Permian–Triassic extinction event

Source: Wikipedia, the free encyclopedia.
CambrianOrdovicianSilurianDevonianCarboniferousPermianTriassicJurassicCretaceousPaleogeneNeogene
Marine extinction intensity during 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]

Approximately 251.9 million years ago, the Permian–Triassic (P–T, P–Tr) extinction event (PTME; also known as the Late Permian extinction event,

insects.[16] It is the greatest of the "Big Five" mass extinctions of the Phanerozoic.[17] There is evidence for one to three distinct pulses, or phases, of extinction.[15][18]

The precise causes of the Great Dying remain unknown. The scientific consensus is that the main cause of extinction was the flood basalt volcanic eruptions that created the Siberian Traps,[19] which released sulfur dioxide and carbon dioxide, resulting in euxinia,[20][21] elevating global temperatures,[22][23][24] and acidifying the oceans.[25][26][3] The level of atmospheric carbon dioxide rose from around 400 ppm to 2,500 ppm with approximately 3,900 to 12,000 gigatonnes of carbon being added to the ocean-atmosphere system during this period.[22] Important 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,[27][28] emissions of methane from the gasification of methane clathrates,[29] emissions of methane possibly by novel methanogenic microorganisms nourished by minerals dispersed in the eruptions,[30][31][32] an extraterrestrial impact creating the Araguainha crater and consequent seismic release of methane,[33][34][35] and the destruction of the ozone layer and increase in harmful solar radiation.[36][37][38]

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.[44] 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 first appearance of the conodont Hindeodus parvus has been used to delineate the Permian-Triassic boundary.[45][46]

The extinction occurred between 251.941 ± 0.037 and 251.880 ± 0.031 million years ago, a duration of 60 ± 48 thousand years.

carbon-13 to that of carbon-12, coincides with this extinction,[48][49][50] and is sometimes used to identify the Permian–Triassic boundary in rocks that are unsuitable for radiometric dating.[51] The negative carbon isotope excursion's magnitude was 4-7% and lasted for approximately 500 kyr,[52] though estimating its exact value is challenging due to diagenetic alteration of many sedimentary facies spanning the boundary.[53][54]

Further evidence for environmental change around the Permian-Triassic boundary suggests an 8 °C (14 °F) rise in temperature,

ppm (for comparison, the concentration immediately before the Industrial Revolution was 280 ppm,[42] and the amount today is about 415 ppm[55]). There is also evidence of increased ultraviolet radiation reaching the earth, causing the mutation of plant spores.[42][38]

It has been suggested that the Permian–Triassic boundary is associated with a sharp increase in the abundance of marine and terrestrial

alga;[42][58] the spike did not appear worldwide;[59][60][61] and in many places it did not fall on the Permian–Triassic boundary.[62] 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.[63] Newer chemical evidence agrees better with a fungal origin for Reduviasporonites, diluting these critiques.[64][65]

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[66][67][15] 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.[68][62][69] Statistical analyses of some highly fossiliferous strata in Meishan, Zhejiang Province in southeastern China, suggest that the main extinction was clustered around one peak,[18] while a study of the Liangfengya section found evidence of two extinction waves, MEH-1 and MEH-2, which varied in their causes,[70] and a study of the Shangsi section showed two extinction pulses with different causes too.[71] 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.[72] Palaeoenvironmental analysis of Lopingian strata in the Bowen Basin of Queensland indicates numerous intermittent periods of marine environmental stress from the middle to late Lopingian leading up to the end-Permian extinction proper, supporting aspects of the gradualist hypothesis.[73] Additionally, the decline in marine species richness and the structural collapse of marine ecosystems may have been decoupled as well, with the former preceding the latter by about 61,000 years according to one study.[74]

Whether the terrestrial and marine extinctions were synchronous or asynchronous is another point of controversy. Evidence from a well-preserved sequence in east Greenland suggests that the terrestrial and marine extinctions began simultaneously. In this sequence, 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.[75] Many sedimentary sequences from South China show synchronous terrestrial and marine extinctions.[76] Research in the Sydney Basin of the PTME's duration and course also supports a synchronous occurrence of the terrestrial and marine biotic collapses.[77] Other scientists believe the terrestrial mass extinction began between 60,000 and 370,000 years before the onset of the marine mass extinction.[78][79] Chemostratigraphic analysis from sections in Finnmark and Trøndelag shows the terrestrial floral turnover occurred before the large negative δ13C shift during the marine extinction.[80] Dating of the boundary between the Dicynodon and Lystrosaurus assemblage zones in the Karoo Basin indicates that the terrestrial extinction occurred earlier than the marine extinction.[81] The Sunjiagou Formation of South China also records a terrestrial ecosystem demise predating the marine crisis.[82] Other research still has found that the terrestrial extinction occurred after the marine extinction in the tropics.[83]

