fbpx
Wikipedia

Permian–Triassic extinction event

Approximately 251.9 million years ago, the Permian–Triassic (P–T, P–Tr) extinction event (PTME; also known as the Late Permian extinction event,[3] the Latest Permian extinction event,[4] the End-Permian extinction event,[5][6] and colloquially as the Great Dying)[7][8] forms the boundary between the Permian and Triassic geologic periods, and with them the Paleozoic and Mesozoic eras.[9] It is the Earth's most severe known extinction event,[10][11] with the extinction of 57% of biological families, 83% of genera, 81% of marine species[12][13][14] and 70% of terrestrial vertebrate species.[15] It is also the greatest known mass extinction of 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]

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]

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 (oxygen-starved, sulfurous oceans),[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] Several other contributing factors have been proposed, including the emission of carbon dioxide from the burning of oil and coal deposits ignited by the eruptions;[27][28] emissions of methane from the gasification of methane clathrates;[29] emissions of methane by novel methanogenic microorganisms nourished by minerals dispersed in the eruptions;[30][31][32] and an extraterrestrial impact which created the Araguainha crater and caused seismic release of methane[33][34][35] and the destruction of the ozone layer with increased exposure to solar radiation.[36][37][38]

Dating edit

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.[47] A large, abrupt global decrease in δ13C, the ratio of the stable isotope 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,[42] and an increase in CO
2
levels to 2,500 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 fungi, caused by the sharp increase in the amount of dead plants and animals fed upon by the fungi.[56] This "fungal spike" has been used by some paleontologists to identify a lithological sequence as being on or very close to the Permian–Triassic boundary in rocks that are unsuitable for radiometric dating or have a lack of suitable index fossils.[57] However, 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.[56] 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;[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 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][84] Many of the extinctions once dated to the Permian-Triassic boundary have more 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. This older theory, still supported in some recent papers,[15][85] proposes that there were two major extinction pulses 9.4 million years apart, separated by a period of extinctions that were less extensive, but still well above the background level, and that the final extinction killed off only about 80% of marine species alive at that time, whereas 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.[86][15][87] For example, all dinocephalian genera died out at the end of the Guadalupian,[85] as did the Verbeekinidae, a family of large-size 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 edit

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 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.[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. These organisms were susceptible to the effects of the ocean acidification that resulted from increased atmospheric CO2.[93] Organisms that relied on haemocyanin or haemoglobin for transporting oxygen were more resistant to extinction than those utilising haemerythrin or oxygen diffusion.[94] 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.[95] There was little latitudinal difference in the survival rates of taxa.[96]

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).[97][98] The extinction rate of marine organisms was catastrophic.[18][99][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.[100]

Surviving marine invertebrate groups included articulate brachiopods (those with a hinge),[101] which had undergone a slow decline in numbers since the P–Tr extinction; the Ceratitida order of ammonites;[102] and crinoids ("sea lilies"),[102] 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.[103][104] In the case of the brachiopods, at least, surviving taxa were generally small, rare members of a formerly diverse community.[105]

Conodonts were severely affected both in terms of taxonomic and morphological diversity, although not as severely as during the Capitanian mass extinction.[106]

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.[107] 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.[108]

Ostracods experienced prolonged diversity perturbations during the Changhsingian before the PTME proper, when immense proportions of them abruptly vanished.[109] At least 74% of ostracods died out during the PTME itself.[110]

Bryozoans had been on a long-term decline throughout the Late Permian epoch before they suffered even more catastrophic losses during the PTME.[111]

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.[112]

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.[113] 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.[114]

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 Early Cretaceous cladodontomorphs in deep, outer shelf environments.[115] Ichthyosaurs, which are believed to have evolved immediately before the PTME, were also PTME survivors.[116]

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,[117][118] notably occurring in foraminifera,[119][120][121] brachiopods,[122][123][124] bivalves,[125][126][127] and ostracods.[128][129] Though gastropods that survived the cataclysm were smaller in size than those that did not,[130] it remains debated whether the Lilliput effect truly took hold among gastropods.[131][132][133] Some gastropod taxa, termed "Gulliver gastropods", ballooned in size during and immediately following the mass extinction,[134] exemplifying the Lilliput effect's opposite, which has been dubbed the Brobdingnag effect.[135]

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;[16] according to some sources, it may well be the only mass extinction to significantly affect insect diversity.[136][137] 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 edit

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.[138] 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.[139][42] However, a massive rearrangement of ecosystems does occur, with plant abundances and distributions changing profoundly and all the forests virtually disappearing.[140][42] The dominant floral groups changed, with many groups of land plants entering abrupt decline, such as Cordaites (gymnosperms) and Glossopteris (seed ferns).[141][142] The severity of plant extinction has been disputed.[143][139]

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.[144]

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.[61]

The Cordaites flora, which dominated the Angaran floristic realm corresponding to Siberia, collapsed over the course of the extinction.[145] 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.[146]

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.[147]

In South China, the subtropical Cathaysian gigantopterid dominated rainforests abruptly collapsed.[148][149][150] 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.[151] Wildfires too likely played a role in the fall of Gigantopteris.[152]

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.[153]

Terrestrial vertebrates edit

The terrestrial vertebrate extinction occurred rapidly, taking 50,000 years or less.[154] Aridification induced by global warming was the chief culprit behind terrestrial vertebrate extinctions.[155][156] 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).[157][10] Gorgonopsians are traditionally thought to have gone extinct during the PTME, but some tentative evidence suggests they may have survived into the Triassic.[158]

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.[159]

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.[160] 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.[161] Another study of latest Permian vertebrates in the Karoo Basin found that 54% of them went extinct due to the PTME.[162]

Biotic recovery edit

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;[163] the Permian-Triassic mass extinction marked a key turning point in this ecological shift that began after the Capitanian mass extinction[164] and culminated in the Late Jurassic.[165] Typical taxa of shelly benthic faunas were now bivalves, snails, sea urchins and Malacostraca, whereas bony fishes[166] and marine reptiles[167] 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).[168][169] However, recovery was also differential between taxa. Some survivors became extinct some million years after the extinction event without having rediversified (dead clade walking,[170] e.g. the snail family Bellerophontidae),[171] whereas others rose to dominance over geologic times (e.g., bivalves).[172][173]

Marine ecosystems edit

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

A cosmopolitanism event began immediately after the end-Permian extinction event.[174] Marine post-extinction faunas were mostly species-poor and were dominated by few disaster taxa such as the bivalves Claraia, Unionites, Eumorphotis, and Promyalina,[175] the conodonts Clarkina and Hindeodus,[176] the inarticulate brachiopod Lingularia,[175] and the foraminifera Earlandia and Rectocornuspira kalhori,[177] the latter of which is sometimes classified under the genus Ammodiscus.[178] Their guild diversity was also low.[179]

The speed of recovery from the extinction is disputed. Some scientists estimate that it took 10 million years (until the Middle Triassic)[180] due to the severity of the extinction. However, studies in Bear Lake County, near Paris, Idaho,[181] and nearby sites in Idaho and Nevada[182] showed a relatively quick rebound in a localized Early Triassic marine ecosystem (Paris biota), taking around 1.3 million years to recover,[181] while an unusually diverse and complex ichnobiota is known from Italy less than a million years after the end-Permian extinction.[183] Additionally, the complex Guiyang biota found near Guiyang, China also indicates life thrived in some places just a million years after the mass extinction,[184][185] as does a fossil assemblage known as the Shanggan fauna found in Shanggan, China,[186] the Wangmo biota from the Luolou Formation of Guizhou,[187] and a gastropod fauna from the Al Jil Formation of Oman.[188] Regional differences in the pace of biotic recovery existed,[189] 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.[190][191] 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;[192] 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.[193]

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.[194] Epifaunal benthos took longer to recover than infaunal benthos.[195] 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,[196] and conodonts, which diversified considerably over the first two million years of the Early Triassic.[197]

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.[198] That recovery was slow in the Early Triassic can be explained by low levels of biological competition due to the paucity of taxonomic diversity,[199] and that biotic recovery explosively accelerated in the Anisian can be explained by niche crowding, a phenomenon that would have drastically increased competition, becoming prevalent by the Anisian.[200] Biodiversity rise thus behaved as a positive feedback loop enhancing itself as it took off in the Spathian and Anisian.[201] 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.[199] Other explanations state that life was delayed in its recovery because grim conditions returned periodically over the course of the Early Triassic,[202] causing further extinction events, such as the Smithian-Spathian boundary extinction.[11][203][204] Continual episodes of extremely hot climatic conditions during the Early Triassic have been held responsible for the delayed recovery of oceanic life,[205][206] in particular skeletonised taxa that are most vulnerable to high carbon dioxide concentrations.[207] The relative delay in the recovery of benthic organisms has been attributed to widespread anoxia,[208] but high abundances of benthic species contradict this explanation.[209] 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.[191] Recurrent Early Triassic environmental stresses also acted as a ceiling limiting the maximum ecological complexity of marine ecosystems until the Spathian.[210] Recovery biotas appear to have been ecologically uneven and unstable into the Anisian, making them vulnerable to environmental stresses.[211]

