine organisms was catastrophic.[16][33][65][52] Bioturbators were extremely severely affected, as evidenced by the loss of the sedimentary mixed layer in many marine facies during the end-Permian extinction.[66]

Surviving marine invertebrate groups included articulate brachiopods (those with a hinge),[67] which had undergone a slow decline in numbers since the P–Tr extinction; the Ceratitida order of ammonites;[68] and crinoids ("sea lilies"),[68] which very nearly became extinct but later became abundant and diverse. The groups with the highest survival rates generally had active control of circulation, elaborate gas exchange mechanisms, and light calcification; more heavily calcified organisms with simpler breathing apparatuses suffered the greatest loss of species diversity.[69][70] In the case of the brachiopods, at least, surviving taxa were generally small, rare members of a formerly diverse community.[71]

The ammonoids, which had been in a long-term decline for the 30 million years since the Roadian (middle Permian), suffered a selective extinction pulse 10 million years before the main event, at the end of the Capitanian stage. In this preliminary extinction, which greatly reduced disparity, or the range of different ecological guilds, environmental factors were apparently responsible. Diversity and disparity fell further until the P–Tr boundary; the extinction here (P–Tr) was non-selective, consistent with a catastrophic initiator. During the Triassic, diversity rose rapidly, but disparity remained low.[72] The range of morphospace occupied by the ammonoids, that is, their range of possible forms, shapes or structures, became more restricted as the Permian progressed. A few million years into the Triassic, the original range of ammonoid structures was once again reoccupied, but the parameters were now shared differently among clades.[73]

The Lilliput effect, a term used to describe the phenomenon of dwarfing of species during an immediately following a mass extinction event, has been observed across the Permian-Triassic boundary,[74][75] notably occurring in foraminifera,[76] brachiopods,[77][78] bivalves,[79][80][81] and ostracods.[82] It remains debated whether the Lilliput effect took hold among gastropods.[83][84][85]

Terrestrial invertebrates

The Permian had great diversity in insect and other invertebrate species, including the largest insects ever to have existed. The end-Permian is the largest known mass extinction of insects;[86] according to some sources, it may well be the only mass extinction to significantly affect insect diversity.[87][88] Eight or nine insect orders became extinct and ten more were greatly reduced in diversity. Palaeodictyopteroids (insects with piercing and sucking mouthparts) began to decline during the mid-Permian; these extinctions have been linked to a change in flora. The greatest decline occurred in the Late Permian and was probably not directly caused by weather-related floral transitions.[33]

Terrestrial plants

The geological record of terrestrial plants is sparse and based mostly on pollen and spore studies. Plants are relatively immune to mass extinction, with the impact of all the major mass extinctions "insignificant" at a family level.[31][dubious ] Even the reduction observed in species diversity (of 50%) may be mostly due to taphonomic processes.[89][31] However, a massive rearrangement of ecosystems does occur, with plant abundances and distributions changing profoundly and all the forests virtually disappearing.[90][31] The dominant floral groups changed, with many groups of land plants entering abrupt decline, such as Cordaites (gymnosperms) and Glossopteris (seed ferns);[91][92] the Palaeozoic flora scarcely survived this extinction.[93]

The Glossopteris-dominated flora that characterised high-latitude Gondwana collapsed in Australia around 370,000 years before the Permian-Triassic boundary, with this flora's collapse being less constrained in western Gondwana but still likely occurring a few hundred thousand years before the boundary.[94]

Palynological or pollen studies from East Greenland of sedimentary rock strata laid down during the extinction period indicate dense gymnosperm woodlands before the event. At the same time that marine invertebrate macrofauna declined, these large woodlands died out and were followed by a rise in diversity of smaller herbaceous plants including Lycopodiophyta, both Selaginellales and Isoetales.[45]

The Cordaites flora, which dominated the Angaran floristic realm corresponding to Siberia, collapsed over the course of the extinction.[95] In the Kuznetsk Basin, the aridity-induced extinction of the regions's humid-adapted forest flora dominated by cordaitaleans occurred approximately 252.76 Ma, around 820,000 years before the end-Permian extinction in South China, suggesting that the end-Permian biotic catastrophe may have started earlier on land and that the ecological crisis may have been more gradual and asynchronous on land compared to its more abrupt onset in the marine realm.[96]

In South China, the subtropical Cathaysian gigantopterid dominated rainforests abruptly collapsed.[97][98][95] The floral extinction in South China is associated with bacterial blooms in soil and nearby lacustrine ecosystems, with soil erosion resulting from the die-off of plants being their likely cause.[99]

Terrestrial vertebrates

There is enough evidence to indicate that over two thirds of terrestrial labyrinthodont amphibians, sauropsid ("reptile") and therapsid ("proto-mammal") taxa became extinct. Large herbivores suffered the heaviest losses.

All Permian anapsid reptiles died out except the procolophonids (although testudines have morphologically-anapsid skulls, they are now thought to have separately evolved from diapsid ancestors). Pelycosaurs died out before the end of the Permian. Too few Permian diapsid fossils have been found to support any conclusion about the effect of the Permian extinction on diapsids (the "reptile" group from which lizards, snakes, crocodilians, and dinosaurs (including birds) evolved).[100][9]

The groups that survived suffered extremely heavy losses of species and some terrestrial vertebrate groups very nearly became extinct at the end of the Permian. Some of the surviving groups did not persist for long past this period, but others that barely survived went on to produce diverse and long-lasting lineages. However, it took 30 million years for the terrestrial vertebrate fauna to fully recover both numerically and ecologically.[101]

It is difficult to analyze extinction and survival rates of land organisms in detail because few terrestrial fossil beds span the Permian–Triassic boundary. The best-known record of vertebrate changes across the Permian–Triassic boundary occurs in the Karoo Supergroup of South Africa, but statistical analyses have so far not produced clear conclusions.[102]

Biotic recovery

In the wake of the extinction event, the ecological structure of present-day biosphere evolved from the stock of surviving taxa. In the sea, the "Modern Evolutionary Fauna" became dominant over elements of the "Palaeozoic Evolutionary Fauna".[103] Typical taxa of shelly benthic faunas were now bivalves, snails, sea urchins and Malacostraca, whereas bony fishes[104] and marine reptiles[105] diversified in the pelagic zone. On land, dinosaurs and mammals arose in the course of the Triassic. The profound change in the taxonomic composition was partly a result of the selectivity of the extinction event, which affected some taxa (e.g., brachiopods) more severely than others (e.g., bivalves).[106][107] However, recovery was also differential between taxa. Some survivors became extinct some million years after the extinction event without having rediversified (dead clade walking,[108] e.g. the snail family Bellerophontidae),[109] whereas others rose to dominance over geologic times (e.g., bivalves).[110][111]

Marine ecosystems

 
Shell bed with the bivalve Claraia clarai, a common early Triassic disaster taxon.

Marine post-extinction faunas were mostly species-poor and dominated by few disaster species such as the bivalves Claraia and Unionites. Seafloor communities maintained a comparatively low diversity until the end of the Early Triassic, approximately 4 million years after the extinction event.[112] This slow recovery stands in remarkable contrast with the quick recovery seen in nektonic organisms such as ammonoids, which exceeded pre-extinction diversities already two million years after the crisis.[113] The relative delay in the recovery of benthic organisms has been attributed to widespread anoxia,[114] but high abundances of benthic species contradict this explanation.[115] More recent work suggests that the pace of recovery was intrinsically driven by the intensity of competition among species, which drives rates of niche differentiation and speciation.[116] Accordingly, low levels of interspecific competition in seafloor communities that are dominated by primary consumers correspond to slow rates of diversification and high levels of interspecific competition among nektonic secondary and tertiary consumers to high diversification rates. Other studies point to the recurrent episodes of extremely hot climatic conditions during the Early Triassic for the delayed recovery of oceanic life.[117] A 2019 study attributed the dissimilarity of recovery times between different ecological communities to differences in local environmental stress during the biotic recovery interval, with regions experiencing persistent environmental stress post-extinction recovering more slowly.[29] Whereas most marine communities were fully recovered by the Middle Triassic,[118][119] global marine diversity reached pre-extinction values no earlier than the Middle Jurassic, approximately 75 million years after the extinction event.[120]

 
Sessile filter feeders like this Carboniferous crinoid, the mushroom crinoid (Agaricocrinus americanus), were significantly less abundant after the P–Tr extinction.

Prior to the extinction, about two-thirds of marine animals were sessile and attached to the seafloor. During the Mesozoic, only about half of the marine animals were sessile while the rest were free-living. Analysis of marine fossils from the period indicated a decrease in the abundance of sessile epifaunal suspension feeders such as brachiopods and sea lilies and an increase in more complex mobile species such as snails, sea urchins and crabs.[121]

Before the Permian mass extinction event, both complex and simple marine ecosystems were equally common. After the recovery from the mass extinction, the complex communities outnumbered the simple communities by nearly three to one,[121] and the increase in predation pressure led to the Mesozoic Marine Revolution.

Bivalves were fairly rare before the P–Tr extinction but became numerous and diverse in the Triassic, taking over niches that were filled primarily by brachiopods before the mass extinction event,[122] and one group, the rudist clams, became the Mesozoic's main reef-builders. Some researchers think the change was attributable not only to the end-Permian extinction but also the ecological restructuring that began as a result of the Capitanian extinction.[123]

Linguliform brachiopods were commonplace immediately after the extinction event, their abundance having been essentially unaffected by the crisis. Adaptations for oxygen-poor and warm environments, such as increased lophophoral cavity surface, shell width/length ratio, and shell miniaturisation, are observed in post-extinction linguliforms.[124]

Crinoids ("sea lilies") suffered a selective extinction, resulting in a decrease in the variety of their forms.[125] Their ensuing adaptive radiation was brisk, and resulted in forms possessing flexible arms becoming widespread; motility, predominantly a response to predation pressure, also became far more prevalent.[126]

Microbial-metazoan reefs dominated surviving communities in the immediate wake of the mass extinction. Metazoan-built reefs reemerged during the Olenekian, mainly being composed of sponge biostrome and bivalve builups. "Tubiphytes"-dominated reefs appeared at the end of the Olenekian, representing the earliest platform-margin reefs of the Triassic, though they did not become abundant until the late Anisian, when reefs' species richness increased. The first scleractinian corals appear in the late Anisian as well, although they would not become the dominant reef builders until the end of the Triassic period.[127]

Terrestrial plants

The proto-recovery of terrestrial floras took place from a few tens of thousands of years after the end-Permian extinction to around 350,000 years after it, with the exact timeline varying by region.[128] Dominant gymnosperm genera were replaced post-boundary by lycophytes – extant lycophytes are recolonizers of disturbed areas.[129] The particular post-extinction dominance of lycophytes, which were well adapted for coastal environments, can be explained in part by global marine transgressions during the Early Triassic.[95] The worldwide recovery of gymnosperm forests took approximately 4–5 million years.[31] However, this trend of prolonged lycophyte dominance during the Early Triassic was not universal, as evidenced by the much more rapid recovery of gymnosperms in certain regions,[130] and floral recovery likely didn't follow a congruent, globally universal trend but instead varied by region according to local environmental conditions.[98]

In East Greenland, lycophytes replaced gymnosperms as the dominant plants. Later, other groups of gymnosperms again become dominant but again suffered major die-offs. These cyclical flora shifts occurred a few times over the course of the extinction period and afterward. These fluctuations of the dominant flora between woody and herbaceous taxa indicate chronic environmental stress resulting in a loss of most large woodland plant species. The successions and extinctions of plant communities do not coincide with the shift in δ13C values but occurred many years after.[45]

In what is now the Barents Sea of the coast of Norway, the post-extinction fauna is dominated by pteridophytes and lycopods, which were suited for primary succession and recolonisation of devastated areas, although gymnosperms made a rapid recovery, with the lycopod dominated flora not persisting across most of the Early Triassic as postulated in other regions.[130]

In Europe and North China, the interval of recovery was dominated by the lycopsid Pleuromeia, an opportunistic pioneer plant that filled ecological vacancies until other plants were able to expand out of refugia and recolonise the land. Conifers became common by the early Anisian, while pteridosperms and cycadophytes only fully recovered by the late Anisian.[131]

In southwestern Gondwana, the post-extinction flora was dominated by bennettitaleans and cycads, with memers of Peltaspermales, Ginkgoales, and Umkomasiales being less common constituents of this flora. Around the Induan-Olenekian boundary, as palaeocommunities recovered, a new Dicroidium flora was established, in which Umkomasiales continued to be prominent and in which Equisetales and Cycadales were subordinate forms. The Dicroidium flora further diversified in the Anisian to its peak, wherein Umkomasiales and Ginkgoales constituted most of the tree canopy and Peltaspermales, Petriellales, Cycadales, Umkomasiales, Gnetales, Equisetales, and Dipteridaceae dominated the understory.[94]

The gigantopterid-dominated forests of South China were replaced by low-lying herbaceous vegetation dominated by the isoetalean Tomiostrobus following the collapse of the former. In contrast to the highly biodiverse gigantopterid rainforests, the post-extinction landscape of South China was near-barren and had vastly lower diversity.[98]

Coal gap

No coal deposits are known from the Early Triassic, and those in the Middle Triassic are thin and low-grade. This "coal gap" has been explained in many ways. It has been suggested that new, more aggressive fungi, insects, and vertebrates evolved and killed vast numbers of trees. These decomposers themselves suffered heavy losses of species during the extinction and are not considered a likely cause of the coal gap. It could simply be that all coal-forming plants were rendered extinct by the P–Tr extinction and that it took 10 million years for a new suite of plants to adapt to the moist, acid conditions of peat bogs.[32] Abiotic factors (factors not caused by organisms), such as decreased rainfall or increased input of clastic sediments, may also be to blame.[31]

On the other hand, the lack of coal may simply reflect the scarcity of all known sediments from the Early Triassic. Coal-producing ecosystems, rather than disappearing, may have moved to areas where we have no sedimentary record for the Early Triassic.[31] For example, in eastern Australia a cold climate had been the norm for a long period, with a peat mire ecosystem adapted to these conditions. Approximately 95% of these peat-producing plants went locally extinct at the P–Tr boundary;[132] coal deposits in Australia and Antarctica disappear significantly before the P–Tr boundary.[31]

Terrestrial vertebrates

 
Lystrosaurus was by far the most abundant early Triassic land vertebrate.

Lystrosaurus, a pig-sized herbivorous dicynodont therapsid, constituted as much as 90% of some earliest Triassic land vertebrate fauna. Smaller carnivorous cynodont therapsids also survived, including the ancestors of mammals. In the Karoo region of southern Africa, the therocephalians Tetracynodon, Moschorhinus and Ictidosuchoides survived, but do not appear to have been abundant in the Triassic.[133] In North China, tetrapod body and ichnofossils are extremely rare in Induan facies, but become more abundant in the Olenekian and Anisian, showing a biotic recovery of tetrapods synchronous with the decreasing aridity during the Olenekian and Anisian.[134][135]

Archosaurs (which included the ancestors of dinosaurs and crocodilians) were initially rarer than therapsids, but they began to displace therapsids in the mid-Triassic. In the mid to late Triassic, the dinosaurs evolved from one group of archosaurs, and went on to dominate terrestrial ecosystems during the Jurassic and Cretaceous.[136] This "Triassic Takeover" may have contributed to the evolution of mammals by forcing the surviving therapsids and their mammaliform successors to live as small, mainly nocturnal insectivores; nocturnal life probably forced at least the mammaliforms to develop fur, better hearing and higher metabolic rates,[137] while losing part of the differential color-sensitive retinal receptors reptilians and birds preserved. The archosaur dominance would end again due to the Cretaceous–Paleogene extinction event, after which both birds (only extant dinosaurs) and mammals (only extant synapsids) would diversify and share the world.

Some temnospondyl amphibians made a relatively quick recovery, in spite of nearly becoming extinct. Mastodonsaurus and trematosaurians were the main aquatic and semiaquatic predators during most of the Triassic, some preying on tetrapods and others on fish.[138]

Land vertebrates took an unusually long time to recover from the P–Tr extinction; Palaeontologist Michael Benton estimated the recovery was not complete until 30 million years after the extinction, i.e. not until the Late Triassic, when the first dinosaurs had risen from bipedal archosaurian ancestors and the first mammals from small cynodont ancestors.[12]

Coprolitic evidence indicates that freshwater food webs had recovered by the early Ladinian, with a lacustrine coprolite assemblage from the Ordos Basin of China providing evidence of a trophically multileveled ecosystem containing at least six different trophic levels. The highest trophic levels were filled by vertebrate predators.[139]

Terrestrial invertebrates

Most fossil insect groups found after the Permian–Triassic boundary differ significantly from those before: Of Paleozoic insect groups, only the Glosselytrodea, Miomoptera, and Protorthoptera have been discovered in deposits from after the extinction. The caloneurodeans, monurans, paleodictyopteroids, protelytropterans, and protodonates became extinct by the end of the Permian. Though Triassic insects are very different from those of the Permian, a gap in the insect fossil record spans approximately 15 million years from the late Permian to early Triassic. In well-documented Late Triassic deposits, fossils overwhelmingly consist of modern fossil insect groups.[87]

Hypotheses about cause

Pinpointing the exact causes of the Permian–Triassic extinction event is difficult, mostly because it occurred over 250 million years ago, and since then much of the evidence that would have pointed to the cause has been destroyed or is concealed deep within the Earth under many layers of rock. The sea floor is completely recycled over around 200 million years by the ongoing process of plate tectonics and seafloor spreading, leaving no useful indications beneath the ocean.

Yet, scientists have gathered significant evidence for causes, and several mechanisms have been proposed. The proposals include both catastrophic and gradual processes (similar to those theorized for the Cretaceous–Paleogene extinction event).

  • The catastrophic group includes one or more large bolide impact events, increased volcanism, and sudden release of methane from the seafloor, either due to dissociation of methane hydrate deposits or metabolism of organic carbon deposits by methanogenic microbes.
  • The gradual group includes sea level change, increasing hypoxia, and increasing aridity.

Any hypothesis about the cause must explain the selectivity of the event, which affected organisms with calcium carbonate skeletons most severely; the long period (4 to 6 million years) before recovery started, and the minimal extent of biological mineralization (despite inorganic carbonates being deposited) once the recovery began.[62]

Volcanism

Siberian Traps

The final stages of the Permian had two flood basalt events. A smaller one, the Emeishan Traps in China, occurred at the same time as the end-Guadalupian extinction pulse, in an area close to the equator at the time.[140][141] The flood basalt eruptions that produced the Siberian Traps constituted one of the largest known volcanic events on Earth and covered over 2,000,000 square kilometres (770,000 sq mi) with lava.[142][143][144] Such a vast aerial extent of the flood basalts may have contributed to their exceptionally catastrophic impact.[145] The date of the Siberian Traps eruptions and the extinction event are in good agreement.[34][146] The timeline of the extinction event strongly indicates it was caused by events in the large igneous province of the Siberian Traps.[147][148][149][150] Carbon dioxide levels prior to and after the eruptions are poorly constrained, but may have jumped from between 500 and 4,000 PPM prior to the extinction event to around 8,000 PPM after the extinction.[151] As carbon dioxide levels shot up, extreme temperature rise would have followed.[152] In a 2020 paper, scientists reconstructed the mechanisms that led to the extinction event in a biogeochemical model, showed the consequences of the greenhouse effect on the marine environment, and concluded that the mass extinction can be traced back to volcanic CO2 emissions.[153][8] Further evidence – based on paired coronene-mercury spikes – for a volcanic combustion cause of the mass extinction was published in 2020.[154][20]

The Siberian Traps had unusual features that made them even more dangerous. The Siberian Traps eruptions released sulphur-rich volatiles that caused dust clouds and the formation of acid aerosols, which would have blocked out sunlight and thus disrupted photosynthesis both on land and in the photic zone of the ocean, causing food chains to collapse. These volcanic outbursts of sulphur also induced brief but severe global cooling, leading to glacio-eustatic sea level fall.[155] The eruptions may also have caused acid rain as the aerosols washed out of the atmosphere. That may have killed land plants and mollusks and planktonic organisms which had calcium carbonate shells. Pure flood basalts produce fluid, low-viscosity lava, and do not hurl debris into the atmosphere. It appears, however, that 20% of the output of the Siberian Traps eruptions was pyroclastic (consisted of ash and other debris thrown high into the atmosphere), increasing the short-term cooling effect.[156] When all of the dust and ash clouds and aerosols washed out of the atmosphere, the excess carbon dioxide emitted by the Siberian Traps would have remained and global warming would have proceeded without any mitigating effects.[152]

The Siberian Traps are underlain by thick sequences of Early-Mid Paleozoic aged carbonate and evaporite deposits, as well as Carboniferous-Permian aged coal bearing clastic rocks. When heated, such as by igneous intrusions, these rocks are capable of emitting large amounts of greenhouse and toxic gases. The unique setting of the Siberian Traps over these deposits is likely the reason for the severity of the extinction.[157][158] The basalt lava erupted or intruded into carbonate rocks and into sediments that were in the process of forming large coal beds, both of which would have emitted large amounts of carbon dioxide, leading to stronger global warming after the dust and aerosols settled.[152] The timing of the change of the Siberian Traps from flood basalt dominated emplacement to sill dominated emplacement, the latter of which would have liberated the largest amounts of trapped hydrocarbon deposits, coincides with the onset of the main phase of the mass extinction[159][160] and is linked to a major negative δ13C excursion.[161] A 2011 study led by Stephen E. Grasby reported evidence that volcanism caused massive coal beds to ignite, possibly releasing more than 3 trillion tons of carbon. The team found ash deposits in deep rock layers near what is now the Buchanan Lake Formation. According to their article, "coal ash dispersed by the explosive Siberian Trap eruption would be expected to have an associated release of toxic elements in impacted water bodies where fly ash slurries developed. ... Mafic megascale eruptions are long-lived events that would allow significant build-up of global ash clouds."[162][163] In a statement, Grasby said, "In addition to these volcanoes causing fires through coal, the ash it spewed was highly toxic and was released in the land and water, potentially contributing to the worst extinction event in earth history."[164] A 2013 study led by Q.Y. Yang reported that the total amounts of important volatiles emitted from the Siberian Traps consisted of 8.5 × 107 Tg CO2, 4.4 × 106 Tg CO, 7.0 × 106 Tg H2S, and 6.8 × 107 Tg SO2. The data support a popular notion that the end-Permian mass extinction on the Earth was caused by the emission of enormous amounts of volatiles from the Siberian Traps into the atmosphere.[165]

In addition, these volcanic eruptions released significant quantities of toxic mercury into the atmosphere and ocean, further contributing to large scale die-offs of terrestrial and marine life.[166][167][168] A series of surges in mercury emissions raised environmental mercury concentrations to levels orders of magnitude above normal background levels and caused intervals of extreme environmental toxicity, each lasting for over a thousand years, in both terrestrial and marine ecosystems.[169] Nickel aerosols released by volcanic activity further contributed to metal poisoning.[170] Cobalt and arsenic emissions from the Siberian Traps caused further still environmental stress.[166]

Choiyoi Silicic Large Igneous Province

A second flood basalt event that emplaced what is now known as the Choiyoi Silicic Large Igneous Province in southwestern Gondwana between around 286 Ma and 247 Ma has also been suggested as a possible extinction mechanism.[94] Being about 1,300,000 cubic kilometres in volume[171] and 1,680,000 square kilometres in area, this flood basalt event was approximately 40% the size of the Siberian Traps and thus may have been a significant additional factor explaining the severity of the end-Permian extinction.[94]

Methane clathrate gasification

Methane clathrates, also known as methane hydrates, consist of methane molecules trapped in cages of water molecules. The methane, produced by methanogens (microscopic single-celled organisms), has a 13C 12C ratio about 6.0% below normal (δ13C −6.0%). At the right combination of pressure and temperature, the methane is trapped in clathrates fairly close to the surface of permafrost and, in much larger quantities, on continental shelves and the deeper seabed close to them. Oceanic methane hydrates are usually found buried in sediments where the seawater is at least 300 m (980 ft) deep. They can be found up to about 2,000 m (6,600 ft) below the sea floor, but usually only about 1,100 m (3,600 ft) below the sea floor.[172]

The release of methane from the clathrates has been considered as a cause because scientists have found worldwide evidence of a swift decrease of about 1% in the 13C 12C isotope ratio in carbonate rocks from the end-Permian.[52][173] This is the first, largest, and most rapid of a series of negative and positive excursions (decreases and increases in 13C 12C ratio) that continues until the isotope ratio abruptly stabilised in the middle Triassic, followed soon afterwards by the recovery of calcifying life forms (organisms that use calcium carbonate to build hard parts such as shells).[69] While a variety of factors may have contributed to this drop in the 13C 12C ratio, a 2002 review found most of them to be insufficient to account fully for the observed amount:[174]

  • Gases from volcanic eruptions have a 13C 12C ratio about 0.5 to 0.8% below standard (δ13C about −0.5 to −0.8%), but an assessment made in 1995 concluded that the amount required to produce a reduction of about 1.0% worldwide requires eruptions greater by orders of magnitude than any for which evidence has been found.[175] (However, this analysis addressed only CO2 produced by the magma itself, not from interactions with carbon bearing sediments, as later proposed.)
  • A reduction in organic activity would extract 12C more slowly from the environment and leave more of it to be incorporated into sediments, thus reducing the 13C 12C ratio. Biochemical processes preferentially use the lighter isotopes since chemical reactions are ultimately driven by electromagnetic forces between atoms and lighter isotopes respond more quickly to these forces, but a study of a smaller drop of 0.3 to 0.4% in 13C 12C (δ13C −3 to −4 ‰) at the Paleocene-Eocene Thermal Maximum (PETM) concluded that even transferring all the organic carbon (in organisms, soils, and dissolved in the ocean) into sediments would be insufficient: Even such a large burial of material rich in 12C would not have produced the 'smaller' drop in the 13C 12C ratio of the rocks around the PETM.[175]
  • Buried sedimentary organic matter has a 13C 12C ratio 2.0 to 2.5% below normal (δ13C −2.0 to −2.5%). Theoretically, if the sea level fell sharply, shallow marine sediments would be exposed to oxidation. But 6,500–8,400 gigatons (1 gigaton = 109 metric tons) of organic carbon would have to be oxidized and returned to the ocean-atmosphere system within less than a few hundred thousand years to reduce the 13C 12C ratio by 1.0%, which is not thought to be a realistic possibility.[33] Moreover, sea levels were rising rather than falling at the time of the extinction.[152]
  • Rather than a sudden decline in sea level, intermittent periods of ocean-bottom hyperoxia and anoxia (high-oxygen and low- or zero-oxygen conditions) may have caused the 13C 12C ratio fluctuations in the Early Triassic;[69] and global anoxia may have been responsible for the end-Permian blip. The continents of the end-Permian and early Triassic were more clustered in the tropics than they are now, and large tropical rivers would have dumped sediment into smaller, partially enclosed ocean basins at low latitudes. Such conditions favor oxic and anoxic episodes; oxic/anoxic conditions would result in a rapid release/burial, respectively, of large amounts of organic carbon, which has a low 13C 12C ratio because biochemical processes use the lighter isotopes more.[176] That or another organic-based reason may have been responsible for both that and a late Proterozoic/Cambrian pattern of fluctuating 13C 12C ratios.[69]

Other hypotheses include mass oceanic poisoning, releasing vast amounts of CO2,[177] and a long-term reorganisation of the global carbon cycle.[174]