Studies of the timing and causes of the Permian-Triassic extinction are complicated by the often-overlooked

fusuline foraminifera.[88] 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.[89][90]

Extinction patterns

Marine extinctions Genera extinct Notes
Arthropoda
Eurypterids 100% May have become extinct shortly before the P–Tr boundary
Ostracods 74%  
Trilobites 100% In decline since the Devonian; only 5 genera living before the extinction
Brachiopoda
Brachiopods 96% Orthids, Orthotetids 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 and Prolecantids died out
Bivalves 59%  
Gastropods
98%  
Retaria
Foraminiferans 97%
Fusulinids
died out, but were almost extinct before the catastrophe
Radiolarians 99%[91]

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.[18] The decrease in diversity was probably caused by a sharp increase in extinctions, rather than a decrease in speciation.[92]

The extinction primarily affected organisms with calcium carbonate skeletons, especially those reliant on stable CO2 levels to produce their skeletons.[93] These organisms were susceptible to the effects of the ocean acidification that resulted from increased atmospheric CO2. There is also evidence that endemism was a strong risk factor influencing a taxon's likelihood of extinction. Bivalve taxa that were endemic and localised to a specific region were more likely to go extinct than cosmopolitan taxa.[94] There was little latitudinal difference in the survival rates of taxa.[95]

Among

turnover).[96][97] The extinction rate of marine organisms was catastrophic.[18][44][98][75] Bioturbators were extremely severely affected, as evidenced by the loss of the sedimentary mixed layer in many marine facies during the end-Permian extinction.[99]

Surviving marine invertebrate groups included articulate

ammonites;[101] and crinoids ("sea lilies"),[101] 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.[102][103] In the case of the brachiopods, at least, surviving taxa were generally small, rare members of a formerly diverse community.[104]

The

Ostracods experienced prolonged diversity perturbations during the Changhsingian before the PTME proper, when immense proportions of them abruptly vanished.[107] At least 74% of ostracods died out during the PTME itself.[108]

Bryozoans had been on a long-term decline throughout the Late Permian epoch before they suffered even more catastrophic losses during the PTME.[109]

Deep water sponges suffered a significant diversity loss and exhibited a decrease in spicule size over the course of the PTME. Shallow water sponges were affected much less strongly; they experienced an increase in spicule size and much lower loss of morphological diversity compared to their deep water counterparts.[110]

Foraminifera suffered a severe bottleneck in diversity.[5] Approximately 93% of latest Permian foraminifera became extinct, with 50% of the clade Textulariina, 92% of Lagenida, 96% of Fusulinida, and 100% of Miliolida disappearing.[111] The reason why lagenides survived while fusulinoidean fusulinides went completely extinct may have been due to the greater range of environmental tolerance and greater geographic distribution of the former compared to the latter.[112]

Cladodontomorph sharks likely survived the extinction because of their ability to survive in refugia in the deep oceans. This hypothesis is based on the discovery of

Ichthyosaurs, which are believed to have evolved immediately before the PTME, were also PTME survivors.[114]

The Lilliput effect, the phenomenon of dwarfing of species during and immediately following a mass extinction event, has been observed across the Permian-Triassic boundary,[115][116] notably occurring in foraminifera,[117][118][119] brachiopods,[120][121][122] bivalves,[123][124][125] and ostracods.[126][127] Though gastropods that survived the cataclysm were smaller in size than those that did not,[128] it remains debated whether the Lilliput effect truly took hold among gastropods.[129][130][131] Some gastropod taxa, termed "Gulliver gastropods", ballooned in size during and immediately following the mass extinction,[132] exemplifying the Lilliput effect's opposite, which has been dubbed the Brobdingnag effect.[133]

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;[16] according to some sources, it may well be the only mass extinction to significantly affect insect diversity.[134][135] 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.[44]