Whereas most marine communities were fully recovered by the Middle Triassic,[212][213] global marine diversity reached pre-extinction values no earlier than the Middle Jurassic, approximately 75 million years after the extinction event.[214]

 
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.[215] 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,[215] and the increase in predation pressure and durophagy led to the Mesozoic Marine Revolution.[216]

Bivalves rapidly recolonised many marine environments in the wake of the catastrophe.[217] 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.[218] Bivalves were once thought to have outcompeted brachiopods, but this outdated hypothesis about the brachiopod-bivalve transition has been disproven by Bayesian analysis.[219] 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.[165] 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.[220] 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.[221] Infaunal habits in bivalves became more common after the PTME.[222]

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.[223] The surviving brachiopod fauna was very low in diversity and exhibited no provincialism whatsoever.[224] 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.[225] 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.[226] 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.[227]

Ostracods were extremely rare during the basalmost Early Triassic.[228] Taxa associated with microbialites were disproportionately represented among ostracod survivors.[110] Ostracod recovery began in the Spathian.[229] Despite high taxonomic turnover, the ecological life modes of Early Triassic ostracods remained rather similar to those of pre-PTME ostracods.[230]

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.[231]

Crinoids ("sea lilies") suffered a selective extinction, resulting in a decrease in the variety of their forms.[232] 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.[233] 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.[234] Though their taxonomic diversity remained relatively low, crinoids regained much of their ecological dominance by the Middle Triassic epoch.[220]

Stem-group echinoids survived the PTME.[235] The survival of miocidarid echinoids such as Eotiaris is likely attributable to their ability to thrive in a wide range of environmental conditions.[236]

Conodonts saw a rapid recovery during the Induan,[237] 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.[238] Segminiplanate conodonts again experienced a brief period of domination in the early Spathian, probably related to a transient oxygenation of deep waters.[239] Neospathodid conodonts survived the crisis but underwent proteromorphosis.[240]

In the PTME's aftermath, disaster taxa of benthic foraminifera filled many of their vacant niches. The recovery of benthic foraminifera was very slow and frequently interrupted until the Spathian.[241]

Microbial reefs predominated across shallow seas for a short time during the earliest Triassic,[242] occurring both in anoxic and oxic waters.[243] Polybessurus-like microfossils often dominated these earliest Triassic microbialites.[244] Microbial-metazoan reefs appeared very early in the Early Triassic;[245] and they dominated many surviving communities across the recovery from the mass extinction.[246] Microbialite deposits appear to have declined in the early Griesbachian synchronously with a significant sea level drop that occurred then.[243] Metazoan-built reefs reemerged during the Olenekian, mainly being composed of sponge biostrome and bivalve builups.[246] Keratose sponges were particularly noteworthy in their integral importance to Early Triassic microbial-metazoan reef communities.[247][248] "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.[246] Bryozoans, after sponges, were the most numerous organisms in Tethyan reefs during the Anisian.[249] Metazoan reefs became common again during the Anisian because the oceans cooled down then from their overheated state during the Early Triassic.[250] Biodiversity amongst metazoan reefs did not recover until well into the Anisian, millions of years after non-reef ecosystems recovered their diversity.[251] 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.[252]

Ichnocoenoses show that marine ecosystems recovered to pre-extinction levels of ecological complexity by the late Olenekian.[253] 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.[254]

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.[255][256] 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.[257] 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.[258]

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.[259] Furthermore, severe extinction pulses continued to occur after the Permian-Triassic boundary, causing additional floral turnovers. Gymnosperms recovered within a few thousand years after the Permian-Triassic boundary, but around 500,000 years after it, the Dominant gymnosperm genera were replaced by lycophytes – extant lycophytes are recolonizers of disturbed areas – during an extinction pulse at the Griesbachian-Dienerian boundary.[260] 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.[145] The worldwide recovery of gymnosperm forests took approximately 4–5 million years.[261][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,[262] and floral recovery likely did not follow a congruent, globally universal trend but instead varied by region according to local environmental conditions.[150]

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.[262]

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.[263]

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.[264] 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.[150] Plant survivors of the PTME in South China experienced extremely high rates of mutagenesis induced by heavy metal poisoning.[265] 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.[264]

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.[266]

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.[267]

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.[144]

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.[43] 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.[268] Approximately 95% of these peat-producing plants went locally extinct at the P–Tr boundary;[269] coal deposits in Australia and Antarctica disappear significantly before the P–Tr boundary.[42]

Terrestrial vertebrates edit

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.[13] A tetrapod gap may have existed from the Induan until the early Spathian between ~30 °N and ~ 40 °S due to extreme heat making these low latitudes uninhabitable for these animals. During the hottest phases of this interval, the gap would have spanned an even greater latitudinal range.[270] East-central Pangaea, with its relatively wet climate, served as a dispersal corridor for PTME survivors during their Early Triassic recolonisation of the supercontinent.[271] 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.[272][273] In Russia, even after 15 Myr of recovery, during which ecosystems were rebuilt and remodelled, many terrestrial vertebrate guilds were absent, including small insectivores, small piscivores, large herbivores, and apex predators.[274] 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.[275] 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.[276]

Synapsids 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, although some recent evidence has called into question its status as a post-PTME disaster taxon.[277] The evolutionary success of Lystrosaurus in the aftermath of the PTME is believed to be attributable to the dicynodont taxon's grouping behaviour and tolerance for extreme and highly variable climatic conditions.[278] Other likely factors behind the success of Lystrosaurus included extremely fast growth rate exhibited by the dicynodont genus,[279] along with its early onset of sexual maturity.[280] Antarctica may have served as a refuge for dicynodonts during the PTME from which surviving dicynodonts spread out of in its aftermath.[281] Ichnological evidence from the earliest Triassic of the Karoo Basin shows dicynodonts were abundant in the immediate aftermath of the biotic crisis.[282] Smaller carnivorous cynodont therapsids also survived, a group that included the ancestors of mammals.[283] As with dicynodonts, selective pressures favoured endothermic epicynodonts.[284] Therocephalians likewise survived; burrowing may have been a key adaptation that helped them make it through the PTME.[285] 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.[286] Early Triassic therocephalians were mostly survivors of the PTME rather than newly evolved taxa that originated during the evolutionary radiation in its aftermath.[287] Both therocephalians and cynodonts, known collectively as eutheriodonts, decreased in body size from the Late Permian to the Early Triassic.[283] This decrease in body size has been interpreted as an example of the Lilliput effect.[288]

Sauropsids edit

Archosaurs (which included the ancestors of dinosaurs and crocodilians) were initially rarer than therapsids, but they began to displace therapsids in the mid-Triassic. Olenekian tooth fossil assemblages from the Karoo Basin indicate that archosauromorphs were already highly diverse by this point in time, though not very ecologically specialised.[289] 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.[290] 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,[291] while losing part of the differential color-sensitive retinal receptors reptilians and birds preserved. Archosaurs also experienced an increase in metabolic rates over time during the Early Triassic.[292] 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.

Temnospondyls edit

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.[293] During the Induan, tupilakosaurids in particular thrived as disaster taxa,[294] including Tupilakosaurus itself,[295] though they gave way to other temnospondyls as ecosystems recovered.[294] Temnospondyls were reduced in size during the Induan, but their body size rebounded to pre-PTME levels during the Olenekian.[296] Mastodonsaurus and trematosaurians were the main aquatic and semiaquatic predators during most of the Triassic, some preying on tetrapods and others on fish.[297]

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, 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.[136]

Microbially induced sedimentary structures (MISS) dominated North Chinese terrestrial fossil assemblages in the Early Triassic.[298][299] In Arctic Canada as well, MISS became a common occurrence following the Permian-Triassic extinction.[300] 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.[299]

Hypotheses about cause edit

Explaining an event from 250 million years ago is inherently difficult, with much of the evidence on land eroded or deeply buried, while the spreading seafloor is completely recycled over 200 million years, 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, but with much less current consensus).