Prior to consideration of the inclusion of roasting carbonate sediments by volcanism, the only proposed mechanism sufficient to cause a global 1% reduction in the 13C 12C ratio was the release of methane from methane clathrates.[33] Carbon-cycle models confirm that it would have had enough effect to produce the observed reduction.[174][177] It was also suggested that a large-scale release of methane and other greenhouse gases from the ocean into the atmosphere was connected to the anoxic events and euxinic (i.e. sulfidic) events at the time, with the exact mechanism compared to the 1986 Lake Nyos disaster [178]

The area covered by lava from the Siberian Traps eruptions is about twice as large as was originally thought, and most of the additional area was shallow sea at the time. The seabed probably contained methane hydrate deposits, and the lava caused the deposits to dissociate, releasing vast quantities of methane.[179] A vast release of methane might cause significant global warming since methane is a very powerful greenhouse gas. Strong evidence suggests the global temperatures increased by about 6 °C (10.8 °F) near the equator and therefore by more at higher latitudes: a sharp decrease in oxygen isotope ratios (18O 16O);[180] the extinction of Glossopteris flora (Glossopteris and plants that grew in the same areas), which needed a cold climate, with its replacement by floras typical of lower paleolatitudes.[181]

However, the pattern of isotope shifts expected to result from a massive release of methane does not match the patterns seen throughout the Early Triassic. Not only would such a cause require the release of five times as much methane as postulated for the PETM,[69] but would it also have to be reburied at an unrealistically high rate to account for the rapid increases in the 13C 12C ratio (episodes of high positive δ13C) throughout the early Triassic before it was released several times again.[69] The latest research suggests that greenhouse gas release during the extinction event was dominated by volcanic carbon dioxide,[182] and while methane release had to have contributed, isotopic signatures show that thermogenic methane released from the Siberian Traps had consistently played a larger role than methane from clathrates and any other biogenic sources such as wetlands during the event.[183]

Hypercapnia and acidification

Marine organisms are more sensitive to changes in CO2 (carbon dioxide) levels than terrestrial organisms for a variety of reasons. CO2 is 28 times more soluble in water than is oxygen. Marine animals normally function with lower concentrations of CO2 in their bodies than land animals, as the removal of CO2 in air-breathing animals is impeded by the need for the gas to pass through the respiratory system's membranes (lungs' alveolus, tracheae, and the like), even when CO2 diffuses more easily than oxygen. In marine organisms, relatively modest but sustained increases in CO2 concentrations hamper the synthesis of proteins, reduce fertilization rates, and produce deformities in calcareous hard parts. An analysis of marine fossils from the Permian's final Changhsingian stage found that marine organisms with a low tolerance for hypercapnia (high concentration of carbon dioxide) had high extinction rates, and the most tolerant organisms had very slight losses. The most vulnerable marine organisms were those that produced calcareous hard parts (from calcium carbonate) and had low metabolic rates and weak respiratory systems, notably calcareous sponges, rugose and tabulate corals, calcite-depositing brachiopods, bryozoans, and echinoderms; about 81% of such genera became extinct. Close relatives without calcareous hard parts suffered only minor losses, such as sea anemones, from which modern corals evolved. Animals with high metabolic rates, well-developed respiratory systems, and non-calcareous hard parts had negligible losses except for conodonts, in which 33% of genera died out. This pattern is also consistent with what is known about the effects of hypoxia, a shortage but not total absence of oxygen. However, hypoxia cannot have been the only killing mechanism for marine organisms. Nearly all of the continental shelf waters would have had to become severely hypoxic to account for the magnitude of the extinction, but such a catastrophe would make it difficult to explain the very selective pattern of the extinction. Mathematical models of the Late Permian and Early Triassic atmospheres show a significant but protracted decline in atmospheric oxygen levels, with no acceleration near the P–Tr boundary. Minimum atmospheric oxygen levels in the Early Triassic are never less than present-day levels and so the decline in oxygen levels does not match the temporal pattern of the extinction.[102]

In addition, an increase in CO2 concentration is inevitably linked to ocean acidification, consistent with the preferential extinction of heavily calcified taxa and other signals in the rock record that suggest a more acidic ocean.[184] The decrease in ocean pH is calculated to be up to 0.7 units.[185] Ocean acidification was most extreme at mid-latitudes, and the major marine transgression associated with the end-Permian extinction is believed to have devastated shallow shelf communities in conjunction with anoxia.[5] Evidence from paralic facies spanning the Permian-Triassic boundary in western Guizhou and eastern Yunnan, however, shows a local marine transgression dominated by carbonate deposition, suggesting that ocean acidification did not occur across the entire globe and was likely limited to certain regions of the world's oceans.[186]

On land, the increasing acidification of rainwater caused increased soil erosion as a result of the increased acidity of forest soils, evidenced by the increased influx of terrestrially derived organic sediments found in marine sedimentary deposits during the end-Permian extinction.[187] A positive feedback loop further enhancing and prolonging soil acidification may have resulted from the decline of infaunal invertebrates like tubificids and chironomids, which remove acid metabolites from the soil.[188] The increased abundance of vermiculitic clays in Shansi, South China coinciding with the Permian-Triassic boundary strongly suggests a sharp drop in soil pH causally related to volcanogenic emissions of carbon dioxide and sulphur dioxide.[189]

Anoxia and euxinia

Evidence for widespread ocean anoxia (severe deficiency of oxygen) and euxinia (presence of hydrogen sulfide) is found from the Late Permian to the Early Triassic.[190][191][192] Throughout most of the Tethys and Panthalassic Oceans, evidence for anoxia, including fine laminations in sediments, small pyrite framboids, high uranium/thorium ratios, and biomarkers for green sulfur bacteria, appear at the extinction event.[193] However, evidence for anoxia precedes the extinction at some other sites, including Spiti, India,[194] Meishan, China,[195] Opal Creek, Alberta,[196] and Kap Stosch, Greenland.[197] Biomarkers for green sulfur bacteria, such as isorenieratane, the diagenetic product of isorenieratene, are widely used as indicators of photic zone euxinia because green sulfur bacteria require both sunlight and hydrogen sulfide to survive. Their abundance in sediments from the P–T boundary indicates euxinic conditions were present even in the shallow waters of the photic zone.[198] The disproportionate extinction of high-latitude marine species provides further evidence for oxygen depletion as a killing mechanism; low-latitude species living in warmer, less oxygenated waters are naturally better adapted to lower levels of oxygen and are able to migrate to higher latitudes during periods of global warming, whereas high-latitude organisms are unable to escape from warming, hypoxic waters at the poles.[199]

This spread of toxic, oxygen-depleted water would have devastated marine life, causing widespread die-offs. Models of ocean chemistry suggest that anoxia and euxinia were closely associated with hypercapnia (high levels of carbon dioxide).[200] This suggests that poisoning from hydrogen sulfide, anoxia, and hypercapnia acted together as a killing mechanism. Hypercapnia best explains the selectivity of the extinction, but anoxia and euxinia probably contributed to the high mortality of the event. The persistence of anoxia through the Early Triassic may explain the slow recovery of marine life and low levels of biodiversity after the extinction.[201] Models also show that anoxic events can cause catastrophic hydrogen sulfide emissions into the atmosphere (see below).[202]

The sequence of events leading to anoxic oceans may have been triggered by carbon dioxide emissions from the eruption of the Siberian Traps.[202] In that scenario, warming from the enhanced greenhouse effect would reduce the solubility of oxygen in seawater, causing the concentration of oxygen to decline. Increased weathering of the continents due to warming and the acceleration of the water cycle would increase the riverine flux of phosphate to the ocean. The phosphate would have supported greater primary productivity in the surface oceans. The increase in organic matter production would have caused more organic matter to sink into the deep ocean, where its respiration would further decrease oxygen concentrations. Once anoxia became established, it would have been sustained by a positive feedback loop because deep water anoxia tends to increase the recycling efficiency of phosphate, leading to even higher productivity.

A severe anoxic event at the end of the Permian would have allowed sulfate-reducing bacteria to thrive, causing the production of large amounts of hydrogen sulfide in the anoxic ocean, turning it euxinic. Upwelling of this water may have released massive hydrogen sulfide emissions into the atmosphere and would poison terrestrial plants and animals and severely weaken the ozone layer, exposing much of the life that remained to fatal levels of UV radiation.[202] Indeed, biomarker evidence for anaerobic photosynthesis by Chlorobiaceae (green sulfur bacteria) from the Late-Permian into the Early Triassic indicates that hydrogen sulfide did upwell into shallow waters because these bacteria are restricted to the photic zone and use sulfide as an electron donor. The hypothesis has the advantage of explaining the mass extinction of plants, which would have added to the methane levels and should otherwise have thrived in an atmosphere with a high level of carbon dioxide. Fossil spores from the end-Permian further support the theory; many spores show deformities that could have been caused by ultraviolet radiation, which would have been more intense after hydrogen sulfide emissions weakened the ozone layer.[203]

Some scientists have challenged the anoxia hypothesis on the grounds that long-lasting anoxic conditions could not have been supported if Late Permian thermohaline ocean circulation conformed to the "thermal mode" characterised by cooling at high latitudes. Anoxia may have persisted under a "haline mode" in which circulation was driven by subtropical evaporation, although the "haline mode" is highly unstable and was unlikely to have represented Late Permian oceanic circulation.[204]

Aridification

Analysis of the fossil river deposits of the floodplains indicate a shift from meandering to braided river patterns, indicating a very abrupt drying of the climate.[205] The climate change may have taken as little as 100,000 years, prompting the extinction of the unique Glossopteris flora and its associated herbivores, followed by the carnivorous guild.[206] In the North China Basin, highly arid climatic conditions are recorded during the latest Permian, near the Permian-Triassic boundary, with a swing towards increased precipitation during the Early Triassic, the latter likely assisting biotic recovery following the mass extinction.[135][134]

Evidence from the Sydney Basin of eastern Australia, on the other hand, suggests that the expansion of semi-arid and arid climatic belts across Pangaea was not immediate but was instead a gradual, prolonged process. Apart from the disappearance of peatlands, there was little evidence of significant sedimentological changes in depositional style across the Permian-Triassic boundary.[207] Instead, a modest shift to amplified seasonality and hotter summers is suggested by palaeoclimatological models based on weathering proxies from the region's Late Permian and Early Triassic deposits.[208]

In the Kuznetsk Basin of southwestern Siberia, an increase in aridity led to the demise of the humid-adapted cordaites forests in the region a few hundred thousand years before the Permian-Triassic boundary. This has been attributed to a broader poleward shift of drier, more arid climates during the late Changhsingian before the more abrupt main phase of the extinction at the Permian-Triassic boundary that disproportionately affected tropical and subtropical species.[96]

Asteroid impact

 
Artist's impression of a major impact event: A collision between Earth and an asteroid a few kilometers in diameter would release as much energy as the detonation of several million nuclear weapons.

Evidence that an impact event may have caused the Cretaceous–Paleogene extinction has led to speculation that similar impacts may have been the cause of other extinction events, including the P–Tr extinction, and thus to a search for evidence of impacts at the times of other extinctions, such as large impact craters of the appropriate age.

Reported evidence for an impact event from the P–Tr boundary level includes rare grains of shocked quartz in Australia and Antarctica;[209][210] fullerenes trapping extraterrestrial noble gases;[211] meteorite fragments in Antarctica;[212] and grains rich in iron, nickel, and silicon, which may have been created by an impact.[213] However, the accuracy of most of these claims has been challenged.[214][215][216][217] For example, quartz from Graphite Peak in Antarctica, once considered "shocked", has been re-examined by optical and transmission electron microscopy. The observed features were concluded to be due not to shock, but rather to plastic deformation, consistent with formation in a tectonic environment such as volcanism.[218]

An impact crater on the seafloor would be evidence of a possible cause of the P–Tr extinction, but such a crater would by now have disappeared. As 70% of the Earth's surface is currently sea, an asteroid or comet fragment is now perhaps more than twice as likely to hit the ocean as it is to hit land. However, Earth's oldest ocean-floor crust is the only 200 million years old as it is continually being destroyed and renewed by spreading and subduction. Furthermore, craters produced by very large impacts may be masked by extensive flood basalting from below after the crust is punctured or weakened.[219] Yet, subduction should not be entirely accepted as an explanation for the lack of evidence: as with the K-T event, an ejecta blanket stratum rich in siderophilic elements (such as iridium) would be expected in formations from the time.

A large impact might have triggered other mechanisms of extinction described above,[152] such as the Siberian Traps eruptions at either an impact site[220] or the antipode of an impact site.[152][221] The abruptness of an impact also explains why more species did not rapidly evolve to survive, as would be expected if the Permian–Triassic event had been slower and less global than a meteorite impact.

Bolide impact claims have been criticised on the grounds that they are unnecessary as explanations for the extinctions, and they do not fit the known data compatible with a protracted extinction spanning thousands of years.[222]

Possible impact sites

Possible impact craters proposed as the site of an impact causing the P–Tr extinction include the 250 km (160 mi) Bedout structure off the northwest coast of Australia[210] and the hypothesized 480 km (300 mi) Wilkes Land crater of East Antarctica.[223][224] An impact has not been proved in either case, and the idea has been widely criticized. The Wilkes Land geophysical feature is of very uncertain age, possibly later than the Permian–Triassic extinction.

Another impact hypothesis postulates that the impact event which formed the Araguainha crater, whose formation has been dated to 254.7 ± 2.5 million, a possible temporal range overlapping with the end-Permian extinction,[225] precipitated the mass extinction.[226] The impact occurred around extensive deposits of oil shale in the shallow marine Paraná–Karoo Basin, whose perturbation by the seismicity resulting from impact likely discharged about 1.6 teratonnes of methane into Earth's atmosphere, buttressing the already rapid warming caused by hydrocarbon release due to the Siberian Traps.[227] The large earthquakes generated by the impact would have additionally generated massive tsunamis across much of the globe.[226][228] Despite this, most palaeontologists reject the impact as being a significant driver of the extinction, citing the relatively low energy (equivalent to 105 to 106 of TNT, around two orders of magnitude lower than the impact energy believed to be required to induce mass extinctions) released by the impact.[227]

A 2017 paper noted the discovery of a circular gravity anomaly near the Falkland Islands which might correspond to an impact crater with a diameter of 250 km (160 mi),[229] as supported by seismic and magnetic evidence. Estimates for the age of the structure range up to 250 million years old. This would be substantially larger than the well-known 180 km (110 mi) Chicxulub impact crater associated with a later extinction. However, Dave McCarthy and colleagues from the British Geological Survey illustrated that the gravity anomaly is not circular and also that the seismic data presented by Rocca, Rampino and Baez Presser did not cross the proposed crater or provide any evidence for an impact crater.[230]

Methanogenic microbes

A hypothesis published in 2014 posits that a genus of anaerobic methanogenic archaea known as Methanosarcina was responsible for the event.[21] Three lines of evidence suggest that these microbes acquired a new metabolic pathway via gene transfer at about that time, enabling them to efficiently metabolize acetate into methane. That would have led to their exponential reproduction, allowing them to rapidly consume vast deposits of organic carbon that had accumulated in the marine sediment. The result would have been a sharp buildup of methane and carbon dioxide in the Earth's oceans and atmosphere, in a manner that may be consistent with the 13C/12C isotopic record. Massive volcanism facilitated this process by releasing large amounts of nickel, a scarce metal which is a cofactor for enzymes involved in producing methane.[21] On the other hand, in the canonical Meishan sections, the nickel concentration increases somewhat after the δ13C concentrations have begun to fall.[231]

Supercontinent Pangaea

 
Map of Pangaea showing where today's continents were at the Permian–Triassic boundary

In the mid-Permian (during the Kungurian age of the Permian's Cisuralian epoch), Earth's major continental plates joined, forming a supercontinent called Pangaea, which was surrounded by the superocean, Panthalassa.

Oceanic circulation and atmospheric weather patterns during the mid-Permian produced seasonal monsoons near the coasts and an arid climate in the vast continental interior.[232]

As the supercontinent formed, the ecologically diverse and productive coastal areas shrank. The shallow aquatic environments were eliminated and exposed formerly protected organisms of the rich continental shelves to increased environmental volatility.

Pangaea's formation depleted marine life at near catastrophic rates. However, Pangaea's effect on land extinctions is thought to have been smaller. In fact, the advance of the therapsids and increase in their diversity is attributed to the late Permian, when Pangaea's global effect was thought to have peaked.

While Pangaea's formation certainly initiated a long period of marine extinction, its impact on the "Great Dying" and the end of the Permian is uncertain.

Interstellar dust

John Gribbin argues that the Solar System last passed through a spiral arm of the Milky Way around 250 million years ago and that the resultant dusty gas clouds may have caused a dimming of the Sun, which combined with the effect of Pangaea to produce an ice age.[233]

Combination of causes

Possible causes supported by strong evidence appear to describe a sequence of catastrophes, each worse than the last: the Siberian Traps eruptions were bad enough alone, but because they occurred near coal beds and the continental shelf, they also triggered very large releases of carbon dioxide and methane. The resultant global warming may have caused perhaps the most severe anoxic event in the oceans' history: according to this theory, the oceans became so anoxic, anaerobic sulfur-reducing organisms dominated the chemistry of the oceans and caused massive emissions of toxic hydrogen sulfide.[102]

However, there may be some weak links in this chain of events: the changes in the 13C/12C ratio expected to result from a massive release of methane do not match the patterns seen throughout the early Triassic;[69] and the types of oceanic thermohaline circulation that may have existed at the end of the Permian are not likely to have supported deep-sea anoxia.[204]