Terrestrial plants

The geological record of terrestrial plants is sparse and based mostly on pollen and spore studies. Floral changes across the Permian-Triassic boundary are highly variable depending on the location and preservation quality of any given site.[136] Plants are relatively immune to mass extinction, with the impact of all the major mass extinctions "insignificant" at a family level.[42][dubious ] Even the reduction observed in species diversity (of 50%) may be mostly due to taphonomic processes.[137][42] However, a massive rearrangement of ecosystems does occur, with plant abundances and distributions changing profoundly and all the forests virtually disappearing.[138][42] The dominant floral groups changed, with many groups of land plants entering abrupt decline, such as Cordaites (gymnosperms) and Glossopteris (seed ferns).[139][140] The severity of plant extinction has been disputed.[141][137]

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.[142]

Selaginellales and Isoetales.[61]

The Cordaites flora, which dominated the Angaran floristic realm corresponding to Siberia, collapsed over the course of the extinction.[143] 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.[144]

In North China, the transition between the Upper Shihhotse and Sunjiagou Formations and their lateral equivalents marked a very large extinction of plants in the region. Those plant genera that did not go extinct still experienced a great reduction in their geographic range. Following this transition, coal swamps vanished. The North Chinese floral extinction correlates with the decline of the Gigantopteris flora of South China.[145]

In South China, the subtropical Cathaysian gigantopterid dominated rainforests abruptly collapsed.[146][147][148] 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.[149] Wildfires too likely played a role in the fall of Gigantopteris.[150]

A conifer flora in what is now Jordan, known from fossils near the Dead Sea, showed unusual stability over the Permian-Triassic transition, and appears to have been only minimally affected by the crisis.[151]

Terrestrial vertebrates

The terrestrial vertebrate extinction occurred rapidly, taking 50,000 years or less.

taxa became extinct. Large herbivores
suffered the heaviest losses.

All Permian

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).[155][10]

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.[156]

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.[157] One study of the Karoo Basin found that 69% of terrestrial vertebrates went extinct over 300,000 years leading up to the Permian-Triassic boundary, followed by a minor extinction pulse involving four taxa that survived the previous extinction interval.[158] Another study of latest Permian vertebrates in the Karoo Basin found that 54% of them went extinct due to the PTME.[159]

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 "Palaeozoic evolutionary fauna" declined while the "modern evolutionary fauna" achieved greater dominance;

dead clade walking,[167] e.g. the snail family Bellerophontidae),[168] whereas others rose to dominance over geologic times (e.g., bivalves).[169][170]

Marine ecosystems

Claraia clarai
, a common early Triassic disaster taxon

A cosmopolitanism event began immediately after the end-Permian extinction event.[171] Marine post-extinction faunas were mostly species-poor and were dominated by few disaster taxa such as the bivalves Claraia, Unionites, Eumorphotis, and Promyalina,[172] the conodonts Clarkina and Hindeodus,[173] the inarticulate brachiopod Lingularia,[172] and the foraminifera Earlandia and Rectocornuspira kalhori,[174] the latter of which is sometimes classified under the genus Ammodiscus.[175] Their guild diversity was also low.[176]

The speed of recovery from the extinction is disputed. Some scientists estimate that it took 10 million years (until the Middle Triassic)[177] due to the severity of the extinction. However, studies in Bear Lake County, near Paris, Idaho,[178] and nearby sites in Idaho and Nevada[179] showed a relatively quick rebound in a localized Early Triassic marine ecosystem (Paris biota), taking around 1.3 million years to recover,[178] while an unusually diverse and complex ichnobiota is known from Italy less than a million years after the end-Permian extinction.[180] Additionally, the complex Guiyang biota found near Guiyang, China also indicates life thrived in some places just a million years after the mass extinction,[181][182] as does a fossil assemblage known as the Shanggan fauna found in Shanggan, China[183] and a gastropod fauna from the Al Jil Formation of Oman.[184] Regional differences in the pace of biotic recovery existed,[185] which suggests that the impact of the extinction may have been felt less severely in some areas than others, with differential environmental stress and instability being the source of the variance.[186][187] In addition, it has been proposed that although overall taxonomic diversity rebounded rapidly, functional ecological diversity took much longer to return to its pre-extinction levels;[188] one study concluded that marine ecological recovery was still ongoing 50 million years after the extinction, during the latest Triassic, even though taxonomic diversity had rebounded in a tenth of that time.[189]