  • 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.[93]

Volcanism edit

Siberian Traps edit

The flood basalt eruptions that produced the large igneous province of the Siberian Traps were among the largest known volcanic events, extruding lava over 2,000,000 square kilometres (770,000 sq mi), roughly the size of Saudi Arabia, producing a catastrophic impact.[301][302][303][304][305] The date of the Siberian Traps eruptions matches well with the extinction event.[47][306][307][19][308] A study of the Norilsk and Maymecha-Kotuy regions of the northern Siberian platform indicates that volcanic activity occurred during a few enormous pulses of magma, as opposed to more regular flows.[309]

The Siberian Traps caused one of the most rapid rises of atmospheric carbon dioxide levels in the geologic record,[310] with the rate of carbon dioxide emissions estimated as five times faster than during the preceeding catastrophic Capitanian mass extinction[311] during the eruption of the Emeishan Traps.[312][313][314] Overwhelming inorganic carbon sinks, carbon dioxide levels might have jumped from between 500 and 4,000 ppm prior to the extinction to around 8,000 ppm after, according to one estimate.[21] Another study estimated pre-extinction carbon dioxide levels at 400 ppm, which then rose to 2,500 ppm, with 3,900 to 12,000 gigatonnes of carbon added to the ocean-atmosphere system.[22] Extreme temperature rise would have followed,[315] though some evidence suggests a lag of 12,000 to 128,000 years between the rise in volcanic carbon dioxide emissions and global warming.[316] During the latest Permian before the extinction, global average surface temperatures were about 18.2 °C,[317] which shot up to as much as 35 °C, this hyperthermal condition lasting 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 surged about 8 °C.[24] In present-day Iran, tropical sea surface temperatures were between 27 and 33 °C during the Changhsingian but jumped to over 35 °C during the PTME.[318]

These extremely high atmospheric carbon dioxide concentrations persisted over a long period.[319] The position and alignment of Pangaea at the time made the inorganic carbon cycle very inefficient at burying carbon.[320] 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.[321][9] Evidence also points to volcanic combustion of undergroud fossil fuel deposits, based on paired coronene-mercury spikes[322][30] coinciding with geographically widespread mercury anomalies and the rise in isotopically light carbon.[323] Te/Th values increase twentyfold over the PTME, further indicating it was concomitant with extreme volcanism.[324] 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.[325]

The Siberian Traps eruptions had unusual features that made them even more dangerous. The Siberian lithosphere is rich in halogens extremely destructive to the ozone layer, and evidence from subcontinental lithospheric xenoliths indicates that as much as 70% of their halogen content was released into the atmosphere.[326] Around 18 teratonnes of hydrochloric acid were emitted,[327] along with sulphur-rich volatiles that caused dust clouds and acid aerosols, which would have blocked out sunlight and disrupted photosynthesis 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 punctuating the broader trend of rapid global warming,[328] with glacio-eustatic sea level fall.[326][329]

The eruptions may also have caused acid rain as the aerosols washed out of the atmosphere.[330] That may have killed land plants and mollusks and planktonic organisms with 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 ash thrown high into the atmosphere, increasing the short-term cooling effect.[331] When this had washed out of the atmosphere, the excess carbon dioxide would have remained and global warming would have proceeded unchecked.[315]

Burning of hydrocarbon deposits may have exacerbated the extinction. 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 may emit large amounts of greenhouse and toxic gases.[332] The unique setting of the Siberian Traps over these deposits is likely the reason for the severity of the extinction.[333][334][335] The basalt lava erupted or intruded into carbonate rocks and sediments in the process of forming large coal beds, which would have emitted large amounts of carbon dioxide, leading to stronger global warming after the dust and aerosols settled.[315] The change of the eruptions from flood basalt to sill dominated emplacement, liberating even more trapped hydrocarbon deposits, coincides with the main onset of the extinction[27] and is linked to a major negative δ13C excursion.[336]

Venting of coal-derived methane was accompanied by explosive combustion of coal and discharge of coal-fly ash.[28] 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. They found ash deposits in deep rock layers near what is now the Buchanan Lake Formation: "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."[337][338] 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."[339] However, some researchers propose that these supposed fly ashes were actually the result of wildfires not related to massive coal combustion by intrusive volcanism.[340] 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.[341]

The sill-dominated emplacement of the Siberian Traps prolonged their warming effects; 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.[342] 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.[206] 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.[343]

Mercury anomalies corresponding to the time of Siberian Traps activity have been found in many geographically disparate sites,[344] indicating that these volcanic eruptions released significant quantities of toxic mercury into the atmosphere and ocean, causing even larger terrestrial and marine die-offs.[345][346][347] A series of surges raised terrestrial and marine environmental mercury concentrations by orders of magnitude above normal background levels and caused mercury poisoning over periods of a thousand years each.[348] Mutagenesis in surviving plants after the PTME coeval with mercury and copper loading confirms volcanically induced heavy metal toxicity.[265] Increased bioproductivity may have sequestered mercury and party mitigated poisoning.[349] Immense volumes of nickel aerosols and cobalt and arsenic emisions, were also released,[350][351][345] further contributing to metal poisoning.[352]

The devastation wrought by the Siberian Traps did not end following the Permian-Triassic boundary. Carbon isotope fluctuations suggest that massive Siberian Traps activity recurred many times during the Early Triassic,[353] causing further extinction events during the epoch.[354]

Choiyoi Silicic Large Igneous Province edit

A second flood basalt event that produced the Choiyoi Silicic Large Igneous Province in southwestern Gondwana between around 286 Ma and 247 Ma has also been suggested as a significant additional extinction mechanism.[144] At 1,300,000 cubic kilometres in volume[355] and 1,680,000 square kilometres in area, this event was 40% the size of the Siberian Traps.[144] Specifically, this flood basalt has been implicated in the regional demise of the Gondwanan Glossopteris flora.[356]

Indochina-South China subduction-zone volcanic arc edit

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, a subduction zone with a volcanic arc. Hafnium isotopes from syndepositional magmatic zircons found in ash beds created by this volcanic pulse confirm its origin in subduction-zone volcanism rather than large igneous province activity.[357] The enrichment of copper samples from these deposits in isotopically light copper provide additional confirmation.[358] This volcanism has been speculated to have caused local biotic stress among radiolarians, sponges, and brachiopods over the 60,000 years preceding the end-Permian marine extinction, as well as an ammonoid crisis with decreased morphological complexity and size and increased rate of turnover beginning in the lower C. yini biozone, around 200,000 years before the extinction.[357]

Methane clathrate gasification edit

Methane clathrates, also known as methane hydrates, consist of molecules of methane trapped in the crystal lattice of ice. This methane, produced by methanogen microbes, has a 13C 12C isotope ratio about 6% below normal (δ13C −6.0%). At the right combination of pressure and temperature, clathrates form near the surface of permafrost and in large quantities on continental shelves and nearby seabed at water depths of at least 300 m (980 ft), buried in sediments up to 2,000 m (6,600 ft) below the sea floor.[359]

Massive release of methane from these clathrates may have contributed to the PTME, as scientists have found worldwide evidence of a swift decrease of about 1% in the 13C 12C ratio in carbonate rocks from the end-Permian.[75][360] This is the first, largest, and fastest of a series of excursions (decreases and increases) in the ratio, until it abruptly stabilised in the middle Triassic, followed soon afterwards by the recovery of calcifying shelled sealife.[103] The seabed probably contained methane hydrate deposits, and the lava caused the deposits to dissociate, releasing vast quantities of methane.[361] 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);[362] 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.[363] 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 (sulfidic) events at the time, with the exact mechanism compared to the 1986 Lake Nyos disaster.[364]

The clathrate hypothesis seemed the only proposed mechanism sufficient to cause a global 1% reduction in the 13C 12C ratio .[365][44] While a variety of factors may have contributed to the ratio drop, a 2002 review found most of them insufficient to account for the observed amount:[29]

  • Gases from volcanic eruptions have a 13C 12C ratio about 0.5 to 0.8% below standard (δ13C −0.5 to −0.8%), but a 1995 assessment concluded that the observed 1.0% worldwide reduction would have required eruptions massively larger than any found.[366] (However, this analysis addressed only CO2 produced by the magma itself, not from interactions with carbon bearing sediments, as described below.)
  • 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.[366]
  • 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 gigatonnes (1 gigatonne = 1012 kg) 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.[44] Moreover, sea levels were rising rather than falling at the time of the extinction.[315]
  • 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;[103] 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.[367] That or another organic-based reason may have been responsible for both that and a late Proterozoic/Cambrian pattern of fluctuating 13C 12C ratios.[103]

However, the clathrate hypothesis has also been criticized. Carbon-cycle models which include consideration of roasting carbonate sediments by volcanism confirm that it would have had enough effect to produce the observed reduction.[29][368] Also, 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, 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.[103] The latest research suggests that greenhouse gas release during the extinction event was dominated by volcanic carbon dioxide,[369] 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.[22]

Adding to the evidence against methane clathrate release as the central driver of warming, the main rapid warming event is also associated with marine transgression rather than regression; the former would not normally have initiated methane release, which would have instead required a decrease in pressure, something that would be generated by a retreat of shallow seas.[370] The configuration of the world's landmasses into one supercontinent would also mean that the global gas hydrate reservoir was lower than today, further damaging the case for methane clathrate dissolution as a major cause of the carbon cycle disruption.[371]