See also

References

  1. ^ a b 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. Retrieved 14 January 2023.
  2. ^ McLoughlin, Steven (8 January 2021). "Age and Paleoenvironmental Significance of the Frazer Beach Member – A New Lithostratigraphic Unit Overlying the End-Permian Extinction Horizon in the Sydney Basin, Australia". Frontiers in Earth Science. 8 (600976): 605. Bibcode:2021FrEaS...8..605M. doi:10.3389/feart.2020.600976.
  3. ^ Algeo, Thomas J. (5 February 2012). . Astrobiology Magazine. Archived from the original on 2021-03-08.{{cite journal}}: CS1 maint: unfit URL (link)
  4. ^ Li, Dirson Jian (18 December 2012). "The tectonic cause of mass extinctions and the genomic contribution to biodiversification". Quantitative Biology. arXiv:1212.4229. Bibcode:2012arXiv1212.4229L.
  5. ^ a b Beauchamp, Benoit; Grasby, Stephen E. (15 September 2012). "Permian lysocline shoaling and ocean acidification along NW Pangea led to carbonate eradication and chert expansion". Palaeogeography, Palaeoclimatology, Palaeoecology. 350–352: 73–90. Bibcode:2012PPP...350...73B. doi:10.1016/j.palaeo.2012.06.014. Retrieved 21 December 2022.
  6. ^ ""Great Dying" lasted 200,000 years". National Geographic. 23 November 2011. Retrieved 1 April 2014.
  7. ^ St. Fleur, Nicholas (16 February 2017). "After Earth's worst mass extinction, life rebounded rapidly, fossils suggest". The New York Times. Retrieved 17 February 2017.
  8. ^ a b Jurikova, Hana; Gutjahr, Marcus; Wallmann, Klaus; Flögel, Sascha; Liebetrau, Volker; Posenato, Renato; et al. (November 2020). "Permian–Triassic mass extinction pulses driven by major marine carbon cycle perturbations". Nature Geoscience. 13 (11): 745–750. Bibcode:2020NatGe..13..745J. doi:10.1038/s41561-020-00646-4. ISSN 1752-0908. S2CID 224783993. Retrieved 8 November 2020.
  9. ^ a b Erwin, D.H. (1990). "The End-Permian Mass Extinction". Annual Review of Ecology and Systematics. 21: 69–91. doi:10.1146/annurev.es.21.110190.000441.
  10. ^ a b Chen, Yanlong; Richoz, Sylvain; Krystyn, Leopold; Zhang, Zhifei (August 2019). "Quantitative stratigraphic correlation of Tethyan conodonts across the Smithian-Spathian (Early Triassic) extinction event". Earth-Science Reviews. 195: 37–51. Bibcode:2019ESRv..195...37C. doi:10.1016/j.earscirev.2019.03.004. S2CID 135139479. Retrieved 28 October 2022.
  11. ^ Stanley, Steven M. (2016-10-18). "Estimates of the magnitudes of major marine mass extinctions in earth history". Proceedings of the National Academy of Sciences. 113 (42): E6325–E6334. Bibcode:2016PNAS..113E6325S. doi:10.1073/pnas.1613094113. ISSN 0027-8424. PMC 5081622. PMID 27698119.
  12. ^ a b Benton, M.J. (2005). When Life Nearly Died: The greatest mass extinction of all time. London: Thames & Hudson. ISBN 978-0-500-28573-2.
  13. ^ Bergstrom, Carl T.; Dugatkin, Lee Alan (2012). Evolution. Norton. p. 515. ISBN 978-0-393-92592-0.
  14. ^ a b c d e 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.
  15. ^ Yin, Hongfu; Feng, Qinglai; Lai, Xulong; Baud, Aymon; Tong, Jinnan (January 2007). "The protracted Permo-Triassic crisis and multi-episode extinction around the Permian–Triassic boundary". Global and Planetary Change. 55 (1–3): 1–20. Bibcode:2007GPC....55....1Y. doi:10.1016/j.gloplacha.2006.06.005. Retrieved 29 October 2022.
  16. ^ a b c d Jin YG, Wang Y, Wang W, Shang QH, Cao CQ, Erwin DH (2000). "Pattern of marine mass extinction near the Permian–Triassic boundary in south China". Science. 289 (5478): 432–436. Bibcode:2000Sci...289..432J. doi:10.1126/science.289.5478.432. PMID 10903200.
  17. ^ Yin H, Zhang K, Tong J, Yang Z, Wu S (2001). "The global stratotype section and point (GSSP) of the Permian–Triassic boundary". Episodes. 24 (2): 102–114. doi:10.18814/epiiugs/2001/v24i2/004.
  18. ^ Yin HF, Sweets WC, Yang ZY, Dickins JM (1992). "Permo–Triassic events in the eastern Tethys – an overview". In Sweet WC (ed.). Permo–Triassic Events in the Eastern Tethys: Stratigraphy, classification, and relations with the western Tethys. Cambridge: Cambridge University Press. pp. 1–7. ISBN 978-0-521-54573-0.
  19. ^ Darcy E. Ogdena & Norman H. Sleep (2011). "Explosive eruption of coal and basalt and the end-Permian mass extinction". Proceedings of the National Academy of Sciences of the United States of America. 109 (1): 59–62. Bibcode:2012PNAS..109...59O. doi:10.1073/pnas.1118675109. PMC 3252959. PMID 22184229.
  20. ^ a b Kaiho, Kunio; Aftabuzzaman, Md; Jones, David S.; Tian, Li (4 November 2020). "Pulsed volcanic combustion events coincident with the end-Permian terrestrial disturbance and the following global crisis". Geology. 49 (3): 289–293. doi:10.1130/G48022.1. ISSN 0091-7613.   Available under CC BY 4.0.
  21. ^ a b c Rothman, D.H.; Fournier, G.P.; French, K.L.; Alm, E.J.; Boyle, E.A.; Cao, C.; Summons, R.E. (2014-03-31). "Methanogenic burst in the end-Permian carbon cycle". Proceedings of the National Academy of Sciences. 111 (15): 5462–5467. Bibcode:2014PNAS..111.5462R. doi:10.1073/pnas.1318106111. PMC 3992638. PMID 24706773. – Lay summary: Chandler, David L. (March 31, 2014). "Ancient whodunit may be solved: Methane-producing microbes did it!". Science Daily.
  22. ^ Stanley, Steven M. (8 September 2009). "Evidence from ammonoids and conodonts for multiple Early Triassic mass extinctions". Proceedings of the National Academy of Sciences of the United States of America. 106 (36): 15264–15267. Bibcode:2009PNAS..10615264S. doi:10.1073/pnas.0907992106. PMC 2741239. PMID 19721005.
  23. ^ Hochuli, Peter A.; Sanson-Barrera, Anna; Schneebeli-Hermann, Elke; Bucher, Hugo (24 June 2016). "Severest crisis overlooked—Worst disruption of terrestrial environments postdates the Permian–Triassic mass extinction". Scientific Reports. 6 (1): 28372. Bibcode:2016NatSR...628372H. doi:10.1038/srep28372. PMC 4920029. PMID 27340926.
  24. ^ Caravaca, Gwénaël; Thomazo, Christophe; Vennin, Emmanuelle; Olivier, Nicolas; Cocquerez, Théophile; Escarguel, Gilles; Fara, Emmanuel; Jenks, James F.; Bylund, Kevin G.; Stephen, Daniel A.; Brayard, Arnaud (July 2017). "Early Triassic fluctuations of the global carbon cycle: New evidence from paired carbon isotopes in the western USA basin". Global and Planetary Change. 154: 10–22. Bibcode:2017GPC...154...10C. doi:10.1016/j.gloplacha.2017.05.005. S2CID 135330761. Retrieved 13 November 2022.
  25. ^ "It took Earth ten million years to recover from greatest mass extinction". ScienceDaily. 27 May 2012. Retrieved 28 May 2012.
  26. ^ Brayard, Arnaud; Krumenacker, L. J.; Botting, Joseph P.; Jenks, James F.; Bylund, Kevin G.; Fara, Emmanuel; Vennin, Emmanuelle; Olivier, Nicolas; Goudemand, Nicolas; Saucède, Thomas; Charbonnier, Sylvain; Romano, Carlo; Doguzhaeva, Larisa; Thuy, Ben; Hautmann, Michael; Stephen, Daniel A.; Thomazo, Christophe; Escarguel, Gilles (15 February 2017). "Unexpected Early Triassic marine ecosystem and the rise of the Modern evolutionary fauna". Science Advances. 13 (2): e1602159. Bibcode:2017SciA....3E2159B. doi:10.1126/sciadv.1602159. PMC 5310825. PMID 28246643.
  27. ^ Smith, Christopher P. A.; Laville, Thomas; Fara, Emmanuel; Escarguel, Gilles; Olivier, Nicolas; Vennin, Emmanuelle; et al. (2021-10-04). "Exceptional fossil assemblages confirm the existence of complex Early Triassic ecosystems during the early Spathian". Scientific Reports. 11 (1): 19657. Bibcode:2021NatSR..1119657S. doi:10.1038/s41598-021-99056-8. ISSN 2045-2322. PMC 8490361. PMID 34608207.
  28. ^ Hofmann, Richard; Goudemand, Nicolas; Wasmer, Martin; Bucher, Hugo; Hautmann, Michael (1 October 2011). "New trace fossil evidence for an early recovery signal in the aftermath of the end-Permian mass extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 310 (3–4): 216–226. Bibcode:2011PPP...310..216H. doi:10.1016/j.palaeo.2011.07.014. Retrieved 20 December 2022.
  29. ^ a b Woods, Adam D.; Alms, Paul D.; Monarrez, Pedro M.; Mata, Scott (1 January 2019). "The interaction of recovery and environmental conditions: An analysis of the outer shelf edge of western North America during the early Triassic". Palaeogeography, Palaeoclimatology, Palaeoecology. 513: 52–64. Bibcode:2019PPP...513...52W. doi:10.1016/j.palaeo.2018.05.014. Retrieved 20 December 2022.
  30. ^ a b Payne, J.L.; Lehrmann, D.J.; Wei, J.; Orchard, M.J.; Schrag, D.P.; Knoll, A.H. (2004). "Large Perturbations of the Carbon Cycle During Recovery from the End-Permian Extinction" (PDF). Science. 305 (5683): 506–9. Bibcode:2004Sci...305..506P. CiteSeerX 10.1.1.582.9406. doi:10.1126/science.1097023. PMID 15273391. S2CID 35498132.
  31. ^ a b c d e f g h i j k l m McElwain, J. C.; Punyasena, S. W. (2007). "Mass extinction events and the plant fossil record". Trends in Ecology & Evolution. 22 (10): 548–557. doi:10.1016/j.tree.2007.09.003. PMID 17919771.
  32. ^ a b Retallack, G. J.; Veevers, J. J.; Morante, R. (1996). "Global coal gap between Permian–Triassic extinctions and middle Triassic recovery of peat forming plants". GSA Bulletin. 108 (2): 195–207. Bibcode:1996GSAB..108..195R. doi:10.1130/0016-7606(1996)108<0195:GCGBPT>2.3.CO;2.
  33. ^ a b c d e Erwin, D.H. (1993). The great Paleozoic crisis; Life and death in the Permian. Columbia University Press. ISBN 978-0-231-07467-4.
  34. ^ a b Burgess, S.D. (2014). "High-precision timeline for Earth's most severe extinction". Proceedings of the National Academy of Sciences of the United States of America. 111 (9): 3316–3321. Bibcode:2014PNAS..111.3316B. doi:10.1073/pnas.1317692111. PMC 3948271. PMID 24516148.
  35. ^ Magaritz M (1989). "13C minima follow extinction events: A clue to faunal radiation". Geology. 17 (4): 337–340. Bibcode:1989Geo....17..337M. doi:10.1130/0091-7613(1989)017<0337:CMFEEA>2.3.CO;2.
  36. ^ Krull SJ, Retallack JR (2000). "13C depth profiles from paleosols across the Permian–Triassic boundary: Evidence for methane release". GSA Bulletin. 112 (9): 1459–1472. Bibcode:2000GSAB..112.1459K. doi:10.1130/0016-7606(2000)112<1459:CDPFPA>2.0.CO;2. ISSN 0016-7606.
  37. ^ a b Dolenec T, Lojen S, Ramovs A (2001). "The Permian-Triassic boundary in Western Slovenia (Idrijca Valley section): magnetostratigraphy, stable isotopes, and elemental variations". Chemical Geology. 175 (1–2): 175–190. Bibcode:2001ChGeo.175..175D. doi:10.1016/S0009-2541(00)00368-5.
  38. ^ Musashi M, Isozaki Y, Koike T, Kreulen R (2001). "Stable carbon isotope signature in mid-Panthalassa shallow-water carbonates across the Permo–Triassic boundary: Evidence for 13C-depleted ocean". Earth and Planetary Science Letters. 193 (1–2): 9–20. Bibcode:2001E&PSL.191....9M. doi:10.1016/S0012-821X(01)00398-3.
  39. ^ "Daily CO2". Mauna Loa Observatory.
  40. ^ Visscher, Henk; Looy, Cindy V.; Collinson, Margaret E.; Brinkhuis, Henk; Cittert, Johanna H. A. van Konijnenburg-van; Kürschner, Wolfram M.; Sephton, Mark A. (2004-08-31). "Environmental mutagenesis during the end-Permian ecological crisis". Proceedings of the National Academy of Sciences of the United States of America. 101 (35): 12952–12956. Bibcode:2004PNAS..10112952V. doi:10.1073/pnas.0404472101. ISSN 0027-8424. PMC 516500. PMID 15282373.
  41. ^ a b Visscher H, Brinkhuis H, Dilcher DL, Elsik WC, Eshet Y, Looy CW, Rampino MR, Traverse A (1996). "The terminal Paleozoic fungal event: Evidence of terrestrial ecosystem destabilization and collapse". Proceedings of the National Academy of Sciences. 93 (5): 2155–2158. Bibcode:1996PNAS...93.2155V. doi:10.1073/pnas.93.5.2155. PMC 39926. PMID 11607638.
  42. ^ Foster, C.B.; Stephenson, M.H.; Marshall, C.; Logan, G.A.; Greenwood, P.F. (2002). "A revision of Reduviasporonites Wilson 1962: Description, illustration, comparison and biological affinities". Palynology. 26 (1): 35–58. doi:10.2113/0260035.
  43. ^ Diez, José B.; Broutin, Jean; Grauvogel-Stamm, Léa; Borquin, Sylvie; Bercovici, Antoine; Ferrer, Javier (October 2010). "Anisian floras from the NE Iberian Peninsula and Balearic Islands: A synthesis". Review of Palaeobotany and Palynology. 162 (3): 522–542. doi:10.1016/j.revpalbo.2010.09.003. Retrieved 8 January 2023.
  44. ^ López-Gómez, J. & Taylor, E.L. (2005). "Permian–Triassic transition in Spain: A multidisciplinary approach". Palaeogeography, Palaeoclimatology, Palaeoecology. 229 (1–2): 1–2. doi:10.1016/j.palaeo.2005.06.028.
  45. ^ a b c Looy CV, Twitchett RJ, Dilcher DL, van Konijnenburg-Van Cittert JH, Visscher H (2005). "Life in the end-Permian dead zone". Proceedings of the National Academy of Sciences. 98 (4): 7879–7883. Bibcode:2001PNAS...98.7879L. doi:10.1073/pnas.131218098. PMC 35436. PMID 11427710. See image 2
  46. ^ a b Ward PD, Botha J, Buick R, de Kock MO, Erwin DH, Garrison GH, Kirschvink JL, Smith R (2005). "Abrupt and gradual extinction among late Permian land vertebrates in the Karoo Basin, South Africa" (PDF). Science. 307 (5710): 709–714. Bibcode:2005Sci...307..709W. CiteSeerX 10.1.1.503.2065. doi:10.1126/science.1107068. PMID 15661973. S2CID 46198018.
  47. ^ Retallack, G.J.; Smith, R.M.H.; Ward, P.D. (2003). "Vertebrate extinction across Permian-Triassic boundary in Karoo Basin, South Africa". Bulletin of the Geological Society of America. 115 (9): 1133–1152. Bibcode:2003GSAB..115.1133R. doi:10.1130/B25215.1.
  48. ^ Sephton, M.A.; Visscher, H.; Looy, C.V.; Verchovsky, A.B.; Watson, J.S. (2009). "Chemical constitution of a Permian-Triassic disaster species". Geology. 37 (10): 875–878. Bibcode:2009Geo....37..875S. doi:10.1130/G30096A.1.
  49. ^ Rampino, Michael R.; Eshet, Yoram (January 2018). "The fungal and acritarch events as time markers for the latest Permian mass extinction: An update". Geoscience Frontiers. 9 (1): 147–154. doi:10.1016/j.gsf.2017.06.005. Retrieved 24 December 2022.
  50. ^ Rampino MR, Prokoph A, Adler A (2000). "Tempo of the end-Permian event: High-resolution cyclostratigraphy at the Permian–Triassic boundary". Geology. 28 (7): 643–646. Bibcode:2000Geo....28..643R. doi:10.1130/0091-7613(2000)28<643:TOTEEH>2.0.CO;2. ISSN 0091-7613.
  51. ^ Wang, S.C.; Everson, P.J. (2007). "Confidence intervals for pulsed mass extinction events". Paleobiology. 33 (2): 324–336. doi:10.1666/06056.1. S2CID 2729020.
  52. ^ a b c Twitchett RJ, Looy CV, Morante R, Visscher H, Wignall PB (2001). "Rapid and synchronous collapse of marine and terrestrial ecosystems during the end-Permian biotic crisis". Geology. 29 (4): 351–354. Bibcode:2001Geo....29..351T. doi:10.1130/0091-7613(2001)029<0351:RASCOM>2.0.CO;2. ISSN 0091-7613.
  53. ^ Fielding, Christopher R.; Frank, Tracy D.; Savatic, Katarina; Mays, Chris; McLoughlin, Stephen; Vajda, Vivi; Nicoll, Robert S. (15 May 2022). "Environmental change in the late Permian of Queensland, NE Australia: The warmup to the end-Permian Extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 594: 110936. Bibcode:2022PPP...594k0936F. doi:10.1016/j.palaeo.2022.110936. S2CID 247514266. Retrieved 23 December 2022.
  54. ^ a b Retallack, G.J.; Metzger, C.A.; Greaver, T.; Jahren, A.H.; Smith, R.M.H.; Sheldon, N.D. (November–December 2006). "Middle-Late Permian mass extinction on land". Bulletin of the Geological Society of America. 118 (11–12): 1398–1411. Bibcode:2006GSAB..118.1398R. doi:10.1130/B26011.1.
  55. ^ Stanley SM, Yang X (1994). "A double mass extinction at the end of the Paleozoic Era". Science. 266 (5189): 1340–1344. Bibcode:1994Sci...266.1340S. doi:10.1126/science.266.5189.1340. PMID 17772839. S2CID 39256134.
  56. ^ 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.
  57. ^ Shen, S. & Shi, G.R. (2002). "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. ISSN 0094-8373. S2CID 35611701.
  58. ^ 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.
  59. ^ Bond, D.P.G.; Wignall, P.B.; Wang, W.; Izon, G.; Jiang, H.S.; Lai, X.L.; et al. (2010). "The mid-Capitanian (Middle Permian) mass extinction and carbon isotope record of South China". Palaeogeography, Palaeoclimatology, Palaeoecology. 292 (1–2): 282–294. Bibcode:2010PPP...292..282B. doi:10.1016/j.palaeo.2010.03.056.
  60. ^ Racki G (1999). "Silica-secreting biota and mass extinctions: survival processes and patterns". Palaeogeography, Palaeoclimatology, Palaeoecology. 154 (1–2): 107–132. Bibcode:1999PPP...154..107R. doi:10.1016/S0031-0182(99)00089-9.
  61. ^ Bambach, R.K.; Knoll, A.H.; Wang, S.C. (December 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. ISSN 0094-8373. S2CID 17279135.
  62. ^ a b Knoll AH (2004). "Biomineralization and evolutionary history". In Dove PM, DeYoreo JJ, Weiner S (eds.). (PDF). Archived from the original (PDF) on 2010-06-20.
  63. ^ Stanley, S.M. (2008). "Predation defeats competition on the seafloor". Paleobiology. 34 (1): 1–21. doi:10.1666/07026.1. S2CID 83713101. Retrieved 2008-05-13.
  64. ^ Stanley, S.M. (2007). "An Analysis of the History of Marine Animal Diversity". Paleobiology. 33 (sp6): 1–55. doi:10.1666/06020.1. S2CID 86014119.
  65. ^ McKinney, M.L. (1987). "Taxonomic selectivity and continuous variation in mass and background extinctions of marine taxa". Nature. 325 (6100): 143–145. Bibcode:1987Natur.325..143M. doi:10.1038/325143a0. S2CID 13473769.
  66. ^ Hofmann, R.; Buatois, L. A.; MacNaughton, R. B.; Mángano, M. G. (15 June 2015). "Loss of the sedimentary mixed layer as a result of the end-Permian extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 428: 1–11. doi:10.1016/j.palaeo.2015.03.036. Retrieved 19 December 2022.
  67. ^ "Permian : The Marine Realm and The End-Permian Extinction". paleobiology.si.edu. Retrieved 2016-01-26.
  68. ^ a b "Permian extinction". Encyclopædia Britannica. Retrieved 2016-01-26.
  69. ^ a b c d e f g Payne, J.L.; Lehrmann, D.J.; Wei, J.; Orchard, M.J.; Schrag, D.P.; Knoll, A.H. (2004). "Large Perturbations of the Carbon Cycle During Recovery from the End-Permian Extinction" (PDF). Science. 305 (5683): 506–9. Bibcode:2004Sci...305..506P. CiteSeerX 10.1.1.582.9406. doi:10.1126/science.1097023. PMID 15273391. S2CID 35498132.
  70. ^ Knoll, A.H.; Bambach, R.K.; Canfield, D.E.; Grotzinger, J.P. (1996). "Comparative Earth history and Late Permian mass extinction". Science. 273 (5274): 452–457. Bibcode:1996Sci...273..452K. doi:10.1126/science.273.5274.452. PMID 8662528. S2CID 35958753.
  71. ^ Leighton, L.R.; Schneider, C.L. (2008). "Taxon characteristics that promote survivorship through the Permian–Triassic interval: transition from the Paleozoic to the Mesozoic brachiopod fauna". Paleobiology. 34 (1): 65–79. doi:10.1666/06082.1. S2CID 86843206.
  72. ^ 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.
  73. ^ Saunders, W. B.; Greenfest-Allen, E.; Work, D. M.; Nikolaeva, S. V. (2008). "Morphologic and taxonomic history of Paleozoic ammonoids in time and morphospace". Paleobiology. 34 (1): 128–154. doi:10.1666/07053.1. S2CID 83650272.
  74. ^ Twitchett, Richard J. (20 August 2007). "The Lilliput effect in the aftermath of the end-Permian extinction event". Palaeogeography, Palaeoclimatology, Palaeoecology. 252 (1–2): 132–144. Bibcode:2007PPP...252..132T. doi:10.1016/j.palaeo.2006.11.038. Retrieved 13 January 2023.
  75. ^ Schaal, Ellen K.; Clapham, Matthew E.; Rego, Brianna L.; Wang, Steve C.; Payne, Jonathan L. (6 November 2015). "Comparative size evolution of marine clades from the Late Permian through Middle Triassic". Paleobiology. 42 (1): 127–142. doi:10.1017/pab.2015.36. S2CID 18552375. Retrieved 13 January 2023.
  76. ^ Song, Haijun; Tong, Jinnan; Chen, Zhong-Qiang (15 July 2011). "Evolutionary dynamics of the Permian–Triassic foraminifer size: Evidence for Lilliput effect in the end-Permian mass extinction and its aftermath". Palaeogeography, Palaeoclimatology, Palaeoecology. 308 (1–2): 98–110. Bibcode:2011PPP...308...98S. doi:10.1016/j.palaeo.2010.10.036. Retrieved 13 January 2023.
  77. ^ He, Weihong; Twitchett, Richard J.; Zhang, Y.; Shi, G. R.; Feng, Q.-L.; Yu, J.-X.; Wu, S.-B.; Peng, X.-F. (9 November 2010). "Controls on body size during the Late Permian mass extinction event". Geobiology. 8 (5): 391–402. doi:10.1111/j.1472-4669.2010.00248.x. PMID 20550584. S2CID 23096333. Retrieved 13 January 2023.
  78. ^ Chen, Jing; Song, Haijun; He, Weihong; Tong, Jinnan; Wang, Fengyu; Wu, Shunbao (1 April 2019). "Size variation of brachiopods from the Late Permian through the Middle Triassic in South China: Evidence for the Lilliput Effect following the Permian-Triassic extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 519: 248–257. Bibcode:2019PPP...519..248C. doi:10.1016/j.palaeo.2018.07.013. S2CID 134806479. Retrieved 13 January 2023.
  79. ^ Huang, Yunfei; Tong, Jinnan; Tian, Li; Song, Haijun; Chu, Daoliang; Miao, Xue; Song, Ting (1 January 2023). "Temporal shell-size variations of bivalves in South China from the Late Permian to the early Middle Triassic". Palaeogeography, Palaeoclimatology, Palaeoecology. 609: 111307. Bibcode:2023PPP...609k1307H. doi:10.1016/j.palaeo.2022.111307. S2CID 253368808. Retrieved 13 January 2023.
  80. ^ Yates, Adam M.; Neumann, Frank H.; Hancox, P. John (2 February 2012). "The Earliest Post-Paleozoic Freshwater Bivalves Preserved in Coprolites from the Karoo Basin, South Africa". PLOS ONE. 7 (2): e30228. Bibcode:2012PLoSO...730228Y. doi:10.1371/journal.pone.0030228. PMC 3271088. PMID 22319562.
  81. ^ Foster, W. J.; Gliwa, J.; Lembke, C.; Pugh, A. C.; Hofmann, R.; Tietje, M.; Varela, S.; Foster, L. C.; Korn, D.; Aberhan, M. (January 2020). "Evolutionary and ecophenotypic controls on bivalve body size distributions following the end-Permian mass extinction". Global and Planetary Change. 185: 103088. Bibcode:2020GPC...18503088F. doi:10.1016/j.gloplacha.2019.103088. S2CID 213294384. Retrieved 13 January 2023.
  82. ^ Chu, Daoliang; Tong, Jinnan; Song, Haijun; Benton, Michael James; Song, Huyue; Yu, Jianxin; Qiu, Xincheng; Huang, Yunfei; Tian, Li (1 October 2015). "Lilliput effect in freshwater ostracods during the Permian–Triassic extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 435: 38–52. Bibcode:2015PPP...435...38C. doi:10.1016/j.palaeo.2015.06.003. Retrieved 13 January 2023.
  83. ^ Brayard, Arnaud; Nützel, Alexander; Stephen, Daniel A.; Bylund, Kevin G.; Jenks, Jim; Bucher, Hugo (1 February 2010). "Gastropod evidence against the Early Triassic Lilliput effect". Geology. 37 (2): 147–150. Bibcode:2010Geo....38..147B. doi:10.1130/G30553.1. Retrieved 13 January 2023.
  84. ^ Fraiser, M. L.; Twitchett, Richard J.; Frederickson, J. A.; Metcalfe, B.; Bottjer, D. J. (1 January 2011). "Gastropod evidence against the Early Triassic Lilliput effect: COMMENT". Geology. 39 (1): e232. Bibcode:2011Geo....39E.232F. doi:10.1130/G31614C.1. Retrieved 13 January 2023.
  85. ^ Brayard, Arnaud; Nützel, Alexander; Kajm, Andrzej; Escarguel, Gilles; Hautmann, Michael; Stephen, Daniel A.; Bylund, Kevin G.; Jenks, Jim; Bucher, Hugo (1 January 2011). "Gastropod evidence against the Early Triassic Lilliput effect: REPLY". Geology. 39 (1): e233. Bibcode:2011Geo....39E.233B. doi:10.1130/G31765Y.1. Retrieved 13 January 2023.
  86. ^ Labandeira, Conrad (1 January 2005), "The fossil record of insect extinction: New approaches and future directions", American Entomologist, 51: 14–29, doi:10.1093/ae/51.1.14
  87. ^ a b Labandeira CC, Sepkoski JJ (1993). "Insect diversity in the fossil record". Science. 261 (5119): 310–315. Bibcode:1993Sci...261..310L. CiteSeerX 10.1.1.496.1576. doi:10.1126/science.11536548. PMID 11536548.
  88. ^ Sole RV, Newman M (2003). "Extinctions and Biodiversity in the Fossil Record". In Canadell JG, Mooney HA (eds.). Encyclopedia of Global Environmental Change, The Earth System. Biological and Ecological Dimensions of Global Environmental Change. Vol. 2. New York: Wiley. pp. 297–391. ISBN 978-0-470-85361-0.
  89. ^ Nowak, Hendrik; Schneebeli-Hermann, Elke; Kustatscher, Evelyn (23 January 2019). "No mass extinction for land plants at the Permian–Triassic transition". Nature Communications. 10 (1): 384. Bibcode:2019NatCo..10..384N. doi:10.1038/s41467-018-07945-w. PMC 6344494. PMID 30674875.
  90. ^ "The Dino Directory – Natural History Museum".
  91. ^ Gulbranson, Erik L.; Mellum, Morgan M.; Corti, Valentina; Dahlseid, Aidan; Atkinson, Brian A.; Ryberg, Patricia E.; Cornamusini, Giancula (24 May 2022). "Paleoclimate-induced stress on polar forested ecosystems prior to the Permian–Triassic mass extinction". Scientific Reports. 12 (1): 8702. Bibcode:2022NatSR..12.8702G. doi:10.1038/s41598-022-12842-w. PMC 9130125. PMID 35610472.
  92. ^ Retallack GJ (1995). "Permian–Triassic life crisis on land". Science. 267 (5194): 77–80. Bibcode:1995Sci...267...77R. doi:10.1126/science.267.5194.77. PMID 17840061. S2CID 42308183.
  93. ^ Cascales-Miñana, B.; Cleal, C. J. (2011). "Plant fossil record and survival analyses". Lethaia. 45: 71–82. doi:10.1111/j.1502-3931.2011.00262.x.
  94. ^ a b c d Bodnar, Josefina; Coturel, Eliana P.; Falco, Juan Ignacio; Beltrána, Marisol (4 March 2021). "An updated scenario for the end-Permian crisis and the recovery of Triassic land flora in Argentina". Historical Biology. 33 (12): 3654–3672. doi:10.1080/08912963.2021.1884245. S2CID 233810158. Retrieved 24 December 2022.
  95. ^ a b c Vajda, Vivi; McLoughlin, Stephen (April 2007). "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 (1–2): 99–112. doi:10.1016/j.revpalbo.2005.09.007. Retrieved 24 December 2022.
  96. ^ a b Davydov, V. I.; Karasev, E. V.; Nurgalieva, N. G.; Schmitz, M. D.; Budnikov, I. V.; Biakov, A. S.; Kuzina, D. M.; Silantiev, V. V.; Urazaeva, M. N.; Zharinova, V. V.; Zorina, S. O.; Gareev, B.; Vasilenko, D. V. (1 July 2021). "Climate and biotic evolution during the Permian-Triassic transition in the temperate Northern Hemisphere, Kuznetsk Basin, Siberia, Russia". Palaeogeography, Palaeoclimatology, Palaeoecology. 573: 110432. Bibcode:2021PPP...573k0432D. doi:10.1016/j.palaeo.2021.110432. S2CID 235530804. Retrieved 19 December 2022.
  97. ^ Chu, Daoliang; Yu, Jianxin; Tong, Jinnan; Benton, Michael James; Song, Haijun; Huang, Yunfei; Song, Ting; Tian, Li (November 2016). "Biostratigraphic correlation and mass extinction during the Permian-Triassic transition in terrestrial-marine siliciclastic settings of South China". Global and Planetary Change. 146: 67–88. Bibcode:2016GPC...146...67C. doi:10.1016/j.gloplacha.2016.09.009. hdl:1983/2d475883-42ea-4988-b01c-0a56912e6aec. Retrieved 12 November 2022.
  98. ^ a b c Feng, Zhuo; Wei, Hai-Bo; Guo, Yun; He, Xiao-Yuan; Sui, Qun; Zhou, Yu; Liu, Hang-Yu; Gou, Xu-Dong; Lv, Yong (May 2020). "From rainforest to herbland: New insights into land plant responses to the end-Permian mass extinction". Earth-Science Reviews. 204: 103153. Bibcode:2020ESRv..20403153F. doi:10.1016/j.earscirev.2020.103153. S2CID 216433847.
  99. ^ Biswas, Raman Kumar; Kaiho, Kunio; Saito, Ryosuke; Tian, Li; Shi, Zhiqiang (December 2020). "Terrestrial ecosystem collapse and soil erosion before the end-Permian marine extinction: Organic geochemical evidence from marine and non-marine records". Global and Planetary Change. 195: 103327. Bibcode:2020GPC...19503327B. doi:10.1016/j.gloplacha.2020.103327. S2CID 224900905. Retrieved 21 December 2022.
  100. ^ Maxwell, W.D. (1992). "Permian and Early Triassic extinction of non-marine tetrapods". Palaeontology. 35: 571–583.
  101. ^ "Bristol University – News – 2008: Mass extinction".
  102. ^ a b c Knoll AH, Bambach RK, Payne JL, Pruss S, Fischer WW (2007). "Paleophysiology and end-Permian mass extinction" (PDF). Earth and Planetary Science Letters. 256 (3–4): 295–313. Bibcode:2007E&PSL.256..295K. doi:10.1016/j.epsl.2007.02.018. Retrieved 2021-12-13.
  103. ^ Sepkoski, J. John (8 February 2016). "A kinetic model of Phanerozoic taxonomic diversity. III. Post-Paleozoic families and mass extinctions". Paleobiology. 10 (2): 246–267. doi:10.1017/S0094837300008186. S2CID 85595559.
  104. ^ Romano, Carlo; Koot, Martha B.; Kogan, Ilja; Brayard, Arnaud; Minikh, Alla V.; Brinkmann, Winand; et al. (February 2016). "Permian–Triassic Osteichthyes (bony fishes): diversity dynamics and body size evolution". Biological Reviews. 91 (1): 106–147. doi:10.1111/brv.12161. PMID 25431138. S2CID 5332637.