The pace and timing of recovery also differed based on clade and mode of life. Seafloor communities maintained a comparatively low diversity until the end of the Early Triassic, approximately 4 million years after the extinction event.

ammonoids, which exceeded pre-extinction diversities already two million years after the crisis,[192] and conodonts, which diversified considerably over the first two million years of the Early Triassic.[193]

Recent work suggests that the pace of recovery was intrinsically driven by the intensity of competition among species, which drives rates of

Smithian-Spathian boundary extinction.[11][199][200] Continual episodes of extremely hot climatic conditions during the Early Triassic have been held responsible for the delayed recovery of oceanic life,[201][202] in particular skeletonised taxa that are most vulnerable to high carbon dioxide concentrations.[203] The relative delay in the recovery of benthic organisms has been attributed to widespread anoxia,[204] but high abundances of benthic species contradict this explanation.[205] 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, supporting the view that recurrent environmental calamities were culpable for retarded biotic recovery.[187] Recurrent Early Triassic environmental stresses also acted as a ceiling limiting the maximum ecological complexity of marine ecosystems until the Spathian.[206] Recovery biotas appear to have been ecologically uneven and unstable into the Anisian, making them vulnerable to environmental stresses.[207]

Whereas most marine communities were fully recovered by the Middle Triassic,[208][209] global marine diversity reached pre-extinction values no earlier than the Middle Jurassic, approximately 75 million years after the extinction event.[210]

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

Mesozoic Marine Revolution.[212]

Bivalves rapidly recolonised many marine environments in the wake of the catastrophe.

Bayesian analysis.[215] The success of bivalves in the aftermath of the extinction event may have been a function of them possessing greater resilience to environmental stress compared to the brachiopods that they coexisted with.[162] The rise of bivalves to taxonomic and ecological dominance over brachiopods was not synchronous, however, and brachiopods retained an outsized ecological dominance into the Middle Triassic even as bivalves eclipsed them in taxonomic diversity.[216] Some researchers think the brachiopod-bivalve transition was attributable not only to the end-Permian extinction but also the ecological restructuring that began as a result of the Capitanian extinction.[217] Infaunal habits in bivalves became more common after the PTME.[218]

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.[219] The surviving brachiopod fauna was very low in diversity and exhibited no provincialism whatsoever.[220] Brachiopods began their recovery around 250.1 ± 0.3 Ma, as marked by the appearance of the genus Meishanorhynchia, believed to be the first of the progenitor brachiopods that evolved after the mass extinction.[221] Major brachiopod rediversification only began in the late Spathian and Anisian in conjunction with the decline of widespread anoxia and extreme heat and the expansion of more habitable climatic zones.[222] Brachiopod taxa during the Anisian recovery interval were only phylogenetically related to Late Permian brachiopods at a familial taxonomic level or higher; the ecology of brachiopods had radically changed from before in the mass extinction's aftermath.[223]

Ostracods were extremely rare during the basalmost Early Triassic.[224] Taxa associated with microbialites were disproportionately represented among ostracod survivors.[108] Ostracod recovery began in the Spathian.[225] Despite high taxonomic turnover, the ecological life modes of Early Triassic ostracods remained rather similar to those of pre-PTME ostracods.[226]

Bryozoans in the Early Triassic were not diverse, represented mainly by members of Trepostomatida. During the Middle Triassic, there was a rise in bryozoan diversity, which peaked in the Carnian.[227]

Crinoids ("sea lilies") suffered a selective extinction, resulting in a decrease in the variety of their forms.[228] Though cladistic analyses suggest the beginning of their recovery to have taken place in the Induan, the recovery of their diversity as measured by fossil evidence was far less brisk, showing up in the late Ladinian.[229] Their adaptive radiation after the extinction event resulted in forms possessing flexible arms becoming widespread; motility, predominantly a response to predation pressure, also became far more prevalent.[230] Though their taxonomic diversity remained relatively low, crinoids regained much of their ecological dominance by the Middle Triassic epoch.[216]

Stem-group echinoids survived the PTME.[231] The survival of miocidarid echinoids such as Eotiaris is likely attributable to their ability to thrive in a wide range of environmental conditions.[232]

Conodonts saw a rapid recovery during the Induan,[233] with anchignathodontids experiencing a diversity peak in the earliest Induan. Gondolellids diversified at the end of the Griesbachian; this diversity spike was most responsible for the overall conodont diversity peak in the Smithian.[234] Segminiplanate conodonts again experienced a brief period of domination in the early Spathian, probably related to a transient oxygenation of deep waters.[235] Neospathodid conodonts survived the crisis but underwent proteromorphosis.[236]