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.[160] Higher concentrations of CO2 also result in decreased activity levels in many active marine animals, hindering their ability to obtain food.[372] 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.[160]

In addition, an increase in CO2 concentration is inevitably linked to ocean acidification,[373] consistent with the preferential extinction of heavily calcified taxa and other signals in the rock record that suggest a more acidic ocean.[26] The decrease in ocean pH is calculated to be up to 0.7 units.[25] 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.[3] 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.[374] One study, published in Scientific Reports, concluded that widespread ocean acidification, if it did occur, was not intense enough to impede calcification and only occurred during the beginning of the extinction event.[375] The relative success of many marine organisms that were very vulnerable to acidification has further been used to argue that acidification was not a major extinction contributor.[376] The persistence of highly elevated carbon dioxide concentrations in the atmosphere during the Early Triassic would have impeded the recovery of biocalcifying organisms after the PTME.[377]

Acidity generated by increased carbon dioxide concentrations in soil and sulphur dioxide dissolution in rainwater was also a kill mechanism on land.[378] 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.[379] Further evidence of an increase in soil acidity comes from elevated Ba/Sr ratios in earliest Triassic soils.[380] 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.[381] 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.[382] Hopane anomalies have also been interpreted as evidence of acidic soils and peats.[383] As with many other environmental stressors, acidity on land episodically persisted well into the Triassic, stunting the recovery of terrestrial ecosystems.[384]

Anoxia and euxinia edit

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.[385][386][387] Throughout most of the Tethys and Panthalassic Oceans, evidence for anoxia appears at the extinction event, including small pyrite framboids,[388] negative δ238U excursions,[389][390] negative δ15N excursions,[391] positive δ82/78Se isotope excursions,[392] relatively positive δ13C ratios in polycyclic aromatic hydrocarbons,[393] high Th/U ratios,[389] positive Ce/Ce* anomalies,[394] and fine laminations in sediments.[388] However, evidence for anoxia precedes the extinction at some other sites, including Spiti, India,[395] Shangsi, China,[396] Meishan, China,[397] Opal Creek, Alberta,[398] and Kap Stosch, Greenland.[399] Biogeochemical evidence also points to the presence of euxinia during the PTME.[400] 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.[401] Negative mercury isotope excursions further bolster evidence for extensive euxinia during the PTME.[402] 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.[403] Evidence of a lag between volcanic mercury inputs and biotic turnovers provides further support for anoxia and euxinia as the key killing mechanism, because extinctions would be expected to be synchronous with volcanic mercury discharge if volcanism and hypercapnia was the primary driver of extinction.[404]

The sequence of events leading to anoxic oceans may have been triggered by carbon dioxide emissions from the eruption of the Siberian Traps.[405] 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 coastal evaporation would have caused the formation of warm saline bottom water (WSBW) depleted in oxygen and nutrients, which spread across the world through the deep oceans. The influx of WSBW caused thermal expansion of water that raised sea levels, bringing anoxic waters onto shallow shelfs and enhancing the formation of WSBW in a positive feedback loop.[406] The flux of terrigeneous material into the oceans increased as a result of soil erosion, which would have facilitated increased eutrophication.[407] Increased chemical weathering of the continents due to warming and the acceleration of the water cycle would increase the riverine flux of nutrients to the ocean.[408] Additionally, the Siberian Traps directly fertilised the oceans with iron and phosphorus as well, triggering bioblooms and marine snowstorms. Increased phosphate levels would have supported greater primary productivity in the surface oceans.[409] 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.[410] Along the Panthalassan coast of South China, oxygen decline was also driven by large-scale upwelling of deep water enriched in various nutrients, causing this region of the ocean to be hit by especially severe anoxia.[411] Convective overturn helped facilitate the expansion of anoxia throughout the water column.[412] 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.[405] In some regions, anoxia briefly disappeared when transient cold snaps resulting from volcanic sulphur emissions occurred.[413]

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).[414] 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,[415][416][417] particularly that of benthic organisms.[418][208] Anoxia disappeared from shallow waters more rapidly than the deep ocean.[419] Reexpansions of oxygen-minimum zones did not cease until the late Spathian, periodically setting back and restarting the biotic recovery process.[420] The decline in continental weathering towards the end of the Spathian at last began ameliorating marine life from recurrent anoxia.[421] In some regions of Panthalassa, pelagic zone anoxia continued to recur as late as the Anisian,[422] probably due to increased productivity and a return of aeolian upwelling.[423] Some sections show a rather quick return to oxic water column conditions, however, so for how long anoxia persisted remains debated.[424] The volatility of the Early Triassic sulphur cycle suggests marine life continued to face returns of euxinia as well.[425]

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.[426]

Oxygen depletion via extensive microbial blooms also played a role in the biological collapse of not just marine ecosystems but freshwater ones as well. Persistent lack of oxygen after the extinction event itself helped delay biotic recovery for much of the Early Triassic epoch.[427]

Aridification edit

Increasing continental aridity, a trend well underway even before the PTME as a result of the coalescence of the supercontinent Pangaea, was drastically exacerbated by terminal Permian volcanism and global warming.[428] The combination of global warming and drying generated an increased incidence of wildfires.[429] Tropical coastal swamp floras such as those in South China may have been very detrimentally impacted by the increase in wildfires,[152] though it is ultimately unclear if an increase in wildfires played a role in driving taxa to extinction.[430]

Aridification trends varied widely in their tempo and regional impact. Analysis of the fossil river deposits of the floodplains of the Karoo Basin indicate a shift from meandering to braided river patterns, indicating a very abrupt drying of the climate.[431] 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.[432] A pattern of aridity-induced extinctions that progressively ascended up the food chain has been deduced from Karoo Basin biostratigraphy.[155] Evidence for aridification in the Karoo across the Permian-Triassic boundary is not, however, universal, as some palaeosol evidence indicates a wettening of the local climate during the transition between the two geologic periods.[433] 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.[434] 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.[435] 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. Drying of this basin 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.[146]

The persistence of hyperaridity varied regionally as well. 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.[273][272] Elsewhere, such as in the Karoo Basin, episodes of dry climate recurred regularly in the Early Triassic, with profound effects on terrestrial tetrapods.[428]

The types and diversity of ichnofossils in a locality has been used as an indicator measuring aridity. Nurra, an ichnofossil site on the island of Sardinia, shows evidence of major drought-related stress among crustaceans. Whereas the Permian subnetwork at Nurra displays extensive horizontal backfilled traces and high ichnodiversity, the Early Triassic subnetwork is characterised by an absence of backfilled traces, an ichnological sign of aridification.[436]

Ozone depletion edit

A collapse of the atmospheric ozone shield has been invoked as an explanation for the mass extinction,[437] particularly that of terrestrial plants.[37] Ozone production may have been reduced by 60-70%, increasing the flux of ultraviolet radiation by 400% at equatorial latitudes and 5,000% at polar latitudes.[438] 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.[439][38] Malformed plant spores from the time of the extinction event show an increase in ultraviolet B absorbing compounds, confirming that increased ultraviolet radiation played a role in the environmental catastrophe and excluding other possible causes of mutagenesis, such as heavy metal toxicity, in these mutated spores.[36] Extremely positive Δ33S anomalies provide evidence of photolysis of volcanic SO2, indicating increased ultraviolet radiation flux.[440]

Multiple mechanisms could have reduced the ozone shield and rendered it ineffective. Computer modelling shows high atmospheric methane levels are associated with ozone shield decline and may have contributed to its reduction during the PTME.[441] Volcanic emissions of sulphate aerosols into the stratosphere would have dealt significant destruction to the ozone layer.[439] As mentioned previously, the rocks in the region where the Siberian Traps were emplaced are extremely rich in halogens.[326] The intrusion of Siberian Traps volcanism into deposits rich in organohalogens, such as methyl bromide and methyl chloride, would have been another source of ozone destruction.[38] An uptick in wildfires, a natural source of methyl chloride, would have had further deleterious effects still on the atmospheric ozone shield.[442] Upwelling of euxinic water may have released massive hydrogen sulphide 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.[443] Indeed, biomarker evidence for anaerobic photosynthesis by Chlorobiaceae (green sulfur bacteria) from the Late-Permian into the Early Triassic indicates that hydrogen sulphide did upwell into shallow waters because these bacteria are restricted to the photic zone and use sulfide as an electron donor.[444]

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.[445] However, suggestions that an asteroid impact was the trigger of the Permian-Triassic extinction are now largely rejected.[446]

Reported evidence for an impact event from the P–Tr boundary level includes rare grains of shocked quartz in Australia and Antarctica;[447][448] fullerenes trapping extraterrestrial noble gases;[449] meteorite fragments in Antarctica;[450] and grains rich in iron, nickel, and silicon, which may have been created by an impact.[451] However, the accuracy of most of these claims has been challenged.[452][453][454][455] 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.[456] Iridium levels in many sites straddling the Permian-Triassic boundaries are not anomalous, providing evidence against an extraterrestrial impact as the PTME's cause.[457]