  105. ^ Scheyer, Torsten M.; Romano, Carlo; Jenks, Jim; Bucher, Hugo (19 March 2014). "Early Triassic Marine Biotic Recovery: The Predators' Perspective". PLOS ONE. 9 (3): e88987. Bibcode:2014PLoSO...988987S. doi:10.1371/journal.pone.0088987. PMC 3960099. PMID 24647136.
  106. ^ Komatsu, Toshifumi; Huyen, Dang Tran; Jin-Hua, Chen (1 June 2008). "Lower Triassic bivalve assemblages after the end-Permian mass extinction in South China and North Vietnam". Paleontological Research. 12 (2): 119–128. doi:10.2517/1342-8144(2008)12[119:LTBAAT]2.0.CO;2. S2CID 131144468. Retrieved 6 November 2022.
  107. ^ Gould, S.J.; Calloway, C.B. (1980). "Clams and brachiopodsships that pass in the night". Paleobiology. 6 (4): 383–396. doi:10.1017/S0094837300003572. S2CID 132467749.
  108. ^ Jablonski, D. (8 May 2001). "Lessons from the past: Evolutionary impacts of mass extinctions". Proceedings of the National Academy of Sciences. 98 (10): 5393–5398. Bibcode:2001PNAS...98.5393J. doi:10.1073/pnas.101092598. PMC 33224. PMID 11344284.
  109. ^ Kaim, Andrzej; Nützel, Alexander (July 2011). "Dead bellerophontids walking – The short Mesozoic history of the Bellerophontoidea (Gastropoda)". Palaeogeography, Palaeoclimatology, Palaeoecology. 308 (1–2): 190–199. Bibcode:2011PPP...308..190K. doi:10.1016/j.palaeo.2010.04.008.
  110. ^ Hautmann, Michael (29 September 2009). "The first scallop" (PDF). Paläontologische Zeitschrift. 84 (2): 317–322. doi:10.1007/s12542-009-0041-5. S2CID 84457522.
  111. ^ Hautmann, Michael; Ware, David; Bucher, Hugo (August 2017). "Geologically oldest oysters were epizoans on Early Triassic ammonoids". Journal of Molluscan Studies. 83 (3): 253–260. doi:10.1093/mollus/eyx018.
  112. ^ Hofmann, Richard; Hautmann, Michael; Brayard, Arnaud; Nützel, Alexander; Bylund, Kevin G.; Jenks, James F.; et al. (May 2014). "Recovery of benthic marine communities from the end-Permian mass extinction at the low latitudes of eastern Panthalassa" (PDF). Palaeontology. 57 (3): 547–589. doi:10.1111/pala.12076. S2CID 6247479.
  113. ^ Brayard, A.; Escarguel, G.; Bucher, H.; Monnet, C.; Bruhwiler, T.; Goudemand, N.; et al. (27 August 2009). "Good genes and good luck: Ammonoid diversity and the end-Permian mass extinction". Science. 325 (5944): 1118–1121. Bibcode:2009Sci...325.1118B. doi:10.1126/science.1174638. PMID 19713525. S2CID 1287762.
  114. ^ Wignall, P. B.; Twitchett, R.J. (24 May 1996). "Oceanic Anoxia and the End Permian Mass Extinction". Science. 272 (5265): 1155–1158. Bibcode:1996Sci...272.1155W. doi:10.1126/science.272.5265.1155. PMID 8662450. S2CID 35032406.
  115. ^ Hofmann, Richard; Hautmann, Michael; Bucher, Hugo (October 2015). "Recovery dynamics of benthic marine communities from the Lower Triassic Werfen Formation, northern Italy". Lethaia. 48 (4): 474–496. doi:10.1111/let.12121.
  116. ^ Hautmann, Michael; Bagherpour, Borhan; Brosse, Morgane; Frisk, Åsa; Hofmann, Richard; Baud, Aymon; et al. (September 2015). "Competition in slow motion: The unusual case of benthic marine communities in the wake of the end-Permian mass extinction". Palaeontology. 58 (5): 871–901. doi:10.1111/pala.12186. S2CID 140688908.
  117. ^ Petsios, Elizabeth; Thompson, Jeffrey R.; Pietsch, Carlie; Bottjer, David J. (1 January 2019). "Biotic impacts of temperature before, during, and after the end-Permian extinction: A multi-metric and multi-scale approach to modeling extinction and recovery dynamics". Palaeogeography, Palaeoclimatology, Palaeoecology. 513: 86–99. Bibcode:2019PPP...513...86P. doi:10.1016/j.palaeo.2017.08.038. S2CID 134149088. Retrieved 2 December 2022.
  118. ^ Friesenbichler, Evelyn; Hautmann, Michael; Nützel, Alexander; Urlichs, Max; Bucher, Hugo (24 July 2018). "Palaeoecology of Late Ladinian (Middle Triassic) benthic faunas from the Schlern/Sciliar and Seiser Alm/Alpe di Siusi area (South Tyrol, Italy)" (PDF). PalZ. 93 (1): 1–29. doi:10.1007/s12542-018-0423-7. S2CID 134192673.
  119. ^ Friesenbichler, Evelyn; Hautmann, Michael; Grădinaru, Eugen; Bucher, Hugo; Brayard, Arnaud (12 October 2019). "A highly diverse bivalve fauna from a Bithynian (Anisian, Middle Triassic) – microbial buildup in North Dobrogea (Romania)" (PDF). Papers in Palaeontology. doi:10.1002/spp2.1286. S2CID 208555999.
  120. ^ Sepkoski, J. John (1997). "Biodiversity: Past, Present, and Future". Journal of Paleontology. 71 (4): 533–539. doi:10.1017/S0022336000040026. PMID 11540302. S2CID 27430390.
  121. ^ a b Wagner PJ, Kosnik MA, Lidgard S (2006). "Abundance Distributions Imply Elevated Complexity of Post-Paleozoic Marine Ecosystems". Science. 314 (5803): 1289–1292. Bibcode:2006Sci...314.1289W. doi:10.1126/science.1133795. PMID 17124319. S2CID 26957610.
  122. ^ Chen, Zhong-Qiang; Tong, Jinnan; Liao, Zhuo-Ting; Chen, Jing (August 2010). "Structural changes of marine communities over the Permian–Triassic transition: Ecologically assessing the end-Permian mass extinction and its aftermath". Global and Planetary Change. 73 (1–2): 123–140. Bibcode:2010GPC....73..123C. doi:10.1016/j.gloplacha.2010.03.011. Retrieved 6 November 2022.
  123. ^ Clapham ME, Bottjer DJ, Shen S (2006). . Geological Society of America Abstracts with Programs. 38 (7): 117. Archived from the original on 2015-12-08. Retrieved 2008-03-28.
  124. ^ Posenato, Renato; Holmer, Lars E.; Prinoth, Herwig (1 April 2014). "Adaptive strategies and environmental significance of lingulid brachiopods across the late Permian extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 399: 373–384. Bibcode:2014PPP...399..373P. doi:10.1016/j.palaeo.2014.01.028. Retrieved 14 January 2023.
  125. ^ Foote, M. (1999). "Morphological diversity in the evolutionary radiation of Paleozoic and post-Paleozoic crinoids". Paleobiology. 25 (sp1): 1–116. doi:10.1666/0094-8373(1999)25[1:MDITER]2.0.CO;2. ISSN 0094-8373. JSTOR 2666042. S2CID 85586709.
  126. ^ Baumiller, T. K. (2008). "Crinoid Ecological Morphology". Annual Review of Earth and Planetary Sciences. 36 (1): 221–249. Bibcode:2008AREPS..36..221B. doi:10.1146/annurev.earth.36.031207.124116.
  127. ^ Martindale, Rowan C.; Foster, William J.; Velledits, Felicitász (1 January 2019). "The survival, recovery, and diversification of metazoan reef ecosystems following the end-Permian mass extinction event". Palaeogeography, Palaeoclimatology, Palaeoecology. 513: 100–115. Bibcode:2019PPP...513..100M. doi:10.1016/j.palaeo.2017.08.014. S2CID 135338869. Retrieved 2 December 2022.
  128. ^ Aftabuzzaman, Md.; Kaiho, Kunio; Biswas, Raman Kumar; Liu, Yuqing; Saito, Ryosuke; Tian, Li; Bhat, Ghulam M.; Chen, Zhong-Qiang (October 2021). "End-Permian terrestrial disturbance followed by the complete plant devastation, and the vegetation proto-recovery in the earliest-Triassic recorded in coastal sea sediments". Global and Planetary Change. 205: 103621. Bibcode:2021GPC...20503621A. doi:10.1016/j.gloplacha.2021.103621. Retrieved 24 December 2022.
  129. ^ Looy CV, Brugman WA, Dilcher DL, Visscher H (1999). "The delayed resurgence of equatorial forests after the Permian–Triassic ecologic crisis". Proceedings of the National Academy of Sciences of the United States of America. 96 (24): 13857–13862. Bibcode:1999PNAS...9613857L. doi:10.1073/pnas.96.24.13857. PMC 24155. PMID 10570163.
  130. ^ a b Hochuli, Peter A.; Hermann, Elke; Vigran, Jorunn Os; Bucher, Hugo; Weissert, Helmut (December 2010). "Rapid demise and recovery of plant ecosystems across the end-Permian extinction event". Global and Planetary Change. 74 (3–4): 144–155. Bibcode:2010GPC....74..144H. doi:10.1016/j.gloplacha.2010.10.004. Retrieved 24 December 2022.
  131. ^ Grauvogel-Stamm, Léa; Ash, Sidney R. (September–October 2005). "Recovery of the Triassic land flora from the end-Permian life crisis". Comptes Rendus Palevol. 4 (6–7): 593–608. doi:10.1016/j.crpv.2005.07.002. Retrieved 24 December 2022.
  132. ^ Michaelsen P (2002). "Mass extinction of peat-forming plants and the effect on fluvial styles across the Permian–Triassic boundary, northern Bowen Basin, Australia". Palaeogeography, Palaeoclimatology, Palaeoecology. 179 (3–4): 173–188. Bibcode:2002PPP...179..173M. doi:10.1016/S0031-0182(01)00413-8.
  133. ^ Botha, J. & Smith, R.M.H. (2007). (PDF). Lethaia. 40 (2): 125–137. doi:10.1111/j.1502-3931.2007.00011.x. Archived from the original (PDF) on 2008-09-10. Retrieved 2008-07-02.
  134. ^ a b Zhu, Zhicai; Liu, Yongqing; Kuang, Hongwei; Newell, Andrew J.; Peng, Nan; Cui, Mingming; Benton, Michael J. (September 2022). "Improving paleoenvironment in North China aided Triassic biotic recovery on land following the end-Permian mass extinction". Global and Planetary Change. 216: 103914. Bibcode:2022GPC...21603914Z. doi:10.1016/j.gloplacha.2022.103914.
  135. ^ a b Yu, Yingyue; Tian, Li; Chu, Daoliang; Song, Huyue; Guo, Wenwei; Tong, Jinnan (1 January 2022). "Latest Permian–Early Triassic paleoclimatic reconstruction by sedimentary and isotopic analyses of paleosols from the Shichuanhe section in central North China Basin". Palaeogeography, Palaeoclimatology, Palaeoecology. 585: 110726. Bibcode:2022PPP...585k0726Y. doi:10.1016/j.palaeo.2021.110726. S2CID 239498183. Retrieved 6 November 2022.
  136. ^ Benton, M.J. (2004). Vertebrate Paleontology. Blackwell Publishers. xii–452. ISBN 978-0-632-05614-9.
  137. ^ Ruben, J.A. & Jones, T.D. (2000). "Selective Factors Associated with the Origin of Fur and Feathers". American Zoologist. 40 (4): 585–596. doi:10.1093/icb/40.4.585.
  138. ^ Yates AM, Warren AA (2000). "The phylogeny of the 'higher' temnospondyls (Vertebrata: Choanata) and its implications for the monophyly and origins of the Stereospondyli". Zoological Journal of the Linnean Society. 128 (1): 77–121. doi:10.1111/j.1096-3642.2000.tb00650.x.
  139. ^ Yao, Mingtao; Sun, Zuoyu; Meng, Qingqiang; Li, Jiachun; Jiang, Dayong (15 August 2022). "Vertebrate coprolites from Middle Triassic Chang 7 Member in Ordos Basin, China: Palaeobiological and palaeoecological implications". Palaeogeography, Palaeoclimatology, Palaeoecology. 600: 111084. Bibcode:2022PPP...600k1084M. doi:10.1016/j.palaeo.2022.111084. S2CID 249186414. Retrieved 8 January 2023.
  140. ^ Zhou MF, Malpas J, Song XY, Robinson PT, Sun M, Kennedy AK, Lesher CM, Keays RR (2002). "A temporal link between the Emeishan large igneous province (SW China) and the end-Guadalupian mass extinction". Earth and Planetary Science Letters. 196 (3–4): 113–122. Bibcode:2002E&PSL.196..113Z. doi:10.1016/S0012-821X(01)00608-2.
  141. ^ Wignall, Paul B.; Sun, Y.; Bond, D.P.G.; Izon, G.; Newton, R.J.; Vedrine, S.; et al. (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.
  142. ^ Andy Saunders; Marc Reichow (2009). "The Siberian Traps – area and volume". Retrieved 2009-10-18.
  143. ^ Saunders, Andy & Reichow, Marc (January 2009). "The Siberian Traps and the End-Permian mass extinction: a critical review" (PDF). Chinese Science Bulletin. 54 (1): 20–37. Bibcode:2009ChSBu..54...20S. doi:10.1007/s11434-008-0543-7. hdl:2381/27540. S2CID 1736350.
  144. ^ Reichow, Marc K.; Pringle, M.S.; Al'Mukhamedov, A.I.; Allen, M.B.; Andreichev, V.L.; Buslov, M.M.; et al. (2009). "The timing and extent of the eruption of the Siberian Traps large igneous province: Implications for the end-Permian environmental crisis" (PDF). Earth and Planetary Science Letters. 277 (1–2): 9–20. Bibcode:2009E&PSL.277....9R. doi:10.1016/j.epsl.2008.09.030. hdl:2381/4204.
  145. ^ Augland, L. E.; Ryabov, V. V.; Vernikovsky, V. A.; Planke, S.; Polozov, A. G.; Callegaro, S.; Jerram, D. A.; Svensen, H. H. (10 December 2019). "The main pulse of the Siberian Traps expanded in size and composition". Scientific Reports. 9 (1): 18723. Bibcode:2019NatSR...918723A. doi:10.1038/s41598-019-54023-2. PMC 6904769. PMID 31822688.
  146. ^ Kamo, SL (2003). "Rapid eruption of Siberian flood-volcanic rocks and evidence for coincidence with the Permian–Triassic boundary and mass extinction at 251 Ma". Earth and Planetary Science Letters. 214 (1–2): 75–91. Bibcode:2003E&PSL.214...75K. doi:10.1016/S0012-821X(03)00347-9.
  147. ^ Burgess, Seth D.; Bowring, Samuel; Shen, Shu-zhong (2014-03-04). "High-precision timeline for Earth's most severe extinction". Proceedings of the National Academy of Sciences. 111 (9): 3316–3321. Bibcode:2014PNAS..111.3316B. doi:10.1073/pnas.1317692111. ISSN 0027-8424. PMC 3948271. PMID 24516148.
  148. ^ Black, Benjamin A.; Weiss, Benjamin P.; Elkins-Tanton, Linda T.; Veselovskiy, Roman V.; Latyshev, Anton (2015-04-30). "Siberian Traps volcaniclastic rocks and the role of magma-water interactions". Geological Society of America Bulletin. 127 (9–10): B31108.1. Bibcode:2015GSAB..127.1437B. doi:10.1130/B31108.1. ISSN 0016-7606.
  149. ^ Burgess, Seth D.; Bowring, Samuel A. (2015-08-01). "High-precision geochronology confirms voluminous magmatism before, during, and after Earth's most severe extinction". Science Advances. 1 (7): e1500470. Bibcode:2015SciA....1E0470B. doi:10.1126/sciadv.1500470. ISSN 2375-2548. PMC 4643808. PMID 26601239.
  150. ^ Fischman, Josh. Giant eruptions and giant extinctions. Scientific American (video). Retrieved 2016-03-11.
  151. ^ Cui, Ying; Kump, Lee R. (October 2015). "Global warming and the end-Permian extinction event: Proxy and modeling perspectives". Earth-Science Reviews. 149: 5–22. Bibcode:2015ESRv..149....5C. doi:10.1016/j.earscirev.2014.04.007.
  152. ^ a b c d e f White, R.V. (2002). "Earth's biggest 'whodunnit': Unravelling the clues in the case of the end-Permian mass extinction" (PDF). Philosophical Transactions of the Royal Society of London. 360 (1801): 2963–2985. Bibcode:2002RSPTA.360.2963W. doi:10.1098/rsta.2002.1097. PMID 12626276. S2CID 18078072. Retrieved 2008-01-12.
  153. ^ "Driver of the largest mass extinction in the history of the Earth identified". phys.org. Retrieved 8 November 2020.
  154. ^ "Large volcanic eruption caused the largest mass extinction". phys.org. Retrieved 8 December 2020.
  155. ^ Baresel, Björn; Bucher, Hugo; Bagherpour, Borhan; Brosse, Morgane; Guodun, Kuang; Schaltegger, Urs (6 March 2017). "Timing of global regression and microbial bloom linked with the Permian-Triassic boundary mass extinction: implications for driving mechanisms". Scientific Reports. 7: 43630. Bibcode:2017NatSR...743630B. doi:10.1038/srep43630. PMC 5338007. PMID 28262815.
  156. ^ "Volcanism". Hooper Museum. hoopermuseum.earthsci.carleton.ca. Ottawa, Ontario, Canada: Carleton University.
  157. ^ Svensen, Henrik; Planke, Sverre; Polozov, Alexander G.; Schmidbauer, Norbert; Corfu, Fernando; Podladchikov, Yuri Y.; Jamtveit, Bjørn (30 January 2009). "Siberian gas venting and the end-Permian environmental crisis". Earth and Planetary Science Letters. 297 (3–4): 490–500. Bibcode:2009E&PSL.277..490S. doi:10.1016/j.epsl.2008.11.015. Retrieved 13 January 2023.
  158. ^ Konstantinov, Konstantin M.; Bazhenov, Mikhail L.; Fetisova, Anna M.; Khutorskoy, Mikhail D. (May 2014). "Paleomagnetism of trap intrusions, East Siberia: Implications to flood basalt emplacement and the Permo–Triassic crisis of biosphere". Earth and Planetary Science Letters. 394: 242–253. Bibcode:2014E&PSL.394..242K. doi:10.1016/j.epsl.2014.03.029.
  159. ^ Burgess, S. D.; Muirhead, J. D.; Bowring, S. A. (31 July 2017). "Initial pulse of Siberian Traps sills as the trigger of the end-Permian mass extinction". Nature Communications. 8 (1): 164. Bibcode:2017NatCo...8..164B. doi:10.1038/s41467-017-00083-9. PMC 5537227. PMID 28761160. S2CID 3312150.
  160. ^ Corso, Jacopo Dal; Song, Haijun; Callegaro, Sara; Chu, Daoliang; Sun, Yadong; Hilton, Jason; Grasby, Stephen E.; Joachimski, Michael M.; Wignall, Paul B. (22 February 2022). "Environmental crises at the Permian–Triassic mass extinction". Nature Reviews Earth & Environment. 3 (3): 197–214. Bibcode:2022NRvEE...3..197D. doi:10.1038/s43017-021-00259-4. S2CID 247013868. Retrieved 20 December 2022.
  161. ^ Corso, Jacopo Dal; Mills, Benjamin J. W.; Chu, Daoling; Newton, Robert J.; Mather, Tamsin A.; Shu, Wenchao; Wu, Yuyang; Tong, Jinnan; Wignall, Paul B. (11 June 2020). "Permo–Triassic boundary carbon and mercury cycling linked to terrestrial ecosystem collapse". Nature Communications. 11 (1): 2962. Bibcode:2020NatCo..11.2962D. doi:10.1038/s41467-020-16725-4. PMC 7289894. PMID 32528009.
  162. ^ Verango, Dan (24 January 2011). "Ancient mass extinction tied to torched coal". USA Today.
  163. ^ Grasby, Stephen E.; Sanei, Hamed & Beauchamp, Benoit (January 23, 2011). "Catastrophic dispersion of coal fly ash into oceans during the latest Permian extinction". Nature Geoscience. 4 (2): 104–107. Bibcode:2011NatGe...4..104G. doi:10.1038/ngeo1069.
  164. ^ "Researchers find smoking gun of world's biggest extinction: Massive volcanic eruption, burning coal and accelerated greenhouse gas choked out life" (Press release). University of Calgary. January 23, 2011. Retrieved 2011-01-26.
  165. ^ Yang, Q.Y. (2013). "The chemical compositions and abundances of volatiles in the Siberian large igneous province: Constraints on magmatic CO2 and SO2 emissions into the atmosphere". Chemical Geology. 339: 84–91. Bibcode:2013ChGeo.339...84T. doi:10.1016/j.chemgeo.2012.08.031.
  166. ^ a b Grasby, Stephen E.; Beauchamp, Benoit; Bond, David P. G.; Wignall, Paul B.; Talavera, Cristina; Galloway, Jennifer M.; Piepjohn, Karsten; Reinhardt, Lutz; Blomeier, Dirk (1 September 2015). "Progressive environmental deterioration in northwestern Pangea leading to the latest Permian extinction". Geological Society of America Bulletin. 127 (9–10): 1331–1347. Bibcode:2015GSAB..127.1331G. doi:10.1130/B31197.1. Retrieved 14 January 2023.
  167. ^ 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.
  168. ^ Sanei, Hamed; Grasby, Stephen E.; Beauchamp, Benoit (1 January 2012). "Latest Permian mercury anomalies". Geology. 40 (1): 63–66. Bibcode:2012Geo....40...63S. doi:10.1130/G32596.1. Retrieved 14 January 2023.
  169. ^ Grasby, Stephen E.; Liu, Xiaojun; Yin, Runsheng; Ernst, Richard E.; Chen, Zhuoheng (19 May 2020). "Toxic mercury pulses into late Permian terrestrial and marine environments". Geology. 48 (8): 830–833. Bibcode:2020Geo....48..830G. doi:10.1130/G47295.1. S2CID 219495628. Retrieved 14 January 2023.
  170. ^ Li, Menghan; Grasby, Stephen E.; Wang, Shui-Jiong; Zhang, Xiaolin; Wasylenki, Laura E.; Xu, Yilun; Sun, Mingzhao; Beauchamp, Benoit; Hu, Dongping; Shen, Yanan (1 April 2021). "Nickel isotopes link Siberian Traps aerosol particles to the end-Permian mass extinction". Nature Communications. 12 (1): 2024. Bibcode:2021NatCo..12.2024L. doi:10.1038/s41467-021-22066-7. PMC 8016954. PMID 33795666.
  171. ^ Nelson, D. A.; Cottle, J. M. (29 March 2019). "Tracking voluminous Permian volcanism of the Choiyoi Province into central Antarctica". Lithosphere. 11 (3): 386–398. Bibcode:2019Lsphe..11..386N. doi:10.1130/L1015.1. S2CID 135130436. Retrieved 14 January 2023.
  172. ^ Dickens, G.R. (2001). "The potential volume of oceanic methane hydrates with variable external conditions". Organic Geochemistry. 32 (10): 1179–1193. doi:10.1016/S0146-6380(01)00086-9.
  173. ^ Palfy J, Demeny A, Haas J, Htenyi M, Orchard MJ, Veto I (2001). "Carbon isotope anomaly at the Triassic–Jurassic boundary from a marine section in Hungary". Geology. 29 (11): 1047–1050. Bibcode:2001Geo....29.1047P. doi:10.1130/0091-7613(2001)029<1047:CIAAOG>2.0.CO;2. ISSN 0091-7613.
  174. ^ a b c Berner, R.A. (2002). "Examination of hypotheses for the Permo-Triassic boundary extinction by carbon cycle modeling". Proceedings of the National Academy of Sciences. 99 (7): 4172–4177. Bibcode:2002PNAS...99.4172B. doi:10.1073/pnas.032095199. PMC 123621. PMID 11917102.
  175. ^ a b Dickens GR, O'Neil JR, Rea DK, Owen RM (1995). "Dissociation of oceanic methane hydrate as a cause of the carbon isotope excursion at the end of the Paleocene". Paleoceanography. 10 (6): 965–971. Bibcode:1995PalOc..10..965D. doi:10.1029/95PA02087.
  176. ^ Schrag DP, Berner RA, Hoffman PF, Halverson GP (2002). "On the initiation of a snowball Earth". Geochemistry, Geophysics, Geosystems. 3 (6): 1–21. Bibcode:2002GGG....3fQ...1S. doi:10.1029/2001GC000219. Preliminary abstract at Schrag, D.P. (June 2001). . Geological Society of America. Archived from the original on 2018-04-25. Retrieved 2008-04-20.
  177. ^ a b Benton, M.J.; Twitchett, R.J. (2003). "How to kill (almost) all life: The end-Permian extinction event". Trends in Ecology & Evolution. 18 (7): 358–365. doi:10.1016/S0169-5347(03)00093-4.
  178. ^ Ryskin, Gregory (September 2003). "Methane-driven oceanic eruptions and mass extinctions". Geology. 31 (9): 741–744. Bibcode:2003Geo....31..741R. doi:10.1130/G19518.1.
  179. ^ Reichow MK, Saunders AD, White RV, Pringle MS, Al'Muhkhamedov AI, Medvedev AI, Kirda NP (2002). " 40Ar 39Ar dates from the West Siberian Basin: Siberian flood basalt province doubled" (PDF). Science. 296 (5574): 1846–1849. Bibcode:2002Sci...296.1846R. doi:10.1126/science.1071671. PMID 12052954. S2CID 28964473.
  180. ^ Holser WT, Schoenlaub HP, Attrep Jr M, Boeckelmann K, Klein P, Magaritz M, Orth CJ, Fenninger A, Jenny C, Kralik M, Mauritsch H, Pak E, Schramm JF, Stattegger K, Schmoeller R (1989). "A unique geochemical record at the Permian/Triassic boundary". Nature. 337 (6202): 39–44. Bibcode:1989Natur.337...39H. doi:10.1038/337039a0. S2CID 8035040.
  181. ^ Dobruskina, I.A. (1987). "Phytogeography of Eurasia during the early Triassic". Palaeogeography, Palaeoclimatology, Palaeoecology. 58 (1–2): 75–86. Bibcode:1987PPP....58...75D. doi:10.1016/0031-0182(87)90007-1.
  182. ^ Cui, Ying; Li, Mingsong; van Soelen, Elsbeth E.; Peterse, Francien; M. Kürschner, Wolfram (7 September 2021). "Massive and rapid predominantly volcanic CO2 emission during the end-Permian mass extinction". Earth, Atmospheric, and Planetary Sciences. 118 (37): e2014701118. Bibcode:2021PNAS..11814701C. doi:10.1073/pnas.2014701118. PMC 8449420. PMID 34493684.
  183. ^ Wu, Yuyang; Chu, Daoliang; Tong, Jinnan; Song, Haijun; Dal Corso, Jacopo; Wignall, Paul B.; Song, Huyue; Du, Yong; Cui, Ying (9 April 2021). "Six-fold increase of atmospheric pCO2 during the Permian–Triassic mass extinction". Nature Communications. 12 (1): 2137. Bibcode:2021NatCo..12.2137W. doi:10.1038/s41467-021-22298-7. PMC 8035180. PMID 33837195. S2CID 233200774.
  184. ^ Payne, J.; Turchyn, A.; Paytan, A.; Depaolo, D.; Lehrmann, D.; Yu, M.; Wei, J. (2010). "Calcium isotope constraints on the end-Permian mass extinction". Proceedings of the National Academy of Sciences of the United States of America. 107 (19): 8543–8548. Bibcode:2010PNAS..107.8543P. doi:10.1073/pnas.0914065107. PMC 2889361. PMID 20421502.
  185. ^ Clarkson, M.; Kasemann, S.; Wood, R.; Lenton, T.; Daines, S.; Richoz, S.; et al. (2015-04-10). "Ocean acidification and the Permo-Triassic mass extinction" (PDF). Science. 348 (6231): 229–232. Bibcode:2015Sci...348..229C. doi:10.1126/science.aaa0193. hdl:10871/20741. PMID 25859043. S2CID 28891777.
  186. ^ Wignall, Paul B.; Chu, Daoliang; Hilton, Jason M.; Corso, Jacopo Dal; Wu, Yuyang; Wang, Yao; Atkinson, Jed; Tong, Jinnan (June 2020). "Death in the shallows: The record of Permo-Triassic mass extinction in paralic settings, southwest China". Global and Planetary Change. 189: 103176. Bibcode:2020GPC...18903176W. doi:10.1016/j.gloplacha.2020.103176. S2CID 216302513. Retrieved 7 November 2022.
  187. ^ Sephton, Mark A.; Jiao, Dan; Engel, Michael H.; Looy, Cindy V.; Visscher, Henk (1 February 2015). "Terrestrial acidification during the end-Permian biosphere crisis?". Geology. 43 (2): 159–162. Bibcode:2015Geo....43..159S. doi:10.1130/G36227.1. Retrieved 23 December 2022.
  188. ^ 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.
  189. ^ Xu, Guozhen; Deconinck, Jean-François; Feng, Qinglai; Baudin, François; Pellenard, Pierre; Shen, Jun; Bruneau, Ludovic (15 May 2017). "Clay mineralogical characteristics at the Permian–Triassic Shangsi section and their paleoenvironmental and/or paleoclimatic significance". Palaeogeography, Palaeoclimatology, Palaeoecology. 474: 152–163. Bibcode:2017PPP...474..152X. doi:10.1016/j.palaeo.2016.07.036. Retrieved 23 December 2022.
  190. ^ Song, Haijun; Wignall, Paul B.; Tong, Jinnan; Bond, David P. G.; Song, Huyue; Lai, Xulong; Zhang, Kexin; Wang, Hongmei; Chen, Yanlong (1 November 2012). "Geochemical evidence from bio-apatite for multiple oceanic anoxic events during Permian–Triassic transition and the link with end-Permian extinction and recovery". Earth and Planetary Science Letters. 353–354: 12–21. Bibcode:2012E&PSL.353...12S. doi:10.1016/j.epsl.2012.07.005. Retrieved 13 January 2023.
  191. ^ Hülse, Dominik; Lau, Kimberly V.; Van de Velde, Sebastiaan J.; Arndt, Sandra; Meyer, Katja M.; Ridgwell, Andy (28 October 2021). "End-Permian marine extinction due to temperature-driven nutrient recycling and euxinia". Nature Geoscience. 14 (11): 862–867. Bibcode:2021NatGe..14..862H. doi:10.1038/s41561-021-00829-7. S2CID 240076553. Retrieved 8 January 2023.
  192. ^ Schobben, Martin; Foster, William J.; Sleveland, Arve R. N.; Zuchuat, Valentin; Svensen, Henrik H.; Planke, Sverre; Bond, David P. G.; Marcelis, Fons; Newton, Robert J.; Wignall, Paul B.; Poulton, Simon W. (17 August 2020). "A nutrient control on marine anoxia during the end-Permian mass extinction". Nature Geoscience. 13 (9): 640–646. Bibcode:2020NatGe..13..640S. doi:10.1038/s41561-020-0622-1. S2CID 221146234. Retrieved 8 January 2023.