Microbial reefs predominated across shallow seas for a short time during the earliest Triassic,

Griesbachian synchronously with a significant sea level drop that occurred then.[238] Metazoan-built reefs reemerged during the Olenekian, mainly being composed of sponge biostrome and bivalve builups.[241] Keratose sponges were particularly noteworthy in their integral importance to Early Triassic microbial-metazoan reef communities.[242][243] "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.[241] Bryozoans, after sponges, were the most numerous organisms in Tethyan reefs during the Anisian.[244] Metazoan reefs became common again during the Anisian because the oceans cooled down then from their overheated state during the Early Triassic.[245] Microbially induced sedimentary structures (MISS) from the earliest Triassic have been found to be associated with abundant opportunistic bivalves and vertical burrows, and it is likely that post-extinction microbial mats played a vital, indispensable role in the survival and recovery of various bioturbating organisms.[246]

Ichnocoenoses show that marine ecosystems recovered to pre-extinction levels of ecological complexity by the late Olenekian.[247] Anisian ichnocoenoses show slightly lower diversity than Spathian ichnocoenoses, although this was likely a taphonomic consequence of increased and deeper bioturbation erasing evidence of shallower bioturbation.[248]

Ichnological evidence suggests that recovery and recolonisation of marine environments may have taken place by way of outward dispersal from refugia that suffered relatively mild perturbations and whose local biotas were less strongly affected by the mass extinction compared to the rest of the world's oceans.[249][250] Although complex bioturbation patterns were rare in the Early Triassic, likely reflecting the inhospitability of many shallow water environments in the extinction's wake, complex ecosystem engineering managed to persist locally in some places, and may have spread from there after harsh conditions across the global ocean were ameliorated over time.[251] Wave-dominated shoreface settings (WDSS) are believed to have served as refugium environments because they appear to have been unusually diverse in the mass extinction's aftermath.[252]

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.

Dienerian boundary.[254] 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.[143] The worldwide recovery of gymnosperm forests took approximately 4–5 million years.[255][42] 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,[256] and floral recovery likely did not follow a congruent, globally universal trend but instead varied by region according to local environmental conditions.[148]

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.[61]

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.[256]

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.[257]

During the survival phase in the terrestrial extinction's immediate aftermath, from the latest Changhsingian to the Griesbachian, South China was dominated by opportunistic lycophytes.[258] Low-lying herbaceous vegetation dominated by the isoetalean Tomiostrobus was ubiquitous following the collapse of the gigantopterid-dominated forests of before. In contrast to the highly biodiverse gigantopterid rainforests, the post-extinction landscape of South China was near-barren and had vastly lower diversity.[148] Plant survivors of the PTME in South China experienced extremely high rates of mutagenesis induced by heavy metal poisoning.[259] From the late Griesbachian to the Smithian, conifers and ferns began to rediversify. After the Smithian, the opportunistic lycophyte flora declined, as the newly radiating conifer and fern species permanently replaced them as the dominant components of South China's flora.[258]

In Tibet, the early Dienerian Endosporites papillatusPinuspollenites thoracatus assemblages closely resemble late Changhsingian Tibetan floras, suggesting that the widespread, dominant latest Permian flora resurged easily after the PTME. However, in the late Dienerian, a major shift towards assemblages dominated by cavate trilete spores took place, heralding widespread deforestation and a rapid change to hotter, more humid conditions. Quillworts and spike mosses dominated Tibetan flora for about a million years after this shift.[260]

In Pakistan, then the northern margin of Gondwana, the flora was rich in lycopods associated with conifers and pteridosperms. Floral turnovers continued to occur due to repeated perturbations arising from recurrent volcanic activity until terrestrial ecosystems stabilised around 2.1 Myr after the PTME.[261]

In southwestern Gondwana, the post-extinction flora was dominated by bennettitaleans and cycads, with members 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.[142]

Coal gap

No

Abiotic factors (factors not caused by organisms), such as decreased rainfall or increased input of clastic sediments, may also be to blame.[42]

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.[42] 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.[262] Approximately 95% of these peat-producing plants went locally extinct at the P–Tr boundary;[263] coal deposits in Australia and Antarctica disappear significantly before the P–Tr boundary.[42]