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 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.[458] 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,[315] such as the Siberian Traps eruptions at either an impact site[459] or the antipode of an impact site.[315][460] 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.[461] Additionally, many sites spanning the Permian-Triassic boundary display a complete lack of evidence of an impact event.[462]

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[448] and the hypothesized 480 km (300 mi) Wilkes Land crater of East Antarctica.[463][464] 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,[33] precipitated the mass extinction.[465] 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.[34] The large earthquakes generated by the impact would have additionally generated massive tsunamis across much of the globe.[465][35] 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.[34]

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),[466] 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.[467]

Methanogens edit

A hypothesis published in 2014 posits that a genus of anaerobic methanogenic archaea known as Methanosarcina was responsible for the event. 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 marine sediment. The result would have been a sharp buildup of methane and carbon dioxide in the 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.[31] Chemostratigraphic analysis of Permian-Triassic boundary sediments in Chaotian demonstrates a methanogenic burst could be responsible for some percentage of the carbon isotopic fluctuations.[32] On the other hand, in the canonical Meishan sections, the nickel concentration increases somewhat after the δ13C concentrations have begun to fall.[468]

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.[469]

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 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.[470]

Comparison to present global warming edit

The PTME has been compared to the current anthropogenic global warming situation and Holocene extinction due to sharing the common characteristic of rapid rates of carbon dioxide release. Though the current rate of greenhouse gas emissions is more than an order of magnitude greater than the rate measured over the course of the PTME, the discharge of greenhouse gases during the PTME is poorly constrained geo-chronologically and was most likely pulsed and constrained to a few key, short intervals, rather than continuously occurring at a constant rate for the whole extinction interval; the rate of carbon release within these intervals was likely to have been similar in timing to modern anthropogenic emissions.[310] As they did during the PTME, oceans in the present day are experiencing drops in pH and in oxygen levels, prompting further comparisons between modern anthropogenic ecological conditions and the PTME.[471] Another biocalcification event similar in its effects on modern marine ecosystems is predicted to occur if carbon dioxide levels continue to rise.[377] The similarities between the two extinction events have led to warnings from geologists about the urgent need for reducing carbon dioxide emissions if an event similar to the PTME is to be prevented from occurring.[310]