permian, triassic, extinction, event, permian, triassic, extinction, event, also, known, latest, permian, extinction, event, permian, extinction, colloquially, great, dying, formed, boundary, between, permian, triassic, geologic, periods, with, them, paleozoic. The Permian Triassic P T P Tr 3 4 extinction event 5 also known as the Latest Permian extinction event the End Permian Extinction 6 and colloquially as the Great Dying 7 formed the boundary between the Permian and Triassic geologic periods and with them the Paleozoic and Mesozoic eras respectively approximately 251 9 million years ago 8 It is the Earth s most severe known extinction event 9 10 with the extinction of 57 of biological families 83 of genera 81 of marine species 11 12 13 and 70 of terrestrial vertebrate species 14 It was the largest known mass extinction of insects Marine extinction intensity during the Phanerozoic Millions of years ago H K Pg Tr J P Tr Cap Late D O S Plot of extinction intensity percentage of marine genera that are present in each interval of time but do not exist in the following interval vs time in the past 1 Geological periods are annotated by abbreviation and colour above The Permian Triassic extinction event is the most significant event for marine genera with just over 50 according to this source perishing source and image info Permian Triassic boundary at Frazer Beach in New South Wales with the End Permian extinction event located just above the coal layer 2 There is evidence for one to three distinct pulses or phases of extinction 15 14 16 17 18 The scientific consensus is that the main cause of extinction was the large amount of carbon dioxide emitted by the volcanic eruptions that created the Siberian Traps which elevated global temperatures and in the oceans led to widespread anoxia and acidification 19 Proposed contributing factors include the emission of much additional carbon dioxide from the thermal decomposition of hydrocarbon deposits including oil and coal triggered by the eruptions and emissions of methane by novel methanogenic microorganisms perhaps nourished by minerals dispersed in the eruptions 20 21 The speed of recovery from the extinction is disputed Some scientists estimate that it took 10 million years until the Middle Triassic due both to the severity of the extinction and because grim conditions returned periodically over the course of the Early Triassic 22 23 24 causing further extinction events such as the Smithian Spathian boundary extinction 10 25 However studies in Bear Lake County near Paris Idaho 26 and nearby sites in Idaho and Nevada 27 showed a relatively quick rebound in a localized Early Triassic marine ecosystem taking around 3 million years to recover while an unusually diverse and complex ichnobiota is known from Italy less than a million years after the end Permian extinction 28 suggesting that the impact of the extinction may have been felt less severely in some areas than in others Differential environmental stress and instability following the extinction event has been theorised as an explanation for the disparity in the pace of biotic recovery between different regions and environments 29 Contents 1 Dating 2 Extinction patterns 2 1 Marine organisms 2 2 Terrestrial invertebrates 2 3 Terrestrial plants 2 4 Terrestrial vertebrates 3 Biotic recovery 3 1 Marine ecosystems 3 2 Terrestrial plants 3 2 1 Coal gap 3 3 Terrestrial vertebrates 3 4 Terrestrial invertebrates 4 Hypotheses about cause 4 1 Volcanism 4 1 1 Siberian Traps 4 1 2 Choiyoi Silicic Large Igneous Province 4 2 Methane clathrate gasification 4 3 Hypercapnia and acidification 4 4 Anoxia and euxinia 4 5 Aridification 4 6 Asteroid impact 4 6 1 Possible impact sites 4 7 Methanogenic microbes 4 8 Supercontinent Pangaea 4 9 Interstellar dust 4 10 Combination of causes 5 See also 6 References 7 Further reading 8 External linksDating EditTimeline of recoveryThis box viewtalkedit 262 260 258 256 254 252 250 248 246 244 242 240 238 Guadal LopingianEarlyMiddleCapitanianWuchiaping Changhsing InduanOlenekianAnisianLadinian Extinction event d 13C stabilises at 2 Calcified green algae return 30 Full recovery of woody trees 31 Corals return 32 Scleractiniancorals amp calcified sponges return 30 MesozoicPaleozoicPermianTriassicAn approximate timeline of recovery after the Permian Triassic extinction Axis scale millions of years ago PreꞒ Ꞓ O S D C P T J K Pg N Previously it was thought that rock sequences spanning the Permian Triassic boundary were too few and contained too many gaps for scientists to reliably determine its details 33 However it is now possible to date the extinction with millennial precision U Pb zircon dates from five volcanic ash beds from the Global Stratotype Section and Point for the Permian Triassic boundary at Meishan China establish a high resolution age model for the extinction allowing exploration of the links between global environmental perturbation carbon cycle disruption mass extinction and recovery at millennial timescales The extinction occurred between 251 941 0 037 and 251 880 0 031 million years ago a duration of 60 48 thousand years 34 A large approximately 0 9 abrupt global decrease in the ratio of the stable isotope carbon 13 to that of carbon 12 coincides with this extinction 31 35 36 37 38 and is sometimes used to identify the Permian Triassic boundary in rocks that are unsuitable for radiometric dating 37 Further evidence for environmental change around the P Tr boundary suggests an 8 C 14 F rise in temperature 31 and an increase in CO2 levels by 2 000 ppm for comparison the concentration immediately before the Industrial Revolution was 280 ppm 31 and the amount today is about 415 ppm 39 There is also evidence of increased ultraviolet radiation reaching the earth causing the mutation of plant spores 31 40 It has been suggested that the Permian Triassic boundary is associated with a sharp increase in the abundance of marine and terrestrial fungi caused by the sharp increase in the amount of dead plants and animals fed upon by the fungi 41 For a while this fungal spike was used by some paleontologists to identify the Permian Triassic boundary in rocks that are unsuitable for radiometric dating or have a lack of suitable index fossils but even the proposers of the fungal spike hypothesis pointed out that fungal spikes may have been a repeating phenomenon created by the post extinction ecosystem during the earliest Triassic 41 The very idea of a fungal spike has been criticized on several grounds including Reduviasporonites the most common supposed fungal spore may be a fossilized alga 31 42 the spike did not appear worldwide 43 44 45 and in many places it did not fall on the Permian Triassic boundary 46 The reduviasporonites may even represent a transition to a lake dominated Triassic world rather than an earliest Triassic zone of death and decay in some terrestrial fossil beds 47 Newer chemical evidence agrees better with a fungal origin for Reduviasporonites diluting these critiques 48 49 Uncertainty exists regarding the duration of the overall extinction and about the timing and duration of various groups extinctions within the greater process Some evidence suggests that there were multiple extinction pulses 14 or that the extinction was long and spread out over a few million years with a sharp peak in the last million years of the Permian 46 50 Statistical analyses of some highly fossiliferous strata in Meishan Zhejiang Province in southeastern China suggest that the main extinction was clustered around one peak 16 Recent research shows that different groups became extinct at different times for example while difficult to date absolutely ostracod and brachiopod extinctions were separated by around 670 000 to 1 17 million years 51 In a well preserved sequence in east Greenland the decline of animal life is concentrated in a period approximately 10 000 to 60 000 years long with plants taking an additional several hundred thousand years to show the full impact of the event 52 Palaeoenvironmental analysis of Lopingian strata in the Bowen Basin of Queensland indicates numerous intermittent periods of environmental stress from the middle to late Lopingian leading up to the end Permian extinction proper 53 An older theory still supported in some recent papers 14 54 is that there were two major extinction pulses 9 4 million years apart separated by a period of extinctions well above the background level and that the final extinction killed off only about 80 of marine species alive at that time while the other losses occurred during the first pulse or the interval between pulses According to this theory one of these extinction pulses occurred at the end of the Guadalupian epoch of the Permian 14 55 For example all dinocephalian genera died out at the end of the Guadalupian 54 as did the Verbeekinidae a family of large size fusuline foraminifera 56 The impact of the end Guadalupian extinction on marine organisms appears to have varied between locations and between taxonomic groups brachiopods and corals had severe losses 57 58 Studies of the timing and causes of the Permian Triassic extinction are complicated by the often overlooked Capitanian extinction also called the Guadalupian extinction just one of perhaps two mass extinctions in the late Permian that closely preceded the Permian Triassic event In short when the Permian Triassic starts it is difficult to know whether the end Capitanian had finished depending on the factor considered 1 59 Some of the extinctions dated to the Permian Triassic boundary have recently been redated to the end Capitanian Further it is unclear whether some species who survived the prior extinction s had recovered well enough for their final demise in the Permian Triassic event to be considered separate from Capitanian event A minority point of view considers the sequence of environmental disasters to have effectively constituted a single prolonged extinction event perhaps depending on which species is considered Extinction patterns EditMarine extinctions Genera extinct NotesArthropodaEurypterids 100 May have become extinct shortly before the P Tr boundaryOstracods 59 Trilobites 100 In decline since the Devonian only 2 genera living before the extinctionBrachiopodaBrachiopods 96 Orthids and productids died outBryozoaBryozoans 79 Fenestrates trepostomes and cryptostomes died outChordataAcanthodians 100 In decline since the Devonian with only one living familyCnidariaAnthozoans 96 Tabulate and rugose corals died outEchinodermataBlastoids 100 May have become extinct shortly before the P Tr boundaryCrinoids 98 Inadunates and camerates died outMolluscaAmmonites 97 Goniatites died outBivalves 59 Gastropods 98 RetariaForaminiferans 97 Fusulinids died out but were almost extinct before the catastropheRadiolarians 99 60 Marine organisms Edit Marine invertebrates suffered the greatest losses during the P Tr extinction Evidence of this was found in samples from south China sections at the P Tr boundary Here 286 out of 329 marine invertebrate genera disappear within the final two sedimentary zones containing conodonts from the Permian 16 The decrease in diversity was probably caused by a sharp increase in extinctions rather than a decrease in speciation 61 The extinction primarily affected organisms with calcium carbonate skeletons especially those reliant on stable CO2 levels to produce their skeletons 62 These organisms were susceptible to the effects of the ocean acidification that resulted from increased atmospheric CO2 Among benthic organisms the extinction event multiplied background extinction rates and therefore caused maximum species loss to taxa that had a high background extinction rate by implication taxa with a high turnover 63 64 The extinction rate of marine organisms was catastrophic 16 33 65 52 Bioturbators were extremely severely affected as evidenced by the loss of the sedimentary mixed layer in many marine facies during the end Permian extinction 66 Surviving marine invertebrate groups included articulate brachiopods those with a hinge 67 which had undergone a slow decline in numbers since the P Tr extinction the Ceratitida order of ammonites 68 and crinoids sea lilies 68 which very nearly became extinct but later became abundant and diverse The groups with the highest survival rates generally had active control of circulation elaborate gas exchange mechanisms and light calcification more heavily calcified organisms with simpler breathing apparatuses suffered the greatest loss of species diversity 69 70 In the case of the brachiopods at least surviving taxa were generally small rare members of a formerly diverse community 71 The ammonoids which had been in a long term decline for the 30 million years since the Roadian middle Permian suffered a selective extinction pulse 10 million years before the main event at the end of the Capitanian stage In this preliminary extinction which greatly reduced disparity or the range of different ecological guilds environmental factors were apparently responsible Diversity and disparity fell further until the P Tr boundary the extinction here P Tr was non selective consistent with a catastrophic initiator During the Triassic diversity rose rapidly but disparity remained low 72 The range of morphospace occupied by the ammonoids that is their range of possible forms shapes or structures became more restricted as the Permian progressed A few million years into the Triassic the original range of ammonoid structures was once again reoccupied but the parameters were now shared differently among clades 73 The Lilliput effect a term used to describe the phenomenon of dwarfing of species during an immediately following a mass extinction event has been observed across the Permian Triassic boundary 74 75 notably occurring in foraminifera 76 brachiopods 77 78 bivalves 79 80 81 and ostracods 82 It remains debated whether the Lilliput effect took hold among gastropods 83 84 85 Terrestrial invertebrates Edit The Permian had great diversity in insect and other invertebrate species including the largest insects ever to have existed The end Permian is the largest known mass extinction of insects 86 according to some sources it may well be the only mass extinction to significantly affect insect diversity 87 88 Eight or nine insect orders became extinct and ten more were greatly reduced in diversity Palaeodictyopteroids insects with piercing and sucking mouthparts began to decline during the mid Permian these extinctions have been linked to a change in flora The greatest decline occurred in the Late Permian and was probably not directly caused by weather related floral transitions 33 Terrestrial plants Edit The geological record of terrestrial plants is sparse and based mostly on pollen and spore studies Plants are relatively immune to mass extinction with the impact of all the major mass extinctions insignificant at a family level 31 dubious discuss Even the reduction observed in species diversity of 50 may be mostly due to taphonomic processes 89 31 However a massive rearrangement of ecosystems does occur with plant abundances and distributions changing profoundly and all the forests virtually disappearing 90 31 The dominant floral groups changed with many groups of land plants entering abrupt decline such as Cordaites gymnosperms and Glossopteris seed ferns 91 92 the Palaeozoic flora scarcely survived this extinction 93 The Glossopteris dominated flora that characterised high latitude Gondwana collapsed in Australia around 370 000 years before the Permian Triassic boundary with this flora s collapse being less constrained in western Gondwana but still likely occurring a few hundred thousand years before the boundary 94 Palynological or pollen studies from East Greenland of sedimentary rock strata laid down during the extinction period indicate dense gymnosperm woodlands before the event At the same time that marine invertebrate macrofauna declined these large woodlands died out and were followed by a rise in diversity of smaller herbaceous plants including Lycopodiophyta both Selaginellales and Isoetales 45 The Cordaites flora which dominated the Angaran floristic realm corresponding to Siberia collapsed over the course of the extinction 95 In the Kuznetsk Basin the aridity induced extinction of the regions s humid adapted forest flora dominated by cordaitaleans occurred approximately 252 76 Ma around 820 000 years before the end Permian extinction in South China suggesting that the end Permian biotic catastrophe may have started earlier on land and that the ecological crisis may have been more gradual and asynchronous on land compared to its more abrupt onset in the marine realm 96 In South China the subtropical Cathaysian gigantopterid dominated rainforests abruptly collapsed 97 98 95 The floral extinction in South China is associated with bacterial blooms in soil and nearby lacustrine ecosystems with soil erosion resulting from the die off of plants being their likely cause 99 Terrestrial vertebrates Edit There is enough evidence to indicate that over two thirds of terrestrial labyrinthodont amphibians sauropsid reptile and therapsid proto mammal taxa became extinct Large herbivores suffered the heaviest losses All Permian anapsid reptiles died out except the procolophonids although testudines have morphologically anapsid skulls they are now thought to have separately evolved from diapsid ancestors Pelycosaurs died out before the end of the Permian Too few Permian diapsid fossils have been found to support any conclusion about the effect of the Permian extinction on diapsids the reptile group from which lizards snakes crocodilians and dinosaurs including birds evolved 100 9 The groups that survived suffered extremely heavy losses of species and some terrestrial vertebrate groups very nearly became extinct at the end of the Permian Some of the surviving groups did not persist for long past this period but others that barely survived went on to produce diverse and long lasting lineages However it took 30 million years for the terrestrial vertebrate fauna to fully recover both numerically and ecologically 101 It is difficult to analyze extinction and survival rates of land organisms in detail because few terrestrial fossil beds span the Permian Triassic boundary The best known record of vertebrate changes across the Permian Triassic boundary occurs in the Karoo Supergroup of South Africa but statistical analyses have so far not produced clear conclusions 102 Biotic recovery EditFurther information Early Triassic Early Triassic life In the wake of the extinction event the ecological structure of present day biosphere evolved from the stock of surviving taxa In the sea the Modern Evolutionary Fauna became dominant over elements of the Palaeozoic Evolutionary Fauna 103 Typical taxa of shelly benthic faunas were now bivalves snails sea urchins and Malacostraca whereas bony fishes 104 and marine reptiles 105 diversified in the pelagic zone On land dinosaurs and mammals arose in the course of the Triassic The profound change in the taxonomic composition was partly a result of the selectivity of the extinction event which affected some taxa e g brachiopods more severely than others e g bivalves 106 107 However recovery was also differential between taxa Some survivors became extinct some million years after the extinction event without having rediversified dead clade walking 108 e g the snail family Bellerophontidae 109 whereas others rose to dominance over geologic times e g bivalves 110 111 Marine ecosystems Edit Shell bed with the bivalve Claraia clarai a common early Triassic disaster taxon Marine post extinction faunas were mostly species poor and dominated by few disaster species such as the bivalves Claraia and Unionites Seafloor communities maintained a comparatively low diversity until the end of the Early Triassic approximately 4 million years after the extinction event 112 This slow recovery stands in remarkable contrast with the quick recovery seen in nektonic organisms such as ammonoids which exceeded pre extinction diversities already two million years after the crisis 113 The relative delay in the recovery of benthic organisms has been attributed to widespread anoxia 114 but high abundances of benthic species contradict this explanation 115 More recent work suggests that the pace of recovery was intrinsically driven by the intensity of competition among species which drives rates of niche differentiation and speciation 116 Accordingly low levels of interspecific competition in seafloor communities that are dominated by primary consumers correspond to slow rates of diversification and high levels of interspecific competition among nektonic secondary and tertiary consumers to high diversification rates Other studies point to the recurrent episodes of extremely hot climatic conditions during the Early Triassic for the delayed recovery of oceanic life 117 A 2019 study attributed the dissimilarity of recovery times between different ecological communities to differences in local environmental stress during the biotic recovery interval with regions experiencing persistent environmental stress post extinction recovering more slowly 29 Whereas most marine communities were fully recovered by the Middle Triassic 118 119 global marine diversity reached pre extinction values no earlier than the Middle Jurassic approximately 75 million years after the extinction event 120 Sessile filter feeders like this Carboniferous crinoid the mushroom crinoid Agaricocrinus americanus were significantly less abundant after the P Tr extinction Prior to the extinction about two thirds of marine animals were sessile and attached to the seafloor During the Mesozoic only about half of the marine animals were sessile while the rest were free living Analysis of marine fossils from the period indicated a decrease in the abundance of sessile epifaunal suspension feeders such as brachiopods and sea lilies and an increase in more complex mobile species such as snails sea urchins and crabs 121 Before the Permian mass extinction event both complex and simple marine ecosystems were equally common After the recovery from the mass extinction the complex communities outnumbered the simple communities by nearly three to one 121 and the increase in predation pressure led to the Mesozoic Marine Revolution Bivalves were fairly rare before the P Tr extinction but became numerous and diverse in the Triassic taking over niches that were filled primarily by brachiopods before the mass extinction event 122 and one group the rudist clams became the Mesozoic s main reef builders Some researchers think the change was attributable not only to the end Permian extinction but also the ecological restructuring that began as a result of the Capitanian extinction 123 Linguliform brachiopods were commonplace immediately after the extinction event their abundance having been essentially unaffected by the crisis Adaptations for oxygen poor and warm environments such as increased lophophoral cavity surface shell width length ratio and shell miniaturisation are observed in post extinction linguliforms 124 Crinoids sea lilies suffered a selective extinction resulting in a decrease in the variety of their forms 125 Their ensuing adaptive radiation was brisk and resulted in forms possessing flexible arms becoming widespread motility predominantly a response to predation pressure also became far more prevalent 126 Microbial metazoan reefs dominated surviving communities in the immediate wake of the mass extinction Metazoan built reefs reemerged during the Olenekian mainly being composed of sponge biostrome and bivalve builups Tubiphytes dominated reefs appeared at the end of the Olenekian representing the earliest platform margin reefs of the Triassic though they did not become abundant until the late Anisian when reefs species richness increased The first scleractinian corals appear in the late Anisian as well although they would not become the dominant reef builders until the end of the Triassic period 127 Terrestrial plants Edit The proto recovery of terrestrial floras took place from a few tens of thousands of years after the end Permian extinction to around 350 000 years after it with the exact timeline varying by region 128 Dominant gymnosperm genera were replaced post boundary by lycophytes extant lycophytes are recolonizers of disturbed areas 129 The particular post extinction dominance of lycophytes which were well adapted for coastal environments can be explained in part by global marine transgressions during the Early Triassic 95 The worldwide recovery of gymnosperm forests took approximately 4 5 million years 31 However this trend of prolonged lycophyte dominance during the Early Triassic was not universal as evidenced by the much more rapid recovery of gymnosperms in certain regions 130 and floral recovery likely didn t follow a congruent globally universal trend but instead varied by region according to local environmental conditions 98 In East Greenland lycophytes replaced gymnosperms as the dominant plants Later other groups of gymnosperms again become dominant but again suffered major die offs These cyclical flora shifts occurred a few times over the course of the extinction period and afterward These fluctuations of the dominant flora between woody and herbaceous taxa indicate chronic environmental stress resulting in a loss of most large woodland plant species The successions and extinctions of plant communities do not coincide with the shift in d 13C values but occurred many years after 45 In what is now the Barents Sea of the coast of Norway the post extinction fauna is dominated by pteridophytes and lycopods which were suited for primary succession and recolonisation of devastated areas although gymnosperms made a rapid recovery with the lycopod dominated flora not persisting across most of the Early Triassic as postulated in other regions 130 In Europe and North China the interval of recovery was dominated by the lycopsid Pleuromeia an opportunistic pioneer plant that filled ecological vacancies until other plants were able to expand out of refugia and recolonise the land Conifers became common by the early Anisian while pteridosperms and cycadophytes only fully recovered by the late Anisian 131 In southwestern Gondwana the post extinction flora was dominated by bennettitaleans and cycads with memers of Peltaspermales Ginkgoales and Umkomasiales being less common constituents of this flora Around the Induan Olenekian boundary as palaeocommunities recovered a new Dicroidium flora was established in which Umkomasiales continued to be prominent and in which Equisetales and Cycadales were subordinate forms The Dicroidium flora further diversified in the Anisian to its peak wherein Umkomasiales and Ginkgoales constituted most of the tree canopy and Peltaspermales Petriellales Cycadales Umkomasiales Gnetales Equisetales and Dipteridaceae dominated the understory 94 The gigantopterid dominated forests of South China were replaced by low lying herbaceous vegetation dominated by the isoetalean Tomiostrobus following the collapse of the former In contrast to the highly biodiverse gigantopterid rainforests the post extinction landscape of South China was near barren and had vastly lower diversity 98 Coal gap Edit No coal deposits are known from the Early Triassic and those in the Middle Triassic are thin and low grade This coal gap has been explained in many ways It has been suggested that new more aggressive fungi insects and vertebrates evolved and killed vast numbers of trees These decomposers themselves suffered heavy losses of species during the extinction and are not considered a likely cause of the coal gap It could simply be that all coal forming plants were rendered extinct by the P Tr extinction and that it took 10 million years for a new suite of plants to adapt to the moist acid conditions of peat bogs 32 Abiotic factors factors not caused by organisms such as decreased rainfall or increased input of clastic sediments may also be to blame 31 On the other hand the lack of coal may simply reflect the scarcity of all known sediments from the Early Triassic Coal producing ecosystems rather than disappearing may have moved to areas where we have no sedimentary record for the Early Triassic 31 For example in eastern Australia a cold climate had been the norm for a long period with a peat mire ecosystem adapted to these conditions Approximately 95 of these peat producing plants went locally extinct at the P Tr boundary 132 coal deposits in Australia and Antarctica disappear significantly before the P Tr boundary 31 Terrestrial vertebrates Edit Lystrosaurus was by far the most abundant early Triassic land vertebrate Lystrosaurus a pig sized herbivorous dicynodont therapsid constituted as much as 90 of some earliest Triassic land vertebrate fauna Smaller carnivorous cynodont therapsids also survived including the ancestors of mammals In the Karoo region of southern Africa the therocephalians Tetracynodon Moschorhinus and Ictidosuchoides survived but do not appear to have been abundant in the Triassic 133 In North China tetrapod body and ichnofossils are extremely rare in Induan facies but become more abundant in the Olenekian and Anisian showing a biotic recovery of tetrapods synchronous with the decreasing aridity during the Olenekian and Anisian 134 135 Archosaurs which included the ancestors of dinosaurs and crocodilians were initially rarer than therapsids but they began to displace therapsids in the mid Triassic In the mid to late Triassic the dinosaurs evolved from one group of archosaurs and went on to dominate terrestrial ecosystems during the Jurassic and Cretaceous 136 This Triassic Takeover may have contributed to the evolution of mammals by forcing the surviving therapsids and their mammaliform successors to live as small mainly nocturnal insectivores nocturnal life probably forced at least the mammaliforms to develop fur better hearing and higher