Terrestrial vertebrates

Land vertebrates took an unusually long time to recover from the P–Tr extinction; palaeontologist

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.[269] Overall, terrestrial faunas after the extinction event tended to be more variable and heterogeneous across space than those of the Late Permian, which exhibited less provincialism, being much more geographically homogeneous.[270]

Synapsids

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

therapsids also survived, a group that included the ancestors of mammals.[277] As with dicynodonts, selective pressures favoured endothermic epicynodonts.[278] Therocephalians likewise survived; burrowing may have been a key adaptation that helped them make it through the PTME.[279] 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.[280] Early Triassic therocephalians were mostly survivors of the PTME rather than newly evolved taxa that originated during the evolutionary radiation in its aftermath.[281] Both therocephalians and cynodonts, known collectively as eutheriodonts, decreased in body size from the Late Permian to the Early Triassic.[277] This decrease in body size has been interpreted as an example of the Lilliput effect.[282]

Sauropsids

birds
(only extant dinosaurs) and mammals (only extant synapsids) would diversify and share the world.

Temnospondyls

Temnospondyl amphibians made a quick recovery; the appearance in the fossil record of so many temnospondyl clades suggests they may have been ideally suited as pioneer species that recolonised decimated ecosystems.[287] During the Induan, tupilakosaurids in particular thrived as disaster taxa,[288] including Tupilakosaurus itself,[289] though they gave way to other temnospondyls as ecosystems recovered.[288] Temnospondyls were reduced in size during the Induan, but their body size rebounded to pre-PTME levels during the Olenekian.[290] Mastodonsaurus and trematosaurians were the main aquatic and semiaquatic predators during most of the Triassic, some preying on tetrapods and others on fish.[291]

Terrestrial invertebrates

Most fossil insect groups found after the Permian–Triassic boundary differ significantly from those before: Of Paleozoic insect groups, only the

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.[134]

Microbially induced sedimentary structures (MISS) dominated North Chinese terrestrial fossil assemblages in the Early Triassic.[292][293] In Arctic Canada as well, MISS became a common occurrence following the Permian-Triassic extinction.[294] The prevalence of MISS in many Early Triassic rocks shows that microbial mats were an important feature of post-extinction ecosystems that were denuded of bioturbators that would have otherwise prevented their widespread occurrence. The disappearance of MISS later in the Early Triassic likely indicated a greater recovery of terrestrial ecosystems and specifically a return of prevalent bioturbation.[293]

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 each 200 million years or so by the ongoing processes 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).

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.[93]

Volcanism

Siberian Traps

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 (roughly the size of Saudi Arabia).[295][296][297][298] Such a vast areal extent of the flood basalts may have contributed to their exceptionally catastrophic impact.[299] The date of the Siberian Traps eruptions and the extinction event are in good agreement.[47][300]

The timeline of the extinction event strongly indicates it was caused by events in the large igneous province of the Siberian Traps.[301][19][302] A study of the Norilsk and Maymecha-Kotuy regions of the northern Siberian platform indicates that volcanic activity occurred during a small number of high intensity pulses that exuded enormous volumes of magma, as opposed to flows emplaced at regular intervals.[303]

The rate of carbon dioxide release from the Siberian Traps represented one of the most rapid rises of carbon dioxide levels in the geologic record,[304] with the rate of carbon dioxide emissions being estimated by one study to be five times faster than the rate during the already catastrophic Capitanian mass extinction event,[305] which occurred as a result of the activity of the Emeishan Traps in southwestern China at the end of the Middle Permian.[306][307][308] 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 according to one estimate.[21]

Another study estimated pre-PTME carbon dioxide levels at 400 ppm that then rose to around 2,500 ppm during the extinction event, with approximately 3,900 to 12,000 gigatonnes of carbon being added to the ocean-atmosphere system.[22] As carbon dioxide levels shot up, extreme temperature rise would have followed,[309] though some evidence suggests a lag of 12,000 to 128,000 years between the rise in volcanic carbon dioxide emissions and global warming.[310] During the latest Permian, before the PTME, global average surface temperatures were about 18.2 °C.[311] Global temperatures shot up to as much as 35 °C, and this hyperthermal condition may have lasted as long as 500,000 years.[22] Air temperatures at Gondwana's high southern latitudes experienced a warming of ~10–14 °C.[23] According to oxygen isotope shifts from conodont apatite in South China, low latitude surface water temperatures skyrocketed by about 8 °C.[24] In Iran, tropical SSTs were between 27 and 33 °C during the Changhsingian but jumped to over 35 °C during the PTME.[312]