See also edit

References edit

  1. ^ a b Rohde RA, Muller, RA (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. ^ a b c 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 2024-03-26.
  4. ^ Jouault, Corentin; Nel, André; Perrichot, Vincent; Legendre, Frédéric; Condamine, Fabien L. (6 December 2011). "Multiple drivers and lineage-specific insect extinctions during the Permo-Triassic". Nature Communications. 13 (1): 7512. doi:10.1038/s41467-022-35284-4. PMC 9726944. PMID 36473862.
  5. ^ a b Delfini, Massimo; Kustatscher, Evelyn; Lavezzi, Fabrizio; Bernardi, Massimo (29 July 2021). "The End-Permian Mass Extinction: Nature's Revolution". In Martinetto, Edoardo; Tschopp, Emanuel; Gastaldo, Robert A. (eds.). Nature through Time. Springer Textbooks in Earth Sciences, Geography and Environment. Springer Cham. pp. 253–267. doi:10.1007/978-3-030-35058-1_10. ISBN 978-3-030-35060-4. S2CID 226405085.
  6. ^ . National Geographic. 23 November 2011. Archived from the original on November 24, 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. ^ Algeo, Thomas J. (5 February 2012). . Astrobiology Magazine. Archived from the original on 2021-03-08.{{cite journal}}: CS1 maint: unfit URL (link)
  9. ^ 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.
  10. ^ a b Erwin, D.H. (1990). "The End-Permian Mass Extinction". Annual Review of Ecology, Evolution, and Systematics. 21: 69–91. doi:10.1146/annurev.es.21.110190.000441.
  11. ^ 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.
  12. ^ Stanley, Steven M. (18 October 2016). "Estimates of the magnitudes of major marine mass extinctions in earth history". Proceedings of the National Academy of Sciences of the United States of America. 113 (42): E6325–E6334. Bibcode:2016PNAS..113E6325S. doi:10.1073/pnas.1613094113. ISSN 0027-8424. PMC 5081622. PMID 27698119.
  13. ^ 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.
  14. ^ Bergstrom, Carl T.; Dugatkin, Lee Alan (2012). Evolution. Norton. p. 515. ISBN 978-0-393-92592-0.
  15. ^ 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.
  16. ^ a b 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
  17. ^ Marshall, Charles R. (5 January 2023). "Forty years later: The status of the "Big Five" mass extinctions". Cambridge Prisms: Extinction. 1: 1–13. doi:10.1017/ext.2022.4. S2CID 255710815.
  18. ^ a b c d Jin, Y. G.; Wang, Y.; Wang, W.; Shang, Q. H.; Cao, C. Q.; Erwin, D. H. (21 July 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. Retrieved 5 March 2023.
  19. ^ a b Burgess, Seth D.; Bowring, Samuel A. (1 August 2015). "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.
  20. ^ Hulse, D; Lau, K.V.; Sebastiaan, J.V.; Arndt, S; Meyer, K.M.; Ridgwell, A (28 Oct 2021). "End-Permian marine extinction due to temperature-driven nutrient recycling and euxinia". Nat Geosci. 14 (11): 862–867. Bibcode:2021NatGe..14..862H. doi:10.1038/s41561-021-00829-7. S2CID 240076553.
  21. ^ a b 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.
  22. ^ a b c d e 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. Retrieved 2024-03-26.
  23. ^ a b Frank, T. D.; Fielding, Christopher R.; Winguth, A. M. E.; Savatic, K.; Tevyaw, A.; Winguth, C.; McLoughlin, Stephen; Vajda, Vivi; Mays, C.; Nicoll, R.; Bocking, M.; Crowley, J. L. (19 May 2021). "Pace, magnitude, and nature of terrestrial climate change through the end-Permian extinction in southeastern Gondwana". Geology. 49 (9): 1089–1095. Bibcode:2021Geo....49.1089F. doi:10.1130/G48795.1. S2CID 236381390. Retrieved 2024-03-26.
  24. ^ a b Joachimski, Michael M.; Lai, Xulong; Shen, Shuzhong; Jiang, Haishui; Luo, Genming; Chen, Bo; Chen, Jun; Sun, Yadong (1 March 2012). "Climate warming in the latest Permian and the Permian–Triassic mass extinction". Geology. 40 (3): 195–198. Bibcode:2012Geo....40..195J. doi:10.1130/G32707.1. Retrieved 2024-03-26.
  25. ^ a b 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.
  26. ^ a b 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.
  27. ^ a b 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.
  28. ^ a b 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.
  29. ^ 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 of the United States of America. 99 (7): 4172–4177. Bibcode:2002PNAS...99.4172B. doi:10.1073/pnas.032095199. PMC 123621. PMID 11917102.
  30. ^ 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.
  31. ^ a b Rothman, D.H.; Fournier, G.P.; French, K.L.; Alm, E.J.; Boyle, E.A.; Cao, C.; Summons, R.E. (31 March 2014). "Methanogenic burst in the end-Permian carbon cycle". Proceedings of the National Academy of Sciences of the United States of America. 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.
  32. ^ a b Saitoh, Masafumi; Isozaki, Yukio (5 February 2021). "Carbon Isotope Chemostratigraphy Across the Permian-Triassic Boundary at Chaotian, China: Implications for the Global Methane Cycle in the Aftermath of the Extinction". Frontiers in Earth Science. 8: 665. Bibcode:2021FrEaS...8..665S. doi:10.3389/feart.2020.596178.
  33. ^ a b Tohver, Eric; Lana, Cris; Cawood, P.A.; Fletcher, I.R.; Jourdan, F.; Sherlock, S.; et al. (1 June 2012). "Geochronological constraints on the age of a Permo–Triassic impact event: U–Pb and 40Ar / 39Ar results for the 40 km Araguainha structure of central Brazil". Geochimica et Cosmochimica Acta. 86: 214–227. Bibcode:2012GeCoA..86..214T. doi:10.1016/j.gca.2012.03.005.
  34. ^ a b c Tohver, Eric; Cawood, P. A.; Riccomini, Claudio; Lana, Cris; Trindade, R. I. F. (1 October 2013). "Shaking a methane fizz: Seismicity from the Araguainha impact event and the Permian–Triassic global carbon isotope record". Palaeogeography, Palaeoclimatology, Palaeoecology. 387: 66–75. Bibcode:2013PPP...387...66T. doi:10.1016/j.palaeo.2013.07.010. Retrieved 2024-03-26.
  35. ^ a b Tohver, Eric; Schmieder, Martin; Lana, Cris; Mendes, Pedro S. T.; Jourdan, Fred; Warren, Lucas; Riccomini, Claudio (2 January 2018). "End-Permian impactogenic earthquake and tsunami deposits in the intracratonic Paraná Basin of Brazil". Geological Society of America Bulletin. 130 (7–8): 1099–1120. Bibcode:2018GSAB..130.1099T. doi:10.1130/B31626.1. Retrieved 2024-03-26.
  36. ^ a b Liu, Feng; Peng, Huiping; Marshall, John E. A.; Lomax, Barry H.; Bomfleur, Benjamin; Kent, Matthew S.; Fraser, Wesley T.; Jardine, Phillip E. (6 January 2023). "Dying in the Sun: Direct evidence for elevated UV-B radiation at the end-Permian mass extinction". Science Advances. 9 (1): eabo6102. Bibcode:2023SciA....9O6102L. doi:10.1126/sciadv.abo6102. PMC 9821938. PMID 36608140.
  37. ^ a b Benca, Jeffrey P.; Duijnstee, Ivo A. P.; Looy, Cindy V. (7 February 2018). "UV-B–induced forest sterility: Implications of ozone shield failure in Earth's largest extinction". Science Advances. 4 (2): e1700618. Bibcode:2018SciA....4..618B. doi:10.1126/sciadv.1700618. PMC 5810612. PMID 29441357.
  38. ^ a b c d Visscher, Henk; Looy, Cindy V.; Collinson, Margaret E.; Brinkhuis, Henk; Cittert, Johanna H.A. van Konijnenburg; Kürschner, Wolfram M.; Sephton, Mark A. (31 August 2004). "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.
  39. ^ Dai, X.; Davies, J. H. F. L.; Yuan, Z.; Brayard, A.; Ovtcharova, M.; Xu, G.; Liu, X.; Smith, C. P. A.; Schweitzer, C. E.; Li, M.; Perrot, M. G.; Jiang, S.; Miao, L.; Cao, Y.; Yan, J.; Bai, R.; Wang, F.; Guo, W.; Song, H.; Tian, L.; Dal Corso, J.; Liu, Y.; Chu, D.; Song, H. (2023). "A Mesozoic fossil lagerstätte from 250.8 million years ago shows a modern-type marine ecosystem". Science. 379 (6632): 567–572. doi:10.1126/science.adf1622. PMID 36758082.
  40. ^ 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 (2017). "Unexpected Early Triassic marine ecosystem and the rise of the Modern evolutionary fauna". Science Advances. 3 (2): e1602159. Bibcode:2017SciA....3E2159B. doi:10.1126/sciadv.1602159. PMC 5310825. PMID 28246643.
  41. ^ 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.
  42. ^ a b c d e f g h i j k l 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.
  43. ^ 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.
  44. ^ a b c d Erwin, D.H. (1993). The great Paleozoic crisis; Life and death in the Permian. Columbia University Press. ISBN 978-0-231-07467-4.
  45. ^ 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.
  46. ^ 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. World and Regional Geology. Cambridge: Cambridge University Press. pp. 1–7. doi:10.1017/CBO9780511529498.002. ISBN 978-0-521-54573-0. Retrieved 3 July 2023.
  47. ^ a b Burgess, Seth D.; Bowring, Samuel; Shen, Shu-zhong (10 February 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.
  48. ^ Yuan, Dong-xun; Shen, Shu-zhong; Henderson, Charles M.; Chen, Jun; Zhang, Hua; Feng, Hong-zhen (1 September 2014). "Revised conodont-based integrated high-resolution timescale for the Changhsingian Stage and end-Permian extinction interval at the Meishan sections, South China". Lithos. 204: 220–245. Bibcode:2014Litho.204..220Y. doi:10.1016/j.lithos.2014.03.026. Retrieved 10 August 2023.
  49. ^ 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.
  50. ^ Musashi, Masaaki; Isozaki, Yukio; Koike, Toshio; Kreulen, Rob (30 August 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. Retrieved 3 July 2023.
  51. ^ 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.
  52. ^ Korte, Christoph; Kozur, Heinz W. (9 September 2010). "Carbon-isotope stratigraphy across the Permian–Triassic boundary: A review". Journal of Asian Earth Sciences. 39 (4): 215–235. Bibcode:2010JAESc..39..215K. doi:10.1016/j.jseaes.2010.01.005. Retrieved 26 June 2023.