metabolic rates 137 while losing part of the differential color sensitive retinal receptors reptilians and birds preserved The archosaur dominance would end again due to the Cretaceous Paleogene extinction event after which both birds only extant dinosaurs and mammals only extant synapsids would diversify and share the world Some temnospondyl amphibians made a relatively quick recovery in spite of nearly becoming extinct Mastodonsaurus and trematosaurians were the main aquatic and semiaquatic predators during most of the Triassic some preying on tetrapods and others on fish 138 Land vertebrates took an unusually long time to recover from the P Tr extinction Palaeontologist Michael Benton estimated the recovery was not complete until 30 million years after the extinction i e not until the Late Triassic when the first dinosaurs had risen from bipedal archosaurian ancestors and the first mammals from small cynodont ancestors 12 Coprolitic evidence indicates that freshwater food webs had recovered by the early Ladinian with a lacustrine coprolite assemblage from the Ordos Basin of China providing evidence of a trophically multileveled ecosystem containing at least six different trophic levels The highest trophic levels were filled by vertebrate predators 139 Terrestrial invertebrates Edit Most fossil insect groups found after the Permian Triassic boundary differ significantly from those before Of Paleozoic insect groups only the Glosselytrodea Miomoptera and Protorthoptera have been discovered in deposits from after the extinction The caloneurodeans monurans paleodictyopteroids protelytropterans and protodonates became extinct by the end of the Permian Though Triassic insects are very different from those of the Permian a gap in the insect fossil record spans approximately 15 million years from the late Permian to early Triassic In well documented Late Triassic deposits fossils overwhelmingly consist of modern fossil insect groups 87 Hypotheses about cause EditPinpointing the exact causes of the Permian Triassic extinction event is difficult mostly because it occurred over 250 million years ago and since then much of the evidence that would have pointed to the cause has been destroyed or is concealed deep within the Earth under many layers of rock The sea floor is completely recycled over around 200 million years by the ongoing process of plate tectonics and seafloor spreading leaving no useful indications beneath the ocean Yet scientists have gathered significant evidence for causes and several mechanisms have been proposed The proposals include both catastrophic and gradual processes similar to those theorized for the Cretaceous Paleogene extinction event The catastrophic group includes one or more large bolide impact events increased volcanism and sudden release of methane from the seafloor either due to dissociation of methane hydrate deposits or metabolism of organic carbon deposits by methanogenic microbes The gradual group includes sea level change increasing hypoxia and increasing aridity Any hypothesis about the cause must explain the selectivity of the event which affected organisms with calcium carbonate skeletons most severely the long period 4 to 6 million years before recovery started and the minimal extent of biological mineralization despite inorganic carbonates being deposited once the recovery began 62 Volcanism Edit Siberian Traps Edit The final stages of the Permian had two flood basalt events A smaller one the Emeishan Traps in China occurred at the same time as the end Guadalupian extinction pulse in an area close to the equator at the time 140 141 The flood basalt eruptions that produced the Siberian Traps constituted one of the largest known volcanic events on Earth and covered over 2 000 000 square kilometres 770 000 sq mi with lava 142 143 144 Such a vast aerial extent of the flood basalts may have contributed to their exceptionally catastrophic impact 145 The date of the Siberian Traps eruptions and the extinction event are in good agreement 34 146 The timeline of the extinction event strongly indicates it was caused by events in the large igneous province of the Siberian Traps 147 148 149 150 Carbon dioxide levels prior to and after the eruptions are poorly constrained but may have jumped from between 500 and 4 000 PPM prior to the extinction event to around 8 000 PPM after the extinction 151 As carbon dioxide levels shot up extreme temperature rise would have followed 152 In a 2020 paper scientists reconstructed the mechanisms that led to the extinction event in a biogeochemical model showed the consequences of the greenhouse effect on the marine environment and concluded that the mass extinction can be traced back to volcanic CO2 emissions 153 8 Further evidence based on paired coronene mercury spikes for a volcanic combustion cause of the mass extinction was published in 2020 154 20 The Siberian Traps had unusual features that made them even more dangerous The Siberian Traps eruptions released sulphur rich volatiles that caused dust clouds and the formation of acid aerosols which would have blocked out sunlight and thus disrupted photosynthesis both on land and in the photic zone of the ocean causing food chains to collapse These volcanic outbursts of sulphur also induced brief but severe global cooling leading to glacio eustatic sea level fall 155 The eruptions may also have caused acid rain as the aerosols washed out of the atmosphere That may have killed land plants and mollusks and planktonic organisms which had calcium carbonate shells Pure flood basalts produce fluid low viscosity lava and do not hurl debris into the atmosphere It appears however that 20 of the output of the Siberian Traps eruptions was pyroclastic consisted of ash and other debris thrown high into the atmosphere increasing the short term cooling effect 156 When all of the dust and ash clouds and aerosols washed out of the atmosphere the excess carbon dioxide emitted by the Siberian Traps would have remained and global warming would have proceeded without any mitigating effects 152 The Siberian Traps are underlain by thick sequences of Early Mid Paleozoic aged carbonate and evaporite deposits as well as Carboniferous Permian aged coal bearing clastic rocks When heated such as by igneous intrusions these rocks are capable of emitting large amounts of greenhouse and toxic gases The unique setting of the Siberian Traps over these deposits is likely the reason for the severity of the extinction 157 158 The basalt lava erupted or intruded into carbonate rocks and into sediments that were in the process of forming large coal beds both of which would have emitted large amounts of carbon dioxide leading to stronger global warming after the dust and aerosols settled 152 The timing of the change of the Siberian Traps from flood basalt dominated emplacement to sill dominated emplacement the latter of which would have liberated the largest amounts of trapped hydrocarbon deposits coincides with the onset of the main phase of the mass extinction 159 160 and is linked to a major negative d 13C excursion 161 A 2011 study led by Stephen E Grasby reported evidence that volcanism caused massive coal beds to ignite possibly releasing more than 3 trillion tons of carbon The team found ash deposits in deep rock layers near what is now the Buchanan Lake Formation According to their article coal ash dispersed by the explosive Siberian Trap eruption would be expected to have an associated release of toxic elements in impacted water bodies where fly ash slurries developed Mafic megascale eruptions are long lived events that would allow significant build up of global ash clouds 162 163 In a statement Grasby said In addition to these volcanoes causing fires through coal the ash it spewed was highly toxic and was released in the land and water potentially contributing to the worst extinction event in earth history 164 A 2013 study led by Q Y Yang reported that the total amounts of important volatiles emitted from the Siberian Traps consisted of 8 5 107 Tg CO2 4 4 106 Tg CO 7 0 106 Tg H2S and 6 8 107 Tg SO2 The data support a popular notion that the end Permian mass extinction on the Earth was caused by the emission of enormous amounts of volatiles from the Siberian Traps into the atmosphere 165 In addition these volcanic eruptions released significant quantities of toxic mercury into the atmosphere and ocean further contributing to large scale die offs of terrestrial and marine life 166 167 168 A series of surges in mercury emissions raised environmental mercury concentrations to levels orders of magnitude above normal background levels and caused intervals of extreme environmental toxicity each lasting for over a thousand years in both terrestrial and marine ecosystems 169 Nickel aerosols released by volcanic activity further contributed to metal poisoning 170 Cobalt and arsenic emissions from the Siberian Traps caused further still environmental stress 166 Choiyoi Silicic Large Igneous Province Edit A second flood basalt event that emplaced what is now known as the Choiyoi Silicic Large Igneous Province in southwestern Gondwana between around 286 Ma and 247 Ma has also been suggested as a possible extinction mechanism 94 Being about 1 300 000 cubic kilometres in volume 171 and 1 680 000 square kilometres in area this flood basalt event was approximately 40 the size of the Siberian Traps and thus may have been a significant additional factor explaining the severity of the end Permian extinction 94 Methane clathrate gasification Edit Main article Clathrate gun hypothesis Further information Arctic methane emissionsMethane clathrates also known as methane hydrates consist of methane molecules trapped in cages of water molecules The methane produced by methanogens microscopic single celled organisms has a 13C 12C ratio about 6 0 below normal d 13C 6 0 At the right combination of pressure and temperature the methane is trapped in clathrates fairly close to the surface of permafrost and in much larger quantities on continental shelves and the deeper seabed close to them Oceanic methane hydrates are usually found buried in sediments where the seawater is at least 300 m 980 ft deep They can be found up to about 2 000 m 6 600 ft below the sea floor but usually only about 1 100 m 3 600 ft below the sea floor 172 The release of methane from the clathrates has been considered as a cause because scientists have found worldwide evidence of a swift decrease of about 1 in the 13C 12C isotope ratio in carbonate rocks from the end Permian 52 173 This is the first largest and most rapid of a series of negative and positive excursions decreases and increases in 13C 12C ratio that continues until the isotope ratio abruptly stabilised in the middle Triassic followed soon afterwards by the recovery of calcifying life forms organisms that use calcium carbonate to build hard parts such as shells 69 While a variety of factors may have contributed to this drop in the 13C 12C ratio a 2002 review found most of them to be insufficient to account fully for the observed amount 174 Gases from volcanic eruptions have a 13C 12C ratio about 0 5 to 0 8 below standard d 13C about 0 5 to 0 8 but an assessment made in 1995 concluded that the amount required to produce a reduction of about 1 0 worldwide requires eruptions greater by orders of magnitude than any for which evidence has been found 175 However this analysis addressed only CO2 produced by the magma itself not from interactions with carbon bearing sediments as later proposed A reduction in organic activity would extract 12C more slowly from the environment and leave more of it to be incorporated into sediments thus reducing the 13C 12C ratio Biochemical processes preferentially use the lighter isotopes since chemical reactions are ultimately driven by electromagnetic forces between atoms and lighter isotopes respond more quickly to these forces but a study of a smaller drop of 0 3 to 0 4 in 13C 12C d 13C 3 to 4 at the Paleocene Eocene Thermal Maximum PETM concluded that even transferring all the organic carbon in organisms soils and dissolved in the ocean into sediments would be insufficient Even such a large burial of material rich in 12C would not have produced the smaller drop in the 13C 12C ratio of the rocks around the PETM 175 Buried sedimentary organic matter has a 13C 12C ratio 2 0 to 2 5 below normal d 13C 2 0 to 2 5 Theoretically if the sea level fell sharply shallow marine sediments would be exposed to oxidation But 6 500 8 400 gigatons 1 gigaton 109 metric tons of organic carbon would have to be oxidized and returned to the ocean atmosphere system within less than a few hundred thousand years to reduce the 13C 12C ratio by 1 0 which is not thought to be a realistic possibility 33 Moreover sea levels were rising rather than falling at the time of the extinction 152 Rather than a sudden decline in sea level intermittent periods of ocean bottom hyperoxia and anoxia high oxygen and low or zero oxygen conditions may have caused the 13C 12C ratio fluctuations in the Early Triassic 69 and global anoxia may have been responsible for the end Permian blip The continents of the end Permian and early Triassic were more clustered in the tropics than they are now and large tropical rivers would have dumped sediment into smaller partially enclosed ocean basins at low latitudes Such conditions favor oxic and anoxic episodes oxic anoxic conditions would result in a rapid release burial respectively of large amounts of organic carbon which has a low 13C 12C ratio because biochemical processes use the lighter isotopes more 176 That or another organic based reason may have been responsible for both that and a late Proterozoic Cambrian pattern of fluctuating 13C 12C ratios 69 Other hypotheses include mass oceanic poisoning releasing vast amounts of CO2 177 and a long term reorganisation of the global carbon cycle 174 Prior to consideration of the inclusion of roasting carbonate sediments by volcanism the only proposed mechanism sufficient to cause a global 1 reduction in the 13C 12C ratio was the release of methane from methane clathrates 33 Carbon cycle models confirm that it would have had enough effect to produce the observed reduction 174 177 It was also suggested that a large scale release of methane and other greenhouse gases from the ocean into the atmosphere was connected to the anoxic events and euxinic i e sulfidic events at the time with the exact mechanism compared to the 1986 Lake Nyos disaster 178 The area covered by lava from the Siberian Traps eruptions is about twice as large as was originally thought and most of the additional area was shallow sea at the time The seabed probably contained methane hydrate deposits and the lava caused the deposits to dissociate releasing vast quantities of methane 179 A vast release of methane might cause significant global warming since methane is a very powerful greenhouse gas Strong evidence suggests the global temperatures increased by about 6 C 10 8 F near the equator and therefore by more at higher latitudes a sharp decrease in oxygen isotope ratios 18O 16O 180 the extinction of Glossopteris flora Glossopteris and plants that grew in the same areas which needed a cold climate with its replacement by floras typical of lower paleolatitudes 181 However the pattern of isotope shifts expected to result from a massive release of methane does not match the patterns seen throughout the Early Triassic Not only would such a cause require the release of five times as much methane as postulated for the PETM 69 but would it also have to be reburied at an unrealistically high rate to account for the rapid increases in the 13C 12C ratio episodes of high positive d 13C throughout the early Triassic before it was released several times again 69 The latest research suggests that greenhouse gas release during the extinction event was dominated by volcanic carbon dioxide 182 and while methane release had to have contributed isotopic signatures show that thermogenic methane released from the Siberian Traps had consistently played a larger role than methane from clathrates and any other biogenic sources such as wetlands during the event 183 Hypercapnia and acidification Edit Marine organisms are more sensitive to changes in CO2 carbon dioxide levels than terrestrial organisms for a variety of reasons CO2 is 28 times more soluble in water than is oxygen Marine animals normally function with lower concentrations of CO2 in their bodies than land animals as the removal of CO2 in air breathing animals is impeded by the need for the gas to pass through the respiratory system s membranes lungs alveolus tracheae and the like even when CO2 diffuses more easily than oxygen In marine organisms relatively modest but sustained increases in CO2 concentrations hamper the synthesis of proteins reduce fertilization rates and produce deformities in calcareous hard parts An analysis of marine fossils from the Permian s final Changhsingian stage found that marine organisms with a low tolerance for hypercapnia high concentration of carbon dioxide had high extinction rates and the most tolerant organisms had very slight losses The most vulnerable marine organisms were those that produced calcareous hard parts from calcium carbonate and had low metabolic rates and weak respiratory systems notably calcareous sponges rugose and tabulate corals calcite depositing brachiopods bryozoans and echinoderms about 81 of such genera became extinct Close relatives without calcareous hard parts suffered only minor losses such as sea anemones from which modern corals evolved Animals with high metabolic rates well developed respiratory systems and non calcareous hard parts had negligible losses except for conodonts in which 33 of genera died out This pattern is also consistent with what is known about the effects of hypoxia a shortage but not total absence of oxygen However hypoxia cannot have been the only killing mechanism for marine organisms Nearly all of the continental shelf waters would have had to become severely hypoxic to account for the magnitude of the extinction but such a catastrophe would make it difficult to explain the very selective pattern of the extinction Mathematical models of the Late Permian and Early Triassic atmospheres show a significant but protracted decline in atmospheric oxygen levels with no acceleration near the P Tr boundary Minimum atmospheric oxygen levels in the Early Triassic are never less than present day levels and so the decline in oxygen levels does not match the temporal pattern of the extinction 102 In addition an increase in CO2 concentration is inevitably linked to ocean acidification consistent with the preferential extinction of heavily calcified taxa and other signals in the rock record that suggest a more acidic ocean 184 The decrease in ocean pH is calculated to be up to 0 7 units 185 Ocean acidification was most extreme at mid latitudes and the major marine transgression associated with the end Permian extinction is believed to have devastated shallow shelf communities in conjunction with anoxia 5 Evidence from paralic facies spanning the Permian Triassic boundary in western Guizhou and eastern Yunnan however shows a local marine transgression dominated by carbonate deposition suggesting that ocean acidification did not occur across the entire globe and was likely limited to certain regions of the world s oceans 186 On land the increasing acidification of rainwater caused increased soil erosion as a result of the increased acidity of forest soils evidenced by the increased influx of terrestrially derived organic sediments found in marine sedimentary deposits during the end Permian extinction 187 A positive feedback loop further enhancing and prolonging soil acidification may have resulted from the decline of infaunal invertebrates like tubificids and chironomids which remove acid metabolites from the soil 188 The increased abundance of vermiculitic clays in Shansi South China coinciding with the Permian Triassic boundary strongly suggests a sharp drop in soil pH causally related to volcanogenic emissions of carbon dioxide and sulphur dioxide 189 Anoxia and euxinia Edit See also Anoxic event Evidence for widespread ocean anoxia severe deficiency of oxygen and euxinia presence of hydrogen sulfide is found from the Late Permian to the Early Triassic 190 191 192 Throughout most of the Tethys and Panthalassic Oceans evidence for anoxia including fine laminations in sediments small pyrite framboids high uranium thorium ratios and biomarkers for green sulfur bacteria appear at the extinction event 193 However evidence for anoxia precedes the extinction at some other sites including Spiti India 194 Meishan China 195 Opal Creek Alberta 196 and Kap Stosch Greenland 197 Biomarkers for green sulfur bacteria such as isorenieratane the diagenetic product of isorenieratene are widely used as indicators of photic zone euxinia because green sulfur bacteria require both sunlight and hydrogen sulfide to survive Their abundance in sediments from the P T boundary indicates euxinic conditions were present even in the shallow waters of the photic zone 198 The disproportionate extinction of high latitude marine species provides further evidence for oxygen depletion as a killing mechanism low latitude species living in warmer less oxygenated waters are naturally better adapted to lower levels of oxygen and are able to migrate to higher latitudes during periods of global warming whereas high latitude organisms are unable to escape from warming hypoxic waters at the poles 199 This spread of toxic oxygen depleted water would have devastated marine life causing widespread die offs Models of ocean chemistry suggest that anoxia and euxinia were closely associated with hypercapnia high levels of carbon dioxide 200 This suggests that poisoning from hydrogen sulfide anoxia and hypercapnia acted together as a killing mechanism Hypercapnia best explains the selectivity of the extinction but anoxia and euxinia probably contributed to the high mortality of the event The persistence of anoxia through the Early Triassic may explain the slow recovery of marine life and low levels of biodiversity after the extinction 201 Models also show that anoxic events can cause catastrophic hydrogen sulfide emissions into the atmosphere see below 202 The sequence of events leading to anoxic oceans may have been triggered by carbon dioxide emissions from the eruption of the Siberian Traps 202 In that scenario warming from the enhanced greenhouse effect would reduce the solubility of oxygen in seawater causing the concentration of oxygen to decline Increased weathering of the continents due to warming and the acceleration of the water cycle would increase the riverine flux of phosphate to the ocean The phosphate would have supported greater primary productivity in the surface oceans The increase in organic matter production would have caused more organic matter to sink into the deep ocean where its respiration would further decrease oxygen concentrations Once anoxia became established it would have been sustained by a positive feedback loop because deep water anoxia tends to increase the recycling efficiency of phosphate leading to even higher productivity A severe anoxic event at the end of the Permian would have allowed sulfate reducing bacteria to thrive causing the production of large amounts of hydrogen sulfide in the anoxic ocean turning it euxinic Upwelling of this water may have released massive hydrogen sulfide emissions into the atmosphere and would poison terrestrial plants and animals and severely weaken the ozone layer exposing much of the life that remained to fatal levels of UV radiation 202 Indeed biomarker evidence for anaerobic photosynthesis by Chlorobiaceae green sulfur bacteria from the Late Permian into the Early Triassic indicates that hydrogen sulfide did upwell into shallow waters because these bacteria are restricted to the photic zone and use sulfide as an electron donor The hypothesis has the advantage of explaining the mass extinction of plants which would have added to the methane levels and should otherwise have thrived in an atmosphere with a high level of carbon dioxide Fossil spores from the end Permian further support the theory many spores show deformities that could have been caused by ultraviolet radiation which would have been more intense after hydrogen sulfide emissions weakened the ozone layer 203 Some scientists have challenged the anoxia hypothesis on the grounds that long lasting anoxic conditions could not have been supported if Late Permian thermohaline ocean circulation conformed to the thermal mode characterised by cooling at high latitudes Anoxia may have persisted under a haline mode in which circulation was driven by subtropical evaporation although the haline mode is highly unstable and was unlikely to have represented Late Permian oceanic circulation 204 Aridification Edit Analysis of the fossil river deposits of the floodplains indicate a shift from meandering to braided river patterns indicating a very abrupt drying of the climate 205 The climate change may have taken as little as 100 000 years prompting the extinction of the unique Glossopteris flora and its associated herbivores followed by the carnivorous guild 206 In the North China Basin highly arid climatic conditions are recorded during the latest Permian near the Permian Triassic boundary with a swing towards increased precipitation during the Early Triassic the latter likely assisting biotic recovery following the mass extinction 135 134 Evidence from the Sydney Basin of eastern Australia on the other hand suggests that the expansion of semi arid and arid climatic belts across Pangaea was not immediate but was instead a gradual prolonged process Apart from the disappearance of peatlands there was little evidence of significant sedimentological changes in depositional style across the Permian Triassic boundary 207 Instead a modest shift to amplified seasonality and hotter summers is suggested by palaeoclimatological models based on weathering proxies from the region s Late Permian and Early Triassic deposits 208 In the Kuznetsk Basin of southwestern Siberia an increase in aridity led to the demise of the humid adapted cordaites forests in the region a few hundred thousand years before the Permian Triassic boundary This has been attributed to a broader poleward shift of drier more arid climates during the late Changhsingian before the more abrupt main phase of the extinction at the Permian Triassic boundary that disproportionately affected tropical and subtropical species 96 Asteroid impact Edit Artist s impression of a major impact event A collision between Earth and an asteroid a few kilometers in diameter would release as much energy as the detonation of several million nuclear weapons Evidence that an impact event may have caused the Cretaceous Paleogene extinction has led to speculation that similar impacts may have been the cause of other extinction events including the P Tr extinction and thus to a search for evidence of impacts at the times of other extinctions such as large impact craters of the appropriate age Reported evidence for an impact event from the P Tr boundary level includes rare grains of shocked quartz in Australia and Antarctica 209 210 fullerenes trapping extraterrestrial noble gases 211 meteorite fragments in Antarctica 212 and grains rich in iron nickel and silicon which may have been created by an impact 213 However the accuracy of most of these claims has been challenged 214 215 216 217 For example quartz from Graphite Peak in Antarctica once considered shocked has been re examined by optical