So much carbon dioxide was released that inorganic carbon sinks were overwhelmed and depleted, enabling the extremely high carbon dioxide concentrations to persist in the atmosphere for much longer than would have otherwise been possible.[313] The position and alignment of Pangaea at the time made the inorganic carbon cycle very inefficient at returning volcanically emitted carbon back to the lithosphere and thus contributed to the exceptional lethality of carbon dioxide emissions during the PTME.[314] 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.[315][9] Further evidence based on paired coronene-mercury spikes for a volcanic combustion cause of the mass extinction has also been found.[316][30] The synchronicity of geographically disparate mercury anomalies with the environmental enrichment in isotopically light carbon confirms a common volcanogenic cause for these mercury spikes.[317] Te/Th values increase twentyfold over the PTME, further indicating it was concomitant with extreme volcanism.[318] A major volcanogenic influx of isotopically light zinc from the Siberian Traps has also been recorded, further confirming that volcanism was contemporary with the PTME.[319]

The Siberian Traps had unusual features that made them even more dangerous. The Siberian lithosphere is significantly enriched in

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 that interrupted the broader trend of rapid global warming,[322] leading to glacio-eustatic sea level fall.[320][323]

The eruptions may also have caused acid rain as the aerosols washed out of the atmosphere.

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.[325] 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.[309]

The Siberian Traps are underlain by thick sequences of Early-Mid

fly ash slurries developed. ... Mafic megascale eruptions are long-lived events that would allow significant build-up of global ash clouds."[331][332] 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."[333] However, some researchers propose that these supposed fly ashes were actually the result of wildfires instead, and were not related to massive coal combustion by intrusive volcanism.[334] 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.[335]

The sill-dominated mode of emplacement of the Siberian Traps made their warming effects more prolonged; whereas extrusive volcanism generates an abundance of subaerial basalts that efficiently sequester carbon dioxide via the silicate weathering process, underground sills cannot sequester atmospheric carbon dioxide and mitigate global warming.[336]

Mercury anomalies corresponding to the time of Siberian Traps activity have been found in many geographically disparate sites,

nickel aerosols were also released by Siberian Traps volcanic activity,[343][344] further contributing to metal poisoning.[345] Cobalt and arsenic emissions from the Siberian Traps caused further still environmental stress.[338]

The devastation wrought by the Siberian Traps did not end following the Permian-Triassic boundary. Stable carbon isotope fluctuations suggest that massive Siberian Traps activity recurred many times over the course of the Early Triassic;[346] this episodic return of severe volcanism caused further extinction events during the epoch.[347] Additionally, enhanced reverse weathering and depletion of siliceous carbon sinks enabled extreme warmth to persist for much longer than expected if the excess carbon dioxide was sequestered by silicate rock.[202] The decline in biological silicate deposition resulting from the mass extinction of siliceous organisms acted as a positive feedback loop wherein mass death of marine life exacerbated and prolonged extreme hothouse conditions.[348]

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.[142] Being about 1,300,000 cubic kilometres in volume[349] 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.[142] Specifically, this flood basalt has been implicated in the regional demise of the Gondwanan Glossopteris flora.[350]

Indochina-South China subduction-zone volcanic arc

Mercury anomalies preceding the end-Permian extinction have been discovered in what was then the boundary between the South China Craton and the Indochinese plate, which was home to a subduction zone and a corresponding volcanic arc. Hafnium isotopes from syndepositional magmatic zircons found in ash beds created by this pulse of volcanic activity confirm its origin in subduction-zone volcanism rather than large igneous province activity.[351] The enrichment of copper samples from these deposits in isotopically light copper provide additional confirmation of the felsic nature of this volcanism and that its origin was not a large igneous province.[352] This volcanism has been speculated to have caused local episodes of biotic stress among radiolarians, sponges, and brachiopods that took place over the 60,000 years preceding the end-Permian marine extinction, as well as an ammonoid crisis manifested in their decreased morphological complexity and size and their increased rate of turnover that began in the lower C. yini biozone, around 200,000 years prior to the end-Permian extinction.[351]

Methane clathrate gasification