  53. ^ Li, Rong; Jones, Brian (15 February 2017). "Diagenetic overprint on negative δ13C excursions across the Permian/Triassic boundary: A case study from Meishan section, China". Palaeogeography, Palaeoclimatology, Palaeoecology. 468: 18–33. Bibcode:2017PPP...468...18L. doi:10.1016/j.palaeo.2016.11.044. ISSN 0031-0182. Retrieved 20 September 2023.
  54. ^ Schobben, Martin; Heuer, Franziska; Tietje, Melanie; Ghaderi, Abbas; Korn, Dieter; Korte, Christoph; Wignall, Paul B. (19 November 2018). "Chemostratigraphy Across the Permian-Triassic Boundary: The Effect of Sampling Strategies on Carbonate Carbon Isotope Stratigraphic Markers". In Sial, Alcides N.; Gaucher, Claudio; Ramkumar, Muthuvairavasamy; Pinto Ferreira, Valderez (eds.). Chemostratigraphy Across Major Chronological Boundaries. American Geophysical Union. pp. 159–181. doi:10.1002/9781119382508.ch9. ISBN 9781119382508. S2CID 134060610. Retrieved 20 September 2023.
  55. ^ "Daily CO2". Mauna Loa Observatory.
  56. ^ 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 of the United States of America. 93 (5): 2155–2158. Bibcode:1996PNAS...93.2155V. doi:10.1073/pnas.93.5.2155. PMC 39926. PMID 11607638.
  57. ^ Tewari, Rajni; Awatar, Ram; Pandita, Sundeep K.; McLoughlin, Stephen D.; Agnihotri, Deepa; Pillai, Suresh S. K.; et al. (October 2015). "The Permian–Triassic palynological transition in the Guryul Ravine section, Kashmir, India: implications for Tethyan–Gondwanan correlations". Earth-Science Reviews. 149: 53–66. Bibcode:2015ESRv..149...53T. doi:10.1016/j.earscirev.2014.08.018. Retrieved 26 May 2023.
  58. ^ Foster CB, Stephenson MH, Marshall C, Logan GA, Greenwood PF (2002). "A revision of Reduviasporonites Wilson 1962: Description, illustration, comparison and biological affinities". Palynology. 26 (1): 35–58. Bibcode:2002Paly...26...35F. doi:10.2113/0260035.
  59. ^ 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. Bibcode:2010RPaPa.162..522D. doi:10.1016/j.revpalbo.2010.09.003. Retrieved 8 January 2023.
  60. ^ López-Gómez J, Taylor EL (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.
  61. ^ 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 of the United States of America. 98 (4): 7879–7883. Bibcode:2001PNAS...98.7879L. doi:10.1073/pnas.131218098. PMC 35436. PMID 11427710. See image 2
  62. ^ 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.
  63. ^ Retallack GJ, Smith RMH, Ward PD (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.
  64. ^ Sephton MA, Visscher H, Looy CV, Verchovsky AB, Watson JS (2009). "Chemical constitution of a Permian-Triassic disaster species". Geology. 37 (10): 875–878. Bibcode:2009Geo....37..875S. doi:10.1130/G30096A.1.
  65. ^ 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. Bibcode:2018GeoFr...9..147R. doi:10.1016/j.gsf.2017.06.005. Retrieved 24 December 2022.
  66. ^ Yin, Hongfu; Jiang, Haishui; Xia, Wenchen; Feng, Qinglai; Zhang, Ning; Shen, Jun (October 2014). "The end-Permian regression in South China and its implication on mass extinction". Earth-Science Reviews. 137: 19–33. Bibcode:2014ESRv..137...19Y. doi:10.1016/j.earscirev.2013.06.003. Retrieved 20 September 2023.
  67. ^ de Wit, Maarten J.; Ghosh, Joy G.; de Villiers, Stephanie; Rakotosolofo, Nicolas; Alexander, James; Tripathi, Archana; Looy, Cindy (March 2002). "Multiple Organic Carbon Isotope Reversals across the Permo-Triassic Boundary of Terrestrial Gondwana Sequences: Clues to Extinction Patterns and Delayed Ecosystem Recovery". The Journal of Geology. 110 (2): 227–240. Bibcode:2002JG....110..227D. doi:10.1086/338411. ISSN 0022-1376. S2CID 129653925. Retrieved 20 September 2023.
  68. ^ 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.
  69. ^ 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.
  70. ^ Li, Guoshan; Liao, Wei; Li, Sheng; Wang, Yongbiao; Lai, Zhongping (23 March 2021). "Different triggers for the two pulses of mass extinction across the Permian and Triassic boundary". Scientific Reports. 11 (1): 6686. Bibcode:2021NatSR..11.6686L. doi:10.1038/s41598-021-86111-7. PMC 7988102. PMID 33758284.
  71. ^ Shen, Jiaheng; Zhang, Yi Ge; Yang, Huan; Xie, Shucheng; Pearson, Ann (3 October 2022). "Early and late phases of the Permian–Triassic mass extinction marked by different atmospheric CO2 regimes". Nature Geoscience. 15 (1): 839–844. Bibcode:2022NatGe..15..839S. doi:10.1038/s41561-022-01034-w. S2CID 252697822. Retrieved 20 April 2023.
  72. ^ Wang SC, Everson PJ (2007). "Confidence intervals for pulsed mass extinction events". Paleobiology. 33 (2): 324–336. Bibcode:2007Pbio...33..324W. doi:10.1666/06056.1. S2CID 2729020.
  73. ^ 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...59410936F. doi:10.1016/j.palaeo.2022.110936. S2CID 247514266.
  74. ^ Huang, Yuangeng; Chen, Zhong-Qiang; Roopnarine, Peter D.; Benton, Michael James; Zhao, Laishi; Feng, Xueqian; Li, Zhenhua (24 February 2023). "The stability and collapse of marine ecosystems during the Permian-Triassic mass extinction". Current Biology. 33 (6): 1059–1070.e4. doi:10.1016/j.cub.2023.02.007. PMID 36841237. S2CID 257186215.
  75. ^ 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.
  76. ^ Shen, Shu-Zhong; Crowley, James L.; Wang, Yue; Bowring, Samuel R.; Erwin, Douglas H.; Sadler, Peter M.; et al. (17 November 2011). "Calibrating the End-Permian Mass Extinction". Science. 334 (6061): 1367–1372. Bibcode:2011Sci...334.1367S. doi:10.1126/science.1213454. PMID 22096103. S2CID 970244. Retrieved 26 May 2023.
  77. ^ Metcalfe, I.; Crowley, J. L.; Nicoll, Robert S.; Schmitz, M. (August 2015). "High-precision U-Pb CA-TIMS calibration of Middle Permian to Lower Triassic sequences, mass extinction and extreme climate-change in eastern Australian Gondwana". Gondwana Research. 28 (1): 61–81. Bibcode:2015GondR..28...61M. doi:10.1016/j.gr.2014.09.002. Retrieved 31 May 2023.
  78. ^ Dal Corso, Jacopo; Song, Haijun; Callegaro, Sara; Chu, Daoliang; Sun, Yadong; Hilton, Jason; et al. (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. hdl:10852/100010. S2CID 247013868. Retrieved 20 December 2022.
  79. ^ Chen, Zhong-Qiang; Harper, David A. T.; Grasby, Stephen; Zhang, Lei (1 August 2022). "Catastrophic event sequences across the Permian-Triassic boundary in the ocean and on land". Global and Planetary Change. 215: 103890. Bibcode:2022GPC...21503890C. doi:10.1016/j.gloplacha.2022.103890. ISSN 0921-8181. S2CID 250417358. Retrieved 24 November 2023.
  80. ^ Hermann, Elke; Hochuli, Peter A.; Bucher, Hugo; Vigran, Jorunn O.; Weissert, Helmut; Bernasconi, Stefano M. (December 2010). "A close-up view of the Permian–Triassic boundary based on expanded organic carbon isotope records from Norway (Trøndelag and Finnmark Platform)". Global and Planetary Change. 74 (3–4): 156–167. Bibcode:2010GPC....74..156H. doi:10.1016/j.gloplacha.2010.10.007. Retrieved 10 August 2023.
  81. ^ Gastaldo, Robert A.; Kamo, Sandra L.; Neveling, Johann; Geissman, John W.; Bamford, Marion; Looy, Cindy V. (1 October 2015). "Is the vertebrate-defined Permian-Triassic boundary in the Karoo Basin, South Africa, the terrestrial expression of the end-Permian marine event?". Geology. 43 (10): 939–942. Bibcode:2015Geo....43..939G. doi:10.1130/G37040.1. S2CID 129258297.
  82. ^ Zhang, Peixin; Yang, Minfang; Lu, Jing; Bond, David P. G.; Zhou, Kai; Xu, Xiaotao; et al. (March 2023). "End-Permian terrestrial ecosystem collapse in North China: Evidence from palynology and geochemistry". Global and Planetary Change. 222: 104070. Bibcode:2023GPC...22204070Z. doi:10.1016/j.gloplacha.2023.104070. S2CID 256920630.
  83. ^ Wu, Qiong; Zhang, Hua; Ramezani, Jahandar; Zhang, Fei-fei; Erwin, Douglas H.; Feng, Zhuo; Shao, Long-yi; Cai, Yao-feng; Zhang, Shu-han; Xu, Yi-gang; Shen, Shu-zhong (2 February 2024). "The terrestrial end-Permian mass extinction in the paleotropics postdates the marine extinction". Science Advances. 10 (5): eadi7284. doi:10.1126/sciadv.adi7284. ISSN 2375-2548. PMC 10830061. PMID 38295161.
  84. ^ Bond DPG, Wignall PB, Wang W, Izon G, Jiang HS, Lai XL, 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.
  85. ^ a b Retallack GJ, Metzger CA, Greaver T, Jahren AH, Smith RMH, Sheldon ND (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.
  86. ^ Li, Dirson Jian (18 December 2012). "Tectonic cause of mass extinctions and the genomic contribution to biodiversification". arXiv:1212.4229 [q-bio.PE].
  87. ^ 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.
  88. ^ 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.
  89. ^ Shen S, Shi GR (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.
  90. ^ Wang X-D, Sugiyama T (December 2000). "Diversity and extinction patterns of Permian coral faunas of China". Lethaia. 33 (4): 285–294. Bibcode:2000Letha..33..285W. doi:10.1080/002411600750053853.
  91. ^ 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.
  92. ^ 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.
  93. ^ 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.
  94. ^ Song, Haijun; Wu, Yuyang; Dai, Xu; Dal Corso, Jacopo; Wang, Fengyu; Feng, Yan; Chu, Daoliang; Tian, Li; Song, Huyue; Foster, William J. (6 May 2024). "Respiratory protein-driven selectivity during the Permian-Triassic mass extinction". The Innovation. 5 (3): 100618. doi:10.1016/j.xinn.2024.100618. Retrieved 21 May 2024.
  95. ^ Yan, Jia; Song, Haijun; Dai, Xu (1 February 2023). "Increased bivalve cosmopolitanism during the mid-Phanerozoic mass extinctions". Palaeogeography, Palaeoclimatology, Palaeoecology. 611: 111362. Bibcode:2023PPP...61111362Y. doi:10.1016/j.palaeo.2022.111362. Retrieved 20 February 2023.