and transmission electron microscopy The observed features were concluded to be due not to shock but rather to plastic deformation consistent with formation in a tectonic environment such as volcanism 218 An impact crater on the seafloor would be evidence of a possible cause of the P Tr extinction but such a crater would by now have disappeared As 70 of the Earth s surface is currently sea an asteroid or comet fragment is now perhaps more than twice as likely to hit the ocean as it is to hit land However Earth s oldest ocean floor crust is the only 200 million years old as it is continually being destroyed and renewed by spreading and subduction Furthermore craters produced by very large impacts may be masked by extensive flood basalting from below after the crust is punctured or weakened 219 Yet subduction should not be entirely accepted as an explanation for the lack of evidence as with the K T event an ejecta blanket stratum rich in siderophilic elements such as iridium would be expected in formations from the time A large impact might have triggered other mechanisms of extinction described above 152 such as the Siberian Traps eruptions at either an impact site 220 or the antipode of an impact site 152 221 The abruptness of an impact also explains why more species did not rapidly evolve to survive as would be expected if the Permian Triassic event had been slower and less global than a meteorite impact Bolide impact claims have been criticised on the grounds that they are unnecessary as explanations for the extinctions and they do not fit the known data compatible with a protracted extinction spanning thousands of years 222 Possible impact sites Edit Possible impact craters proposed as the site of an impact causing the P Tr extinction include the 250 km 160 mi Bedout structure off the northwest coast of Australia 210 and the hypothesized 480 km 300 mi Wilkes Land crater of East Antarctica 223 224 An impact has not been proved in either case and the idea has been widely criticized The Wilkes Land geophysical feature is of very uncertain age possibly later than the Permian Triassic extinction Another impact hypothesis postulates that the impact event which formed the Araguainha crater whose formation has been dated to 254 7 2 5 million a possible temporal range overlapping with the end Permian extinction 225 precipitated the mass extinction 226 The impact occurred around extensive deposits of oil shale in the shallow marine Parana Karoo Basin whose perturbation by the seismicity resulting from impact likely discharged about 1 6 teratonnes of methane into Earth s atmosphere buttressing the already rapid warming caused by hydrocarbon release due to the Siberian Traps 227 The large earthquakes generated by the impact would have additionally generated massive tsunamis across much of the globe 226 228 Despite this most palaeontologists reject the impact as being a significant driver of the extinction citing the relatively low energy equivalent to 105 to 106 of TNT around two orders of magnitude lower than the impact energy believed to be required to induce mass extinctions released by the impact 227 A 2017 paper noted the discovery of a circular gravity anomaly near the Falkland Islands which might correspond to an impact crater with a diameter of 250 km 160 mi 229 as supported by seismic and magnetic evidence Estimates for the age of the structure range up to 250 million years old This would be substantially larger than the well known 180 km 110 mi Chicxulub impact crater associated with a later extinction However Dave McCarthy and colleagues from the British Geological Survey illustrated that the gravity anomaly is not circular and also that the seismic data presented by Rocca Rampino and Baez Presser did not cross the proposed crater or provide any evidence for an impact crater 230 Methanogenic microbes Edit A hypothesis published in 2014 posits that a genus of anaerobic methanogenic archaea known as Methanosarcina was responsible for the event 21 Three lines of evidence suggest that these microbes acquired a new metabolic pathway via gene transfer at about that time enabling them to efficiently metabolize acetate into methane That would have led to their exponential reproduction allowing them to rapidly consume vast deposits of organic carbon that had accumulated in the marine sediment The result would have been a sharp buildup of methane and carbon dioxide in the Earth s oceans and atmosphere in a manner that may be consistent with the 13C 12C isotopic record Massive volcanism facilitated this process by releasing large amounts of nickel a scarce metal which is a cofactor for enzymes involved in producing methane 21 On the other hand in the canonical Meishan sections the nickel concentration increases somewhat after the d 13C concentrations have begun to fall 231 Supercontinent Pangaea Edit Map of Pangaea showing where today s continents were at the Permian Triassic boundary In the mid Permian during the Kungurian age of the Permian s Cisuralian epoch Earth s major continental plates joined forming a supercontinent called Pangaea which was surrounded by the superocean Panthalassa Oceanic circulation and atmospheric weather patterns during the mid Permian produced seasonal monsoons near the coasts and an arid climate in the vast continental interior 232 As the supercontinent formed the ecologically diverse and productive coastal areas shrank The shallow aquatic environments were eliminated and exposed formerly protected organisms of the rich continental shelves to increased environmental volatility Pangaea s formation depleted marine life at near catastrophic rates However Pangaea s effect on land extinctions is thought to have been smaller In fact the advance of the therapsids and increase in their diversity is attributed to the late Permian when Pangaea s global effect was thought to have peaked While Pangaea s formation certainly initiated a long period of marine extinction its impact on the Great Dying and the end of the Permian is uncertain Interstellar dust Edit John Gribbin argues that the Solar System last passed through a spiral arm of the Milky Way around 250 million years ago and that the resultant dusty gas clouds may have caused a dimming of the Sun which combined with the effect of Pangaea to produce an ice age 233 Combination of causes Edit Possible causes supported by strong evidence appear to describe a sequence of catastrophes each worse than the last the Siberian Traps eruptions were bad enough alone but because they occurred near coal beds and the continental shelf they also triggered very large releases of carbon dioxide and methane The resultant global warming may have caused perhaps the most severe anoxic event in the oceans history according to this theory the oceans became so anoxic anaerobic sulfur reducing organisms dominated the chemistry of the oceans and caused massive emissions of toxic hydrogen sulfide 102 However there may be some weak links in this chain of events the changes in the 13C 12C ratio expected to result from a massive release of methane do not match the patterns seen throughout the early Triassic 69 and the types of oceanic thermohaline circulation that may have existed at the end of the Permian are not likely to have supported deep sea anoxia 204 See also Edit Evolutionary biology portal Paleontology portalCarbon dioxide Extinction event Climate change List of possible impact structures on Earth Silurian hypothesisReferences Edit a b Rohde R A amp 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 Retrieved 14 January 2023 McLoughlin Steven 8 January 2021 Age and Paleoenvironmental Significance of the Frazer Beach Member A New Lithostratigraphic Unit Overlying the End Permian Extinction Horizon in the Sydney Basin Australia Frontiers in Earth Science 8 600976 605 Bibcode 2021FrEaS 8 605M doi 10 3389 feart 2020 600976 Algeo Thomas J 5 February 2012 The P T Extinction was a Slow Death Astrobiology Magazine Archived from the original on 2021 03 08 a href Template Cite journal html title Template Cite journal cite journal a CS1 maint unfit URL link Li Dirson Jian 18 December 2012 The tectonic cause of mass extinctions and the genomic contribution to biodiversification Quantitative Biology arXiv 1212 4229 Bibcode 2012arXiv1212 4229L a b Beauchamp Benoit Grasby Stephen E 15 September 2012 Permian lysocline shoaling and ocean acidification along NW Pangea led to carbonate eradication and chert expansion Palaeogeography Palaeoclimatology Palaeoecology 350 352 73 90 Bibcode 2012PPP 350 73B doi 10 1016 j palaeo 2012 06 014 Retrieved 21 December 2022 Great Dying lasted 200 000 years National Geographic 23 November 2011 Retrieved 1 April 2014 St Fleur Nicholas 16 February 2017 After Earth s worst mass extinction life rebounded rapidly fossils suggest The New York Times Retrieved 17 February 2017 a b Jurikova Hana Gutjahr Marcus Wallmann Klaus Flogel Sascha Liebetrau Volker Posenato Renato et al November 2020 Permian Triassic mass extinction pulses driven by major marine carbon cycle perturbations Nature Geoscience 13 11 745 750 Bibcode 2020NatGe 13 745J doi 10 1038 s41561 020 00646 4 ISSN 1752 0908 S2CID 224783993 Retrieved 8 November 2020 a b Erwin D H 1990 The End Permian Mass Extinction Annual Review of Ecology and Systematics 21 69 91 doi 10 1146 annurev es 21 110190 000441 a b Chen Yanlong Richoz Sylvain Krystyn Leopold Zhang Zhifei August 2019 Quantitative stratigraphic correlation of Tethyan conodonts across the Smithian Spathian Early Triassic extinction event Earth Science Reviews 195 37 51 Bibcode 2019ESRv 195 37C doi 10 1016 j earscirev 2019 03 004 S2CID 135139479 Retrieved 28 October 2022 Stanley Steven M 2016 10 18 Estimates of the magnitudes of major marine mass extinctions in earth history Proceedings of the National Academy of Sciences 113 42 E6325 E6334 Bibcode 2016PNAS 113E6325S doi 10 1073 pnas 1613094113 ISSN 0027 8424 PMC 5081622 PMID 27698119 a b Benton M J 2005 When Life Nearly Died The greatest mass extinction of all time London Thames amp Hudson ISBN 978 0 500 28573 2 Bergstrom Carl T Dugatkin Lee Alan 2012 Evolution Norton p 515 ISBN 978 0 393 92592 0 a b c d e 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 Yin Hongfu Feng Qinglai Lai Xulong Baud Aymon Tong Jinnan January 2007 The protracted Permo Triassic crisis and multi episode extinction around the Permian Triassic boundary Global and Planetary Change 55 1 3 1 20 Bibcode 2007GPC 55 1Y doi 10 1016 j gloplacha 2006 06 005 Retrieved 29 October 2022 a b c d Jin YG Wang Y Wang W Shang QH Cao CQ Erwin DH 2000 Pattern of marine mass extinction near the Permian Triassic boundary in south China Science 289 5478 432 436 Bibcode 2000Sci 289 432J doi 10 1126 science 289 5478 432 PMID 10903200 Yin H Zhang K Tong J Yang Z Wu S 2001 The global stratotype section and point GSSP of the Permian Triassic boundary Episodes 24 2 102 114 doi 10 18814 epiiugs 2001 v24i2 004 Yin HF Sweets WC Yang ZY Dickins JM 1992 Permo Triassic events in the eastern Tethys an overview In Sweet WC ed Permo Triassic Events in the Eastern Tethys Stratigraphy classification and relations with the western Tethys Cambridge Cambridge University Press pp 1 7 ISBN 978 0 521 54573 0 Darcy E Ogdena amp Norman H Sleep 2011 Explosive eruption of coal and basalt and the end Permian mass extinction Proceedings of the National Academy of Sciences of the United States of America 109 1 59 62 Bibcode 2012PNAS 109 59O doi 10 1073 pnas 1118675109 PMC 3252959 PMID 22184229 a b Kaiho Kunio Aftabuzzaman Md Jones David S Tian Li 4 November 2020 Pulsed volcanic combustion events coincident with the end Permian terrestrial disturbance and the following global crisis Geology 49 3 289 293 doi 10 1130 G48022 1 ISSN 0091 7613 Available under CC BY 4 0 a b c Rothman D H Fournier G P French K L Alm E J Boyle E A Cao C Summons R E 2014 03 31 Methanogenic burst in the end Permian carbon cycle Proceedings of the National Academy of Sciences 111 15 5462 5467 Bibcode 2014PNAS 111 5462R doi 10 1073 pnas 1318106111 PMC 3992638 PMID 24706773 Lay summary Chandler David L March 31 2014 Ancient whodunit may be solved Methane producing microbes did it Science Daily Stanley Steven M 8 September 2009 Evidence from ammonoids and conodonts for multiple Early Triassic mass extinctions Proceedings of the National Academy of Sciences of the United States of America 106 36 15264 15267 Bibcode 2009PNAS 10615264S doi 10 1073 pnas 0907992106 PMC 2741239 PMID 19721005 Hochuli Peter A Sanson Barrera Anna Schneebeli Hermann Elke Bucher Hugo 24 June 2016 Severest crisis overlooked Worst disruption of terrestrial environments postdates the Permian Triassic mass extinction Scientific Reports 6 1 28372 Bibcode 2016NatSR 628372H doi 10 1038 srep28372 PMC 4920029 PMID 27340926 Caravaca Gwenael Thomazo Christophe Vennin Emmanuelle Olivier Nicolas Cocquerez Theophile Escarguel Gilles Fara Emmanuel Jenks James F Bylund Kevin G Stephen Daniel A Brayard Arnaud July 2017 Early Triassic fluctuations of the global carbon cycle New evidence from paired carbon isotopes in the western USA basin Global and Planetary Change 154 10 22 Bibcode 2017GPC 154 10C doi 10 1016 j gloplacha 2017 05 005 S2CID 135330761 Retrieved 13 November 2022 It took Earth ten million years to recover from greatest mass extinction ScienceDaily 27 May 2012 Retrieved 28 May 2012 Brayard Arnaud Krumenacker L J Botting Joseph P Jenks James F Bylund Kevin G Fara Emmanuel Vennin Emmanuelle Olivier Nicolas Goudemand Nicolas Saucede Thomas Charbonnier Sylvain Romano Carlo Doguzhaeva Larisa Thuy Ben Hautmann Michael Stephen Daniel A Thomazo Christophe Escarguel Gilles 15 February 2017 Unexpected Early Triassic marine ecosystem and the rise of the Modern evolutionary fauna Science Advances 13 2 e1602159 Bibcode 2017SciA 3E2159B doi 10 1126 sciadv 1602159 PMC 5310825 PMID 28246643 Smith Christopher P A Laville Thomas Fara Emmanuel Escarguel Gilles Olivier Nicolas Vennin Emmanuelle et al 2021 10 04 Exceptional fossil assemblages confirm the existence of complex Early Triassic ecosystems during the early Spathian Scientific Reports 11 1 19657 Bibcode 2021NatSR 1119657S doi 10 1038 s41598 021 99056 8 ISSN 2045 2322 PMC 8490361 PMID 34608207 Hofmann Richard Goudemand Nicolas Wasmer Martin Bucher Hugo Hautmann Michael 1 October 2011 New trace fossil evidence for an early recovery signal in the aftermath of the end Permian mass extinction Palaeogeography Palaeoclimatology Palaeoecology 310 3 4 216 226 Bibcode 2011PPP 310 216H doi 10 1016 j palaeo 2011 07 014 Retrieved 20 December 2022 a b Woods Adam D Alms Paul D Monarrez Pedro M Mata Scott 1 January 2019 The interaction of recovery and environmental conditions An analysis of the outer shelf edge of western North America during the early Triassic Palaeogeography Palaeoclimatology Palaeoecology 513 52 64 Bibcode 2019PPP 513 52W doi 10 1016 j palaeo 2018 05 014 Retrieved 20 December 2022 a b Payne J L Lehrmann D J Wei J Orchard M J Schrag D P Knoll A H 2004 Large Perturbations of the Carbon Cycle During Recovery from the End Permian Extinction PDF Science 305 5683 506 9 Bibcode 2004Sci 305 506P CiteSeerX 10 1 1 582 9406 doi 10 1126 science 1097023 PMID 15273391 S2CID 35498132 a b c d e f g h i j k l m McElwain J C Punyasena S W 2007 Mass extinction events and the plant fossil record Trends in Ecology amp Evolution 22 10 548 557 doi 10 1016 j tree 2007 09 003 PMID 17919771 a b Retallack G J Veevers J J Morante R 1996 Global coal gap between Permian Triassic extinctions and middle Triassic recovery of peat forming plants GSA Bulletin 108 2 195 207 Bibcode 1996GSAB 108 195R doi 10 1130 0016 7606 1996 108 lt 0195 GCGBPT gt 2 3 CO 2 a b c d e Erwin D H 1993 The great Paleozoic crisis Life and death in the Permian Columbia University Press ISBN 978 0 231 07467 4 a b Burgess S D 2014 High precision timeline for Earth s most severe extinction Proceedings of the National Academy of Sciences of the United States of America 111 9 3316 3321 Bibcode 2014PNAS 111 3316B doi 10 1073 pnas 1317692111 PMC 3948271 PMID 24516148 Magaritz M 1989 13C minima follow extinction events A clue to faunal radiation Geology 17 4 337 340 Bibcode 1989Geo 17 337M doi 10 1130 0091 7613 1989 017 lt 0337 CMFEEA gt 2 3 CO 2 Krull SJ Retallack JR 2000 13C depth profiles from paleosols across the Permian Triassic boundary Evidence for methane release GSA Bulletin 112 9 1459 1472 Bibcode 2000GSAB 112 1459K doi 10 1130 0016 7606 2000 112 lt 1459 CDPFPA gt 2 0 CO 2 ISSN 0016 7606 a b Dolenec T Lojen S Ramovs A 2001 The Permian Triassic boundary in Western Slovenia Idrijca Valley section magnetostratigraphy stable isotopes and elemental variations Chemical Geology 175 1 2 175 190 Bibcode 2001ChGeo 175 175D doi 10 1016 S0009 2541 00 00368 5 Musashi M Isozaki Y Koike T Kreulen R 2001 Stable carbon isotope signature in mid Panthalassa shallow water carbonates across the Permo Triassic boundary Evidence for 13C depleted ocean Earth and Planetary Science Letters 193 1 2 9 20 Bibcode 2001E amp PSL 191 9M doi 10 1016 S0012 821X 01 00398 3 Daily CO2 Mauna Loa Observatory Visscher Henk Looy Cindy V Collinson Margaret E Brinkhuis Henk Cittert Johanna H A van Konijnenburg van Kurschner Wolfram M Sephton Mark A 2004 08 31 Environmental mutagenesis during the end Permian ecological crisis Proceedings of the National Academy of Sciences of the United States of America 101 35 12952 12956 Bibcode 2004PNAS 10112952V doi 10 1073 pnas 0404472101 ISSN 0027 8424 PMC 516500 PMID 15282373 a b Visscher H Brinkhuis H Dilcher DL Elsik WC Eshet Y Looy CW Rampino MR Traverse A 1996 The terminal Paleozoic fungal event Evidence of terrestrial ecosystem destabilization and collapse Proceedings of the National Academy of Sciences 93 5 2155 2158 Bibcode 1996PNAS 93 2155V doi 10 1073 pnas 93 5 2155 PMC 39926 PMID 11607638 Foster C B Stephenson M H Marshall C Logan G A Greenwood P F 2002 A revision of Reduviasporonites Wilson 1962 Description illustration comparison and biological affinities Palynology 26 1 35 58 doi 10 2113 0260035 Diez Jose B Broutin Jean Grauvogel Stamm Lea Borquin Sylvie Bercovici Antoine Ferrer Javier October 2010 Anisian floras from the NE Iberian Peninsula and Balearic Islands A synthesis Review of Palaeobotany and Palynology 162 3 522 542 doi 10 1016 j revpalbo 2010 09 003 Retrieved 8 January 2023 Lopez Gomez J amp Taylor E L 2005 Permian Triassic transition in Spain A multidisciplinary approach Palaeogeography Palaeoclimatology Palaeoecology 229 1 2 1 2 doi 10 1016 j palaeo 2005 06 028 a b c Looy CV Twitchett RJ Dilcher DL van Konijnenburg Van Cittert JH Visscher H 2005 Life in the end Permian dead zone Proceedings of the National Academy of Sciences 98 4 7879 7883 Bibcode 2001PNAS 98 7879L doi 10 1073 pnas 131218098 PMC 35436 PMID 11427710 See image 2 a b Ward PD Botha J Buick R de Kock MO Erwin DH Garrison GH Kirschvink JL Smith R 2005 Abrupt and gradual extinction among late Permian land vertebrates in the Karoo Basin South Africa PDF Science 307 5710 709 714 Bibcode 2005Sci 307 709W CiteSeerX 10 1 1 503 2065 doi 10 1126 science 1107068 PMID 15661973 S2CID 46198018 Retallack G J Smith R M H Ward P D 2003 Vertebrate extinction across Permian Triassic boundary in Karoo Basin South Africa Bulletin of the Geological Society of America 115 9 1133 1152 Bibcode 2003GSAB 115 1133R doi 10 1130 B25215 1 Sephton M A Visscher H Looy C V Verchovsky A B Watson J S 2009 Chemical constitution of a Permian Triassic disaster species Geology 37 10 875 878 Bibcode 2009Geo 37 875S doi 10 1130 G30096A 1 Rampino Michael R Eshet Yoram January 2018 The fungal and acritarch events as time markers for the latest Permian mass extinction An update Geoscience Frontiers 9 1 147 154 doi 10 1016 j gsf 2017 06 005 Retrieved 24 December 2022 Rampino MR Prokoph A Adler A 2000 Tempo of the end Permian event High resolution cyclostratigraphy at the Permian Triassic boundary Geology 28 7 643 646 Bibcode 2000Geo 28 643R doi 10 1130 0091 7613 2000 28 lt 643 TOTEEH gt 2 0 CO 2 ISSN 0091 7613 Wang S C Everson P J 2007 Confidence intervals for pulsed mass extinction events Paleobiology 33 2 324 336 doi 10 1666 06056 1 S2CID 2729020 a b c Twitchett RJ Looy CV Morante R Visscher H Wignall PB 2001 Rapid and synchronous collapse of marine and terrestrial ecosystems during the end Permian biotic crisis Geology 29 4 351 354 Bibcode 2001Geo 29 351T doi 10 1130 0091 7613 2001 029 lt 0351 RASCOM gt 2 0 CO 2 ISSN 0091 7613 Fielding Christopher R Frank Tracy D Savatic Katarina Mays Chris McLoughlin Stephen Vajda Vivi Nicoll Robert S 15 May 2022 Environmental change in the late Permian of Queensland NE Australia The warmup to the end Permian Extinction Palaeogeography Palaeoclimatology Palaeoecology 594 110936 Bibcode 2022PPP 594k0936F doi 10 1016 j palaeo 2022 110936 S2CID 247514266 Retrieved 23 December 2022 a b Retallack G J Metzger C A Greaver T Jahren A H Smith R M H Sheldon N D November December 2006 Middle Late Permian mass extinction on land Bulletin of the Geological Society of America 118 11 12 1398 1411 Bibcode 2006GSAB 118 1398R doi 10 1130 B26011 1 Stanley SM Yang X 1994 A double mass extinction at the end of the Paleozoic Era Science 266 5189 1340 1344 Bibcode 1994Sci 266 1340S doi 10 1126 science 266 5189 1340 PMID 17772839 S2CID 39256134 Ota A amp 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 Shen S amp Shi G R 2002 Paleobiogeographical extinction patterns of Permian brachiopods in the Asian western Pacific region Paleobiology 28 4 449 463 doi 10 1666 0094 8373 2002 028 lt 0449 PEPOPB gt 2 0 CO 2 ISSN 0094 8373 S2CID 35611701 Wang X D amp Sugiyama T December 2000 Diversity and extinction patterns of Permian coral faunas of China Lethaia 33 4 285 294 doi 10 1080 002411600750053853 Bond D P G Wignall P B Wang W Izon G Jiang H S Lai X L et al 2010 The mid Capitanian Middle Permian mass extinction and carbon isotope record of South China Palaeogeography Palaeoclimatology Palaeoecology 292 1 2 282 294 Bibcode 2010PPP 292 282B doi 10 1016 j palaeo 2010 03 056 Racki G 1999 Silica secreting biota and mass extinctions survival processes and patterns Palaeogeography Palaeoclimatology Palaeoecology 154 1 2 107 132 Bibcode 1999PPP 154 107R doi 10 1016 S0031 0182 99 00089 9 Bambach R K Knoll A H Wang S C December 2004 Origination extinction and mass depletions of marine diversity Paleobiology 30 4 522 542 doi 10 1666 0094 8373 2004 030 lt 0522 OEAMDO gt 2 0 CO 2 ISSN 0094 8373 S2CID 17279135 a b Knoll AH 2004 Biomineralization and evolutionary history In Dove PM DeYoreo JJ Weiner S eds Reviews in Mineralogy and Geochemistry PDF Archived from the original PDF on 2010 06 20 Stanley S M 2008 Predation defeats competition on the seafloor Paleobiology 34 1 1 21 doi 10 1666 07026 1 S2CID 83713101 Retrieved 2008 05 13 Stanley S M 2007 An Analysis of the History of Marine Animal Diversity Paleobiology 33 sp6 1 55 doi 10 1666 06020 1 S2CID 86014119 McKinney M L 1987 Taxonomic selectivity and continuous variation in mass and background extinctions of marine taxa Nature 325 6100 143 145 Bibcode 1987Natur 325 143M doi 10 1038 325143a0 S2CID 13473769 Hofmann R Buatois L A MacNaughton R B Mangano M G 15 June 2015 Loss of the sedimentary mixed layer as a result of the end Permian extinction Palaeogeography Palaeoclimatology Palaeoecology 428 1 11 doi 10 1016 j palaeo 2015 03 036 Retrieved 19 December 2022 Permian The Marine Realm and The End Permian Extinction paleobiology si edu Retrieved 2016 01 26 a b Permian extinction Encyclopaedia Britannica Retrieved 2016 01 26 a b c d e f g Payne J L Lehrmann D J Wei J Orchard M J Schrag D P Knoll A H 2004 Large Perturbations of the Carbon Cycle During Recovery from the End Permian Extinction PDF Science 305 5683 506 9 Bibcode 2004Sci 305 506P CiteSeerX 10 1 1 582 9406 doi 10 1126 science 1097023 PMID 15273391 S2CID 35498132 Knoll A H Bambach R K Canfield D E Grotzinger J P 1996 Comparative Earth history and Late Permian mass extinction Science 273 5274 452 457 Bibcode 1996Sci 273 452K doi 10 1126 science 273 5274 452 PMID 8662528 S2CID 35958753 Leighton L R Schneider C L 2008 Taxon characteristics that promote survivorship through the Permian Triassic interval transition from the Paleozoic to the Mesozoic brachiopod fauna Paleobiology 34 1 65 79 doi 10 1666 06082 1 S2CID 86843206 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 Saunders W B Greenfest Allen E Work D M Nikolaeva S V 2008 Morphologic and taxonomic history of Paleozoic ammonoids in time and morphospace Paleobiology 34 1 128 154 doi 10 1666 07053 1 S2CID 83650272 Twitchett Richard J 20 August 2007 The Lilliput effect in the aftermath of the end Permian extinction event Palaeogeography Palaeoclimatology Palaeoecology 252 1 2 132 144 Bibcode 2007PPP 252 132T doi 10 1016 j palaeo 2006 11 038 Retrieved 13 January 2023 Schaal Ellen K Clapham Matthew E Rego Brianna L Wang Steve C Payne Jonathan L 6 November 2015 Comparative size evolution of marine clades from the Late Permian through Middle Triassic Paleobiology 42 1 127 142 doi 10 1017 pab 2015 36 S2CID 18552375 Retrieved 13 January 2023 Song Haijun Tong Jinnan Chen Zhong Qiang 15 July 2011 Evolutionary dynamics of the Permian Triassic foraminifer size Evidence for Lilliput effect in the end Permian mass extinction and its aftermath Palaeogeography Palaeoclimatology Palaeoecology 308 1 2 98 110 Bibcode 2011PPP 308 98S doi 10 1016 j palaeo 2010 10 036 Retrieved 13 January 2023 He Weihong Twitchett Richard J Zhang Y Shi G R Feng Q L Yu J X Wu S B Peng X F 9 November 2010 Controls on body size during the Late Permian mass extinction event Geobiology 8 5 391 402 doi 10 1111 j 1472 4669 2010 00248 x PMID 20550584 S2CID 23096333 Retrieved 13 January 2023 Chen Jing Song Haijun He Weihong Tong Jinnan Wang Fengyu Wu Shunbao 1 April 2019 Size variation of brachiopods from the Late Permian through the Middle Triassic in South China Evidence for the Lilliput Effect following the Permian Triassic extinction Palaeogeography Palaeoclimatology Palaeoecology 519 248 257 Bibcode 2019PPP 519 248C doi 10 1016 j palaeo 2018 07 013 S2CID 134806479 Retrieved 13 January 2023 Huang Yunfei Tong Jinnan Tian Li Song Haijun Chu Daoliang Miao Xue Song Ting 1 January 2023 Temporal shell size variations of bivalves in South China from the Late Permian to the early Middle Triassic Palaeogeography Palaeoclimatology Palaeoecology 609 111307 Bibcode 2023PPP 609k1307H doi 10 1016 j palaeo 2022 111307 S2CID 253368808 Retrieved 13 January 2023 Yates Adam M Neumann Frank H Hancox P John 2 February 2012 The Earliest Post Paleozoic Freshwater Bivalves Preserved in Coprolites from the Karoo Basin South Africa PLOS ONE 7 2 e30228 Bibcode 2012PLoSO 730228Y doi 10 1371 journal pone 0030228 PMC 3271088 PMID 22319562 Foster W J Gliwa J Lembke C Pugh A C Hofmann R Tietje M Varela S Foster L C Korn D Aberhan M January 2020 Evolutionary and ecophenotypic controls on bivalve body size distributions following the end Permian mass extinction Global and Planetary Change 185 103088 Bibcode 2020GPC 18503088F doi 10 1016 j gloplacha 2019 103088 S2CID 213294384 Retrieved 13 January 2023 Chu Daoliang Tong Jinnan Song Haijun Benton Michael James Song Huyue Yu Jianxin Qiu Xincheng Huang Yunfei Tian Li 1 October 2015 Lilliput effect in freshwater ostracods during the Permian Triassic extinction Palaeogeography Palaeoclimatology Palaeoecology 435 38 52 Bibcode 2015PPP 435 38C doi 10 1016 j palaeo 2015 06 003 Retrieved 13 January 2023 Brayard Arnaud Nutzel Alexander Stephen Daniel A Bylund Kevin G Jenks Jim Bucher Hugo 1 February 2010 Gastropod evidence against the Early Triassic Lilliput effect Geology 37 2 147 150 Bibcode 2010Geo 38 147B doi 10 1130 G30553 1 Retrieved 13 January 2023 Fraiser M L Twitchett Richard J Frederickson J A Metcalfe B Bottjer D J 1 January 2011 Gastropod evidence against the Early Triassic Lilliput effect COMMENT Geology 39 1 e232 Bibcode 2011Geo 39E 232F doi 10 1130 G31614C 1 Retrieved 13 January 2023 Brayard Arnaud Nutzel Alexander Kajm Andrzej Escarguel Gilles Hautmann Michael Stephen Daniel A Bylund Kevin G Jenks Jim Bucher Hugo 1 January 2011 Gastropod evidence against the Early Triassic Lilliput effect REPLY Geology 39 1 e233 Bibcode 2011Geo 39E 233B doi 10 1130 G31765Y 1 Retrieved 13 January 2023 Labandeira Conrad 1 January 2005 The fossil record of insect extinction New approaches and future directions American Entomologist 51 14 29 doi 10 1093 ae 51 1 14 a b Labandeira CC Sepkoski JJ 1993 Insect diversity in the fossil record Science 261 5119 310 315 Bibcode 1993Sci 261 310L CiteSeerX 10 1 1 496 1576 doi 10 1126 science 11536548 PMID 11536548 Sole RV Newman M 2003 Extinctions and Biodiversity in the Fossil Record In Canadell JG Mooney HA eds Encyclopedia of Global Environmental Change The Earth System Biological and Ecological Dimensions of Global Environmental Change Vol 2 New York Wiley pp 297 391 ISBN 978 0 470 85361 0 Nowak Hendrik Schneebeli Hermann Elke Kustatscher Evelyn 23 January 2019 No mass extinction for land plants at the Permian Triassic transition Nature Communications 10 1 384 Bibcode 2019NatCo 10 384N doi 10 1038 s41467 018 07945 w PMC 6344494 PMID 30674875 The Dino Directory Natural History Museum Gulbranson Erik L Mellum Morgan M Corti Valentina Dahlseid Aidan Atkinson Brian A Ryberg Patricia E Cornamusini Giancula 24 May 2022 Paleoclimate induced stress on polar forested ecosystems prior to the Permian Triassic mass extinction Scientific Reports 12 1 8702 Bibcode 2022NatSR 12 8702G doi 10 1038 s41598 022 12842 w PMC 9130125 PMID 35610472 Retallack GJ 1995 Permian Triassic life crisis on land Science 267 5194 77 80 Bibcode 1995Sci 267 77R doi 10 1126 science 267 5194 77 PMID 17840061 S2CID 42308183 Cascales Minana B Cleal C J 2011 Plant fossil record and survival analyses Lethaia 45 71 82 doi 10 1111 j 1502 3931 2011 00262 x a b c d Bodnar Josefina Coturel Eliana P Falco Juan Ignacio Beltrana Marisol 4 March 2021 An updated scenario for the end Permian crisis and the recovery of Triassic land flora in Argentina Historical Biology 33 12 3654 3672 doi 10 1080 08912963 2021 1884245 S2CID 233810158 Retrieved 24 December 2022 a b c Vajda Vivi McLoughlin Stephen April 2007 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 1 2 99 112 doi 10 1016 j revpalbo 2005 09 007 Retrieved 24 December 2022 a b Davydov V I Karasev E V Nurgalieva N G Schmitz M D Budnikov I V Biakov A S Kuzina D M Silantiev V V Urazaeva M N Zharinova V V Zorina S O Gareev B Vasilenko D V 1 July 2021 Climate and biotic evolution during the Permian Triassic transition in the temperate Northern Hemisphere Kuznetsk Basin Siberia Russia Palaeogeography Palaeoclimatology Palaeoecology 573 110432 Bibcode 2021PPP 573k0432D doi 10 1016 j palaeo 2021 110432 S2CID 235530804 Retrieved 19 December 2022 Chu Daoliang Yu Jianxin Tong Jinnan Benton Michael James Song Haijun Huang Yunfei Song Ting Tian Li November 2016 Biostratigraphic correlation and mass extinction during the Permian Triassic transition in terrestrial marine siliciclastic settings of South China Global and Planetary Change 146 67 88 Bibcode 2016GPC 146 67C doi 10 1016 j gloplacha 2016 09 009 hdl 1983 2d475883 42ea 4988 b01c 0a56912e6aec Retrieved 12 November 2022 a b c Feng Zhuo Wei Hai Bo Guo Yun He Xiao Yuan Sui Qun Zhou Yu Liu Hang Yu Gou Xu Dong Lv Yong May 2020 From rainforest to herbland New insights into land plant responses to the end Permian mass extinction Earth Science Reviews 204 103153 Bibcode 2020ESRv 20403153F doi 10 1016 j earscirev 2020 103153 S2CID 216433847 Biswas Raman Kumar Kaiho Kunio Saito Ryosuke Tian Li Shi Zhiqiang December 2020 Terrestrial ecosystem collapse and soil erosion before the end Permian marine extinction Organic geochemical evidence from marine and non marine records Global and Planetary Change 195 103327 Bibcode 2020GPC 19503327B doi 10 1016 j gloplacha 2020 103327 S2CID 224900905 Retrieved 21 December 2022 Maxwell W D 1992 Permian and Early Triassic extinction of non marine tetrapods Palaeontology 35 571 583 Bristol University News 2008 Mass extinction a b c Knoll AH Bambach RK Payne JL Pruss S Fischer WW 2007 Paleophysiology and end Permian mass extinction PDF Earth and Planetary Science Letters 256 3 4 295 313 Bibcode 2007E amp PSL 256 295K doi 10 1016 j epsl 2007 02 018 Retrieved 2021 12 13 Sepkoski J John 8 February 2016 A kinetic model of Phanerozoic taxonomic diversity III Post Paleozoic families and mass extinctions Paleobiology 10 2 246 267 doi 10 1017 S0094837300008186 S2CID 85595559 Romano Carlo Koot Martha B Kogan Ilja Brayard Arnaud Minikh Alla V Brinkmann Winand et al February 2016 Permian Triassic Osteichthyes bony fishes diversity dynamics and body size evolution Biological Reviews 91 1 106 147 doi 10 1111 brv 12161 PMID 25431138 S2CID 5332637 Scheyer Torsten M Romano Carlo Jenks Jim Bucher Hugo 19 March 2014 Early Triassic Marine Biotic Recovery The Predators Perspective PLOS ONE 9 3 e88987 Bibcode 2014PLoSO 988987S doi 10 1371 journal pone 0088987 PMC 3960099 PMID 24647136 Komatsu Toshifumi Huyen Dang Tran Jin Hua Chen 1 June 2008 Lower Triassic bivalve assemblages after the end Permian mass extinction in South China and North Vietnam Paleontological Research 12 2 119 128 doi 10 2517 1342 8144 2008 12 119 LTBAAT 2 0 CO 2 S2CID 131144468 Retrieved 6 November 2022 Gould S J Calloway C B 1980 Clams and brachiopodsships that pass in the night Paleobiology 6 4 383 396 doi 10 1017 S0094837300003572 S2CID 132467749 Jablonski D 8 May 2001 Lessons from the past Evolutionary impacts of mass extinctions Proceedings of the National Academy of Sciences 98 10 5393 5398 Bibcode 2001PNAS 98 5393J doi 10 1073 pnas 101092598 PMC 33224 PMID 11344284 Kaim Andrzej Nutzel Alexander July 2011 Dead bellerophontids walking The short Mesozoic history of the Bellerophontoidea Gastropoda Palaeogeography Palaeoclimatology Palaeoecology 308 1 2 190 199 Bibcode 2011PPP 308 190K doi 10 1016 j palaeo 2010 04 008 Hautmann Michael 29 September 2009 The first scallop PDF Palaontologische Zeitschrift 84 2 317 322 doi 10 1007 s12542 009 0041 5 S2CID 84457522 Hautmann Michael Ware David Bucher Hugo August 2017 Geologically oldest oysters were epizoans on Early Triassic ammonoids Journal of Molluscan Studies 83 3 253 260 doi 10 1093 mollus eyx018 Hofmann Richard Hautmann Michael Brayard Arnaud Nutzel Alexander Bylund Kevin G Jenks James F et al May 2014 Recovery of benthic marine communities from the end Permian mass extinction at the low latitudes of eastern Panthalassa PDF Palaeontology 57 3 547 589 doi 10 1111 pala 12076 S2CID 6247479 Brayard A Escarguel G Bucher H Monnet C Bruhwiler T Goudemand N et al 27 August 2009 Good genes and good luck Ammonoid diversity and the end Permian mass extinction Science 325 5944 1118 1121 Bibcode 2009Sci 325 1118B doi 10 1126 science 1174638 PMID 19713525 S2CID 1287762 Wignall P B Twitchett R J 24 May 1996 Oceanic Anoxia and the End Permian Mass Extinction Science 272 5265 1155 1158 Bibcode 1996Sci 272 1155W doi 10 1126 science 272 5265 1155 PMID 8662450 S2CID 35032406 Hofmann Richard Hautmann Michael Bucher Hugo October 2015 Recovery dynamics of benthic marine communities from the Lower Triassic Werfen Formation northern Italy Lethaia 48 4 474 496 doi 10 1111 let 12121 Hautmann Michael Bagherpour Borhan Brosse Morgane Frisk Asa Hofmann Richard Baud Aymon et al September 2015 Competition in slow motion The unusual case of benthic marine communities in the wake of the end Permian mass extinction Palaeontology 58 5 871 901 doi 10 1111 pala 12186 S2CID 140688908 Petsios Elizabeth Thompson Jeffrey R Pietsch Carlie Bottjer David J 1 January 2019 Biotic impacts of temperature before during and after the end Permian extinction A multi metric and multi scale approach to modeling extinction and recovery dynamics Palaeogeography Palaeoclimatology Palaeoecology 513 86 99 Bibcode 2019PPP 513 86P doi 10 1016 j palaeo 2017 08 038 S2CID 134149088 Retrieved 2 December 2022 Friesenbichler Evelyn Hautmann Michael Nutzel Alexander Urlichs Max Bucher Hugo 24 July 2018 Palaeoecology of Late Ladinian Middle Triassic benthic faunas from the Schlern Sciliar and Seiser Alm Alpe di Siusi area South Tyrol Italy PDF PalZ 93 1 1 29 doi 10 1007 s12542 018 0423 7 S2CID 134192673 Friesenbichler Evelyn Hautmann Michael Grădinaru Eugen Bucher Hugo Brayard Arnaud 12 October 2019 A highly diverse bivalve fauna from a Bithynian Anisian Middle Triassic microbial buildup in North Dobrogea Romania PDF Papers in Palaeontology doi 10 1002 spp2 1286 S2CID 208555999 Sepkoski J John 1997 Biodiversity Past Present and Future Journal of Paleontology 71 4 533 539 doi 10 1017 S0022336000040026 PMID 11540302 S2CID 27430390 a b Wagner PJ Kosnik MA Lidgard S 2006 Abundance Distributions Imply Elevated Complexity of Post Paleozoic Marine Ecosystems Science 314 5803 1289 1292 Bibcode 2006Sci 314 1289W doi 10 1126 science 1133795 PMID 17124319 S2CID 26957610 Chen Zhong Qiang Tong Jinnan Liao Zhuo Ting Chen Jing August 2010 Structural changes of marine communities over the Permian Triassic transition Ecologically assessing the end Permian mass extinction and its aftermath Global and Planetary Change 73 1 2 123 140 Bibcode 2010GPC 73 123C doi 10 1016 j gloplacha 2010 03 011 Retrieved 6 November 2022 Clapham ME Bottjer DJ 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 2008 03 28 Posenato Renato Holmer Lars E Prinoth Herwig 1 April 2014 Adaptive strategies and environmental significance of lingulid brachiopods across the late Permian extinction Palaeogeography Palaeoclimatology Palaeoecology 399 373 384 Bibcode 2014PPP 399 373P doi 10 1016 j palaeo 2014 01 028 Retrieved 14 January 2023 Foote M 1999 Morphological diversity in the evolutionary radiation of Paleozoic and post Paleozoic crinoids Paleobiology 25 sp1 1 116 doi 10 1666 0094 8373 1999 25 1 MDITER 2 0 CO 2 ISSN 0094 8373 JSTOR 2666042 S2CID 85586709 Baumiller T K 2008 Crinoid Ecological Morphology Annual Review of Earth and Planetary Sciences 36 1 221 249 Bibcode 2008AREPS 36 221B doi 10 1146 annurev earth 36 031207 124116 Martindale Rowan C Foster William J Velledits Felicitasz 1 January 2019 The survival recovery and diversification of metazoan reef ecosystems following the end Permian mass extinction event Palaeogeography Palaeoclimatology Palaeoecology 513 100 115 Bibcode 2019PPP 513 100M doi 10 1016 j palaeo 2017 08 014 S2CID 135338869 Retrieved 2 December 2022 Aftabuzzaman Md Kaiho Kunio Biswas Raman Kumar Liu Yuqing Saito Ryosuke Tian Li Bhat Ghulam M Chen Zhong Qiang October 2021 End Permian terrestrial disturbance followed by the complete plant devastation and the vegetation proto recovery in the earliest Triassic recorded in coastal sea sediments Global and Planetary Change 205 103621 Bibcode 2021GPC 20503621A doi 10 1016 j gloplacha 2021 103621 Retrieved 24 December 2022 Looy CV Brugman WA Dilcher DL Visscher H 1999 The delayed resurgence of equatorial forests after the Permian Triassic ecologic crisis Proceedings of the National Academy of Sciences of the United States of America 96 24 13857 13862 Bibcode 1999PNAS 9613857L doi 10 1073 pnas 96 24 13857 PMC 24155 PMID 10570163 a b Hochuli Peter A Hermann Elke Vigran Jorunn Os Bucher Hugo Weissert Helmut December 2010 Rapid demise and recovery of plant ecosystems across the end Permian extinction event Global and Planetary Change 74 3 4 144 155 Bibcode 2010GPC 74 144H doi 10 1016 j gloplacha 2010 10 004 Retrieved 24 December 2022 Grauvogel Stamm Lea Ash Sidney R September October 2005 Recovery of the Triassic land flora from the end Permian life crisis Comptes Rendus Palevol 4 6 7 593 608 doi 10 1016 j crpv 2005 07 002 Retrieved 24 December 2022 Michaelsen P 2002 Mass extinction of peat forming plants and the effect on fluvial styles across the Permian Triassic boundary northern Bowen Basin Australia Palaeogeography Palaeoclimatology Palaeoecology 179 3 4 173 188 Bibcode 2002PPP 179 173M doi 10 1016 S0031 0182 01 00413 8 Botha J amp Smith R M H 2007 Lystrosaurus species composition across the Permo Triassic boundary in the Karoo Basin of South Africa PDF Lethaia 40 2 125 137 doi 10 1111 j 1502 3931 2007 00011 x Archived from the original PDF on 2008 09 10 Retrieved 2008 07 02 a b Zhu Zhicai Liu Yongqing Kuang Hongwei Newell Andrew J Peng Nan Cui Mingming Benton Michael J September 2022 Improving paleoenvironment in North China aided Triassic biotic recovery on land following the end Permian mass extinction Global and Planetary Change 216 103914 Bibcode 2022GPC 21603914Z doi 10 1016 j gloplacha 2022 103914 a b Yu Yingyue Tian Li Chu Daoliang Song Huyue Guo Wenwei Tong Jinnan 1 January 2022 Latest Permian Early Triassic paleoclimatic reconstruction by sedimentary and isotopic analyses of paleosols from the Shichuanhe section in central North China Basin Palaeogeography Palaeoclimatology Palaeoecology 585 110726 Bibcode 2022PPP 585k0726Y doi 10 1016 j palaeo 2021 110726 S2CID 239498183 Retrieved 6 November 2022 Benton M J 2004 Vertebrate Paleontology Blackwell Publishers xii 452 ISBN 978 0 632 05614 9 Ruben J A amp Jones T D 2000 Selective Factors Associated with the Origin of Fur and Feathers American Zoologist 40 4 585 596 doi 10 1093 icb 40 4 585 Yates AM Warren AA 2000 The phylogeny of the higher temnospondyls Vertebrata Choanata and its implications for the monophyly and origins of the Stereospondyli Zoological Journal of the Linnean Society 128 1 77 121 doi 10 1111 j 1096 3642 2000 tb00650 x Yao Mingtao Sun Zuoyu Meng Qingqiang Li Jiachun Jiang Dayong 15 August 2022 Vertebrate coprolites from Middle Triassic Chang 7 Member in Ordos Basin China Palaeobiological and palaeoecological implications Palaeogeography Palaeoclimatology Palaeoecology 600 111084 Bibcode 2022PPP 600k1084M doi 10 1016 j palaeo 2022 111084 S2CID 249186414 Retrieved 8 January 2023 Zhou MF Malpas J Song XY Robinson PT Sun M Kennedy AK Lesher CM Keays RR 2002 A temporal link between the Emeishan large igneous province SW China and the end Guadalupian mass extinction Earth and Planetary Science Letters 196 3 4 113 122 Bibcode 2002E amp PSL 196 113Z doi 10 1016 S0012 821X 01 00608 2 Wignall Paul B Sun Y Bond D P G Izon G Newton R J Vedrine S et al 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 Andy Saunders Marc Reichow 2009 The Siberian Traps area and volume Retrieved 2009 10 18 Saunders Andy amp Reichow Marc January 2009 The Siberian Traps and the End Permian mass extinction a critical review PDF Chinese Science Bulletin 54 1 20 37 Bibcode 2009ChSBu 54 20S doi 10 1007 s11434 008 0543 7 hdl 2381 27540 S2CID 1736350 Reichow Marc K Pringle M S Al Mukhamedov A I Allen M B Andreichev V L Buslov M M et al 2009 The timing and extent of the eruption of the Siberian Traps large igneous province Implications for the end Permian environmental crisis PDF Earth and Planetary Science Letters 277 1 2 9 20 Bibcode 2009E amp PSL 277 9R doi 10 1016 j epsl 2008 09 030 hdl 2381 4204 Augland L E Ryabov V V Vernikovsky V A Planke S Polozov A G Callegaro S Jerram D A Svensen H H 10 December 2019 The main pulse of the Siberian Traps expanded in size and composition Scientific Reports 9 1 18723 Bibcode 2019NatSR 918723A doi 10 1038 s41598 019 54023 2 PMC 6904769 PMID 31822688 Kamo SL 2003 Rapid eruption of Siberian flood volcanic rocks and evidence for coincidence with the Permian Triassic boundary and mass extinction at 251 Ma Earth and Planetary Science Letters 214 1 2 75 91 Bibcode 2003E amp PSL 214 75K doi 10 1016 S0012 821X 03 00347 9 Burgess Seth D Bowring Samuel Shen Shu zhong 2014 03 04 High precision timeline for Earth s most severe extinction Proceedings of the National Academy of Sciences 111 9 3316 3321 Bibcode 2014PNAS 111 3316B doi 10 1073 pnas 1317692111 ISSN 0027 8424 PMC 3948271 PMID 24516148 Black Benjamin A Weiss Benjamin P Elkins Tanton Linda T Veselovskiy Roman V Latyshev Anton 2015 04 30 Siberian Traps volcaniclastic rocks and the role of magma water interactions Geological Society of America Bulletin 127 9 10 B31108 1 Bibcode 2015GSAB 127 1437B doi 10 1130 B31108 1 ISSN 0016 7606 Burgess Seth D Bowring Samuel A 2015 08 01 High precision geochronology confirms voluminous magmatism before during and after Earth s most severe extinction Science Advances 1 7 e1500470 Bibcode 2015SciA 1E0470B doi 10 1126 sciadv 1500470 ISSN 2375 2548 PMC 4643808 PMID 26601239 Fischman Josh Giant eruptions and giant extinctions Scientific American video Retrieved 2016 03 11 Cui Ying Kump Lee R October 2015 Global warming and the end Permian extinction event Proxy and modeling perspectives Earth Science Reviews 149 5 22 Bibcode 2015ESRv 149 5C doi 10 1016 j earscirev 2014 04 007 a b c d e f White R V 2002 Earth s biggest whodunnit Unravelling the clues in the case of the end Permian mass extinction PDF Philosophical Transactions of the Royal Society of London 360 1801 2963 2985 Bibcode 2002RSPTA 360 2963W doi 10 1098 rsta 2002 1097 PMID 12626276 S2CID 18078072 Retrieved 2008 01 12 Driver of the largest mass extinction in the history of the Earth identified phys org Retrieved 8 November 2020 Large volcanic eruption caused the largest mass extinction phys org Retrieved 8 December 2020 Baresel Bjorn Bucher Hugo Bagherpour Borhan Brosse Morgane Guodun Kuang Schaltegger Urs 6 March 2017 Timing of global regression and microbial bloom linked with the Permian Triassic boundary mass extinction implications for driving mechanisms Scientific Reports 7 43630 Bibcode 2017NatSR 743630B doi 10 1038 srep43630 PMC 5338007 PMID 28262815 Volcanism Hooper Museum hoopermuseum earthsci carleton ca Ottawa Ontario Canada Carleton University Svensen Henrik Planke Sverre Polozov Alexander G Schmidbauer Norbert Corfu Fernando Podladchikov Yuri Y Jamtveit Bjorn 30 January 2009 Siberian gas venting and the end Permian environmental crisis Earth and Planetary Science Letters 297 3 4 490 500 Bibcode 2009E amp PSL 277 490S doi 10 1016 j epsl 2008 11 015 Retrieved 13 January 2023 Konstantinov Konstantin M Bazhenov Mikhail L Fetisova Anna M Khutorskoy Mikhail D May 2014 Paleomagnetism of trap intrusions East Siberia Implications to flood basalt emplacement and the Permo Triassic crisis of biosphere Earth and Planetary Science Letters 394 242 253 Bibcode 2014E amp PSL 394 242K doi 10 1016 j epsl 2014 03 029 Burgess S D Muirhead J D Bowring S A 31 July 2017 Initial pulse of Siberian Traps sills as the trigger of the end Permian mass extinction Nature Communications 8 1 164 Bibcode 2017NatCo 8 164B doi 10 1038 s41467 017 00083 9 PMC 5537227 PMID 28761160 S2CID 3312150 Corso Jacopo Dal Song Haijun Callegaro Sara Chu Daoliang Sun Yadong Hilton Jason Grasby Stephen E Joachimski Michael M Wignall Paul B 22 February 2022 Environmental crises at the Permian Triassic mass extinction Nature Reviews Earth amp Environment 3 3 197 214 Bibcode 2022NRvEE 3 197D doi 10 1038 s43017 021 00259 4 S2CID 247013868 Retrieved 20 December 2022 Corso Jacopo Dal Mills Benjamin J W Chu Daoling Newton Robert J Mather Tamsin A Shu Wenchao Wu Yuyang Tong Jinnan Wignall Paul B 11 June 2020 Permo Triassic boundary carbon and mercury cycling linked to terrestrial ecosystem collapse Nature Communications 11 1 2962 Bibcode 2020NatCo 11 2962D doi 10 1038 s41467 020 16725 4 PMC 7289894 PMID 32528009 Verango Dan 24 January 2011 Ancient mass extinction tied to torched coal USA Today Grasby Stephen E Sanei Hamed amp Beauchamp Benoit January 23 2011 Catastrophic dispersion of coal fly ash into oceans during the latest Permian extinction Nature Geoscience 4 2 104 107 Bibcode 2011NatGe 4 104G doi 10 1038 ngeo1069 Researchers find smoking gun of world s biggest extinction Massive volcanic eruption burning coal and accelerated greenhouse gas choked out life Press release University of Calgary January 23 2011 Retrieved 2011 01 26 Yang Q Y 2013 The chemical compositions and abundances of volatiles in the Siberian large igneous province Constraints on magmatic CO2 and SO2 emissions into the atmosphere Chemical Geology 339 84 91 Bibcode 2013ChGeo 339 84T doi 10 1016 j chemgeo 2012 08 031 a b Grasby Stephen E Beauchamp Benoit Bond David P G Wignall Paul B Talavera Cristina Galloway Jennifer M Piepjohn Karsten Reinhardt Lutz Blomeier Dirk 1 September 2015 Progressive environmental deterioration in northwestern Pangea leading to the latest Permian extinction Geological Society of America Bulletin 127 9 10 1331 1347 Bibcode 2015GSAB 127 1331G doi 10 1130 B31197 1 Retrieved 14 January 2023 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 Sanei Hamed Grasby Stephen E Beauchamp Benoit 1 January 2012 Latest Permian mercury anomalies Geology 40 1 63 66 Bibcode 2012Geo 40 63S doi 10 1130 G32596 1 Retrieved 14 January 2023 Grasby Stephen E Liu Xiaojun Yin Runsheng Ernst Richard E Chen Zhuoheng 19 May 2020 Toxic mercury pulses into late Permian terrestrial and marine environments Geology 48 8 830 833 Bibcode 2020Geo 48 830G doi 10 1130 G47295 1 S2CID 219495628 Retrieved 14 January 2023 Li Menghan Grasby Stephen E Wang Shui Jiong Zhang Xiaolin Wasylenki Laura E Xu Yilun Sun Mingzhao Beauchamp Benoit Hu Dongping Shen Yanan 1 April 2021 Nickel isotopes link Siberian Traps aerosol particles to the end Permian mass extinction Nature Communications 12 1 2024 Bibcode 2021NatCo 12 2024L doi 10 1038 s41467 021 22066 7 PMC 8016954 PMID 33795666 Nelson D A Cottle J M 29 March 2019 Tracking voluminous Permian volcanism of the Choiyoi Province into central Antarctica Lithosphere 11 3 386 398 Bibcode 2019Lsphe 11 386N doi 10 1130 L1015 1 S2CID 135130436 Retrieved 14 January 2023 Dickens G R 2001 The potential volume of oceanic methane hydrates with variable external conditions Organic Geochemistry 32 10 1179 1193 doi 10 1016 S0146 6380 01 00086 9 Palfy J Demeny A Haas J Htenyi M Orchard MJ Veto I 2001 Carbon isotope anomaly at the Triassic Jurassic boundary from a marine section in Hungary Geology 29 11 1047 1050 Bibcode 2001Geo 29 1047P doi 10 1130 0091 7613 2001 029 lt 1047 CIAAOG gt 2 0 CO 2 ISSN 0091 7613 a b c Berner R A 2002 Examination of hypotheses for the Permo Triassic boundary extinction by carbon cycle modeling Proceedings of the National Academy of Sciences 99 7 4172 4177 Bibcode 2002PNAS 99 4172B doi 10 1073 pnas 032095199 PMC 123621 PMID 11917102 a b Dickens GR O Neil JR Rea DK Owen RM 1995 Dissociation of oceanic methane hydrate as a cause of the carbon isotope excursion at the end of the Paleocene Paleoceanography 10 6 965 971 Bibcode 1995PalOc 10 965D doi 10 1029 95PA02087 Schrag DP Berner RA Hoffman PF Halverson GP 2002 On the initiation of a snowball Earth Geochemistry Geophysics Geosystems 3 6 1 21 Bibcode 2002GGG 3fQ 1S doi 10 1029 2001GC000219 Preliminary abstract at Schrag D P June 2001 On the initiation of a snowball Earth Geological Society of America Archived from the original on 2018 04 25 Retrieved 2008 04 20 a b Benton M J Twitchett R J 2003 How to kill almost all life The end Permian extinction event Trends in Ecology amp Evolution 18 7 358 365 doi 10 1016 S0169 5347 03 00093 4 Ryskin Gregory September 2003 Methane driven oceanic eruptions and mass extinctions Geology 31 9 741 744 Bibcode 2003Geo 31 741R doi 10 1130 G19518 1 Reichow MK Saunders AD White RV Pringle MS Al Muhkhamedov AI Medvedev AI Kirda NP 2002 40Ar 39Ar dates from the West Siberian Basin Siberian flood basalt province doubled PDF Science 296 5574 1846 1849 Bibcode 2002Sci 296 1846R doi 10 1126 science 1071671 PMID 12052954 S2CID 28964473 Holser WT Schoenlaub HP Attrep Jr M Boeckelmann K Klein P Magaritz M Orth CJ Fenninger A Jenny C Kralik M Mauritsch H Pak E Schramm JF Stattegger K Schmoeller R 1989 A unique geochemical record at the Permian Triassic boundary Nature 337 6202 39 44 Bibcode 1989Natur 337 39H doi 10 1038 337039a0 S2CID 8035040 Dobruskina I A 1987 Phytogeography of Eurasia during the early Triassic Palaeogeography Palaeoclimatology Palaeoecology 58 1 2 75 86 Bibcode 1987PPP 58 75D doi 10 1016 0031 0182 87 90007 1 Cui Ying Li Mingsong van Soelen Elsbeth E Peterse Francien M Kurschner Wolfram 7 September 2021 Massive and rapid predominantly volcanic CO2 emission during the end Permian mass extinction Earth Atmospheric and Planetary Sciences 118 37 e2014701118 Bibcode 2021PNAS 11814701C doi 10 1073 pnas 2014701118 PMC 8449420 PMID 34493684 Wu Yuyang Chu Daoliang Tong Jinnan Song Haijun Dal Corso Jacopo Wignall Paul B Song Huyue Du Yong Cui Ying 9 April 2021 Six fold increase of atmospheric pCO2 during the Permian Triassic mass extinction Nature Communications 12 1 2137 Bibcode 2021NatCo 12 2137W doi 10 1038 s41467 021 22298 7 PMC 8035180 PMID 33837195 S2CID 233200774 Payne J Turchyn A Paytan A Depaolo D Lehrmann D Yu M Wei J 2010 Calcium isotope constraints on the end Permian mass extinction Proceedings of the National Academy of Sciences of the United States of America 107 19 8543 8548 Bibcode 2010PNAS 107 8543P doi 10 1073 pnas 0914065107 PMC 2889361 PMID 20421502 Clarkson M Kasemann S Wood R Lenton T Daines S Richoz S et al 2015 04 10 Ocean acidification and the Permo Triassic mass extinction PDF Science 348 6231 229 232 Bibcode 2015Sci 348 229C doi 10 1126 science aaa0193 hdl 10871 20741 PMID 25859043 S2CID 28891777 Wignall Paul B Chu Daoliang Hilton Jason M Corso Jacopo Dal Wu Yuyang Wang Yao Atkinson Jed Tong Jinnan June 2020 Death in the shallows The record of Permo Triassic mass extinction in paralic settings southwest China Global and Planetary Change 189 103176 Bibcode 2020GPC 18903176W doi 10 1016 j gloplacha 2020 103176 S2CID 216302513 Retrieved 7 November 2022 Sephton Mark A Jiao Dan Engel Michael H Looy Cindy V Visscher Henk 1 February 2015 Terrestrial acidification during the end Permian biosphere crisis Geology 43 2 159 162 Bibcode 2015Geo 43 159S doi 10 1130 G36227 1 Retrieved 23 December 2022 Buatois Luis A Borruel Abadia Violeta De la Horra Raul Galan Abellan Ana Belen Lopez Gomez Jose Barrenechea Jose 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 Xu Guozhen Deconinck Jean Francois Feng Qinglai Baudin Francois Pellenard Pierre Shen Jun Bruneau Ludovic 15 May 2017 Clay mineralogical characteristics at the Permian Triassic Shangsi section and their paleoenvironmental and or paleoclimatic significance Palaeogeography Palaeoclimatology Palaeoecology 474 152 163 Bibcode 2017PPP 474 152X doi 10 1016 j palaeo 2016 07 036 Retrieved 23 December 2022 Song Haijun Wignall Paul B Tong Jinnan Bond David P G Song Huyue Lai Xulong Zhang Kexin Wang Hongmei Chen Yanlong 1 November 2012 Geochemical evidence from bio apatite for multiple oceanic anoxic events during Permian Triassic transition and the link with end Permian extinction and recovery Earth and Planetary Science Letters 353 354 12 21 Bibcode 2012E amp PSL 353 12S doi 10 1016 j epsl 2012 07 005 Retrieved 13 January 2023 Hulse Dominik Lau Kimberly V Van de Velde Sebastiaan J Arndt Sandra Meyer Katja M Ridgwell Andy 28 October 2021 End Permian marine extinction due to temperature driven nutrient recycling and euxinia Nature Geoscience 14 11 862 867 Bibcode 2021NatGe 14 862H doi 10 1038 s41561 021 00829 7 S2CID 240076553 Retrieved 8 January 2023 Schobben Martin Foster William J Sleveland Arve R N Zuchuat Valentin Svensen Henrik H Planke Sverre Bond David P G Marcelis Fons Newton Robert J Wignall Paul B Poulton Simon W 17 August 2020 A nutrient control on marine anoxia during the end Permian mass extinction Nature Geoscience 13 9 640 646 Bibcode 2020NatGe 13 640S doi 10 1038 s41561 020 0622 1 S2CID 221146234 Retrieved 8 January 2023 span, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.