  96. ^ Allen, Bethany J.; Clapham, Matthew E.; Saupe, Erin E.; Wignall, Paul B.; Hill, Daniel J.; Dunhill, Alexander M. (10 February 2023). "Estimating spatial variation in origination and extinction in deep time: a case study using the Permian–Triassic marine invertebrate fossil record". Paleobiology. 49 (3): 509–526. Bibcode:2023Pbio...49..509A. doi:10.1017/pab.2023.1. hdl:20.500.11850/598054. S2CID 256801383.
  97. ^ Stanley, S. M. (2008). "Predation defeats competition on the seafloor". Paleobiology. 34 (1): 1–21. Bibcode:2008Pbio...34....1S. doi:10.1666/07026.1. S2CID 83713101. Retrieved 2008-05-13.
  98. ^ Stanley, S. M. (2007). "An Analysis of the History of Marine Animal Diversity". Paleobiology. 33 (sp6): 1–55. Bibcode:2007Pbio...33R...1S. doi:10.1666/06020.1. S2CID 86014119.
  99. ^ 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.
  100. ^ 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. Bibcode:2015PPP...428....1H. doi:10.1016/j.palaeo.2015.03.036. Retrieved 19 December 2022.
  101. ^ "Permian: The Marine Realm and The End-Permian Extinction". paleobiology.si.edu. Retrieved 2016-01-26.
  102. ^ a b "Permian extinction". Encyclopædia Britannica. Retrieved 2016-01-26.
  103. ^ a b c d e 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.
  104. ^ 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.
  105. ^ 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.
  106. ^ Xue, Chunling; Yuan, Dong-xun; Chen, Yanlong; Stubbs, Thomas L.; Zhao, Yueli; Zhang, Zhifei (15 May 2024). "Morphological innovation after mass extinction events in Permian and Early Triassic conodonts based on Polygnathacea". Palaeogeography, Palaeoclimatology, Palaeoecology. 642: 112149. doi:10.1016/j.palaeo.2024.112149. Retrieved 21 May 2024 – via Elsevier Science Direct.
  107. ^ 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.
  108. ^ 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. Bibcode:2008Pbio...34..128S. doi:10.1666/07053.1. S2CID 83650272.
  109. ^ Crasquin, Sylvie; Forel, Marie-Béatrice; Qinglai, Feng; Aihua, Yuan; Baudin, François; Collin, Pierre-Yves (30 July 2010). "Ostracods (Crustacea) through the Permian-Triassic boundary in South China: the Meishan stratotype (Zhejiang Province)". Journal of Systematic Palaeontology. 8 (3): 331–370. Bibcode:2010JSPal...8..331C. doi:10.1080/14772011003784992. S2CID 85986762. Retrieved 3 July 2023.
  110. ^ a b Forel, Marie-Béatrice (1 August 2012). "Ostracods (Crustacea) associated with microbialites across the Permian-Triassic boundary in Dajiang (Guizhou Province, South China)". European Journal of Taxonomy (19): 1–34. doi:10.5852/ejt.2012.19. Retrieved 3 July 2023.
  111. ^ Powers, Catherine M.; Bottjer, David J. (1 November 2007). "Bryozoan paleoecology indicates mid-Phanerozoic extinctions were the product of long-term environmental stress". Geology. 35 (11): 995. Bibcode:2007Geo....35..995P. doi:10.1130/G23858A.1. ISSN 0091-7613. Retrieved 30 December 2023.
  112. ^ Liu, Guichun; Feng, Qinglai; Shen, Jun; Yu, Jianxin; He, Weihong; Algeo, Thomas J. (1 August 2013). "Decline of Siliceous Sponges and Spicule Miniaturization Induced by Marine Productivity Collapse and Expanding Anoxia During the Permian-Triassic Crisis in South China". PALAIOS. 28 (8): 664–679. doi:10.2110/palo.2013.p13-035r. S2CID 128751510. Retrieved 31 May 2023.
  113. ^ Song, Haijun; Tong, Jinnan; Chen, Zhong-Qiang; Yang, Hao; Wang, Yongbiao (14 July 2015). "End-Permian mass extinction of foraminifers in the Nanpanjiang basin, South China". Journal of Paleontology. 83 (5): 718–738. doi:10.1666/08-175.1. S2CID 130765890. Retrieved 23 March 2023.
  114. ^ Groves, John R.; Altiner, Demír; Rettori, Roberto (11 August 2017). "Extinction, Survival, and Recovery of Lagenide Foraminifers in the Permian–Triassic Boundary Interval, Central Taurides, Turkey". Journal of Paleontology. 79 (S62): 1–38. doi:10.1666/0022-3360(2005)79[1:ESAROL]2.0.CO;2. S2CID 130620805. Retrieved 23 March 2023.
  115. ^ Guinot, Guillaume; Adnet, Sylvain; Cavin, Lionel; Cappetta, Henri (29 October 2013). "Cretaceous stem chondrichthyans survived the end-Permian mass extinction". Nature Communications. 4: 2669. Bibcode:2013NatCo...4.2669G. doi:10.1038/ncomms3669. PMID 24169620. S2CID 205320689.
  116. ^ Kear, Benjamin P.; Engelschiøn, Victoria S.; Hammer, Øyvind; Roberts, Aubrey J.; Hurum, Jørn H. (13 March 2023). "Earliest Triassic ichthyosaur fossils push back oceanic reptile origins". Current Biology. 33 (5): R178–R179. doi:10.1016/j.cub.2022.12.053. PMID 36917937. S2CID 257498390.
  117. ^ 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.
  118. ^ 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.
  119. ^ Feng, Yan; Song, Haijun; Bond, David P. G. (1 November 2020). "Size variations in foraminifers from the early Permian to the Late Triassic: implications for the Guadalupian–Lopingian and the Permian–Triassic mass extinctions". Paleobiology. 46 (4): 511–532. Bibcode:2020Pbio...46..511F. doi:10.1017/pab.2020.37. S2CID 224855811.
  120. ^ 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.
  121. ^ Rego, Brianna L.; Wang, Steve C.; Altiner, Demír; Payne, Jonathan L. (8 February 2016). "Within- and among-genus components of size evolution during mass extinction, recovery, and background intervals: a case study of Late Permian through Late Triassic foraminifera". Paleobiology. 38 (4): 627–643. doi:10.1666/11040.1. S2CID 17284750. Retrieved 23 March 2023.
  122. ^ 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. Bibcode:2010Gbio....8..391H. doi:10.1111/j.1472-4669.2010.00248.x. PMID 20550584. S2CID 23096333. Retrieved 13 January 2023.
  123. ^ 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.
  124. ^ Shi, Guang R.; Zhang, Yi-chun; Shen, Shu-zhong; He, Wei-hong (15 April 2016). "Nearshore–offshore–basin species diversity and body size variation patterns in Late Permian (Changhsingian) brachiopods". Palaeogeography, Palaeoclimatology, Palaeoecology. 448: 96–107. Bibcode:2016PPP...448...96S. doi:10.1016/j.palaeo.2015.07.046. hdl:10536/DRO/DU:30080099.
  125. ^ 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...60911307H. doi:10.1016/j.palaeo.2022.111307. S2CID 253368808. Retrieved 13 January 2023.
  126. ^ 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.
  127. ^ 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.
  128. ^ 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.
  129. ^ Forel, Marie-Béatrice; Crasquin, Sylvie; Chitnarin, Anisong; Angiolini, Lucia; Gaetani, Maurizio (25 February 2015). "Precocious sexual dimorphism and the Lilliput effect in Neo-Tethyan Ostracoda (Crustacea) through the Permian–Triassic boundary". Palaeontology. 58 (3): 409–454. Bibcode:2015Palgy..58..409F. doi:10.1111/pala.12151. hdl:2434/354270. S2CID 140198533. Retrieved 23 May 2023.
  130. ^ Payne, Jonathan L. (8 April 2016). "Evolutionary dynamics of gastropod size across the end-Permian extinction and through the Triassic recovery interval". Paleobiology. 31 (2): 269–290. doi:10.1666/0094-8373(2005)031[0269:EDOGSA]2.0.CO;2. S2CID 7367192. Retrieved 23 March 2023.
  131. ^ 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.
  132. ^ 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.
  133. ^ 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.
  134. ^ Brayard, Arnaud; Meier, Maximiliano; Escarguel, Gilles; Fara, Emmanuel; Nützel, Alexander; Olivier, Nicolas; Bylund, Kevin G.; Jenks, James F.; Stephen, Daniel A.; Hautmann, Michael; Vennin, Emmanuelle; Bucher, Hugo (July 2015). "Early Triassic Gulliver gastropods: Spatio-temporal distribution and significance for biotic recovery after the end-Permian mass extinction". Earth-Science Reviews. 146: 31–64. Bibcode:2015ESRv..146...31B. doi:10.1016/j.earscirev.2015.03.005. Retrieved 20 January 2023.
  135. ^ Atkinson, Jed W.; Wignall, Paul B.; Morton, Jacob D.; Aze, Tracy (9 January 2019). "Body size changes in bivalves of the family Limidae in the aftermath of the end-Triassic mass extinction: the Brobdingnag effect". Palaeontology. 62 (4): 561–582. Bibcode:2019Palgy..62..561A. doi:10.1111/pala.12415. S2CID 134070316. Retrieved 20 January 2023.
  136. ^ 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.
  137. ^ Sole, R. V.; Newman, M. (2003). "Extinctions and Biodiversity in the Fossil Record". In Canadell, J. G.; Mooney, H. A. (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.
  138. ^ Rees, P. McAllister (1 September 2002). "Land-plant diversity and the end-Permian mass extinction". Geology. 30 (9): 827–830. Bibcode:2002Geo....30..827M. doi:10.1130/0091-7613(2002)030<0827:LPDATE>2.0.CO;2. Retrieved 31 May 2023.
  139. ^ a b 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:
permian, triassic, extinction, event, great, dying, redirects, here, other, uses, great, dying, disambiguation, approximately, million, years, permian, triassic, extinction, event, ptme, also, known, late, permian, extinction, event, latest, permian, extinctio. Great Dying redirects here For other uses see Great Dying disambiguation Approximately 251 9 million years ago the Permian Triassic P T P Tr extinction event PTME also known as the Late Permian extinction event 3 the Latest Permian extinction event 4 the End Permian extinction event 5 6 and colloquially as the Great Dying 7 8 forms the boundary between the Permian and Triassic geologic periods and with them the Paleozoic and Mesozoic eras 9 It is the Earth s most severe known extinction event 10 11 with the extinction of 57 of biological families 83 of genera 81 of marine species 12 13 14 and 70 of terrestrial vertebrate species 15 It is also the greatest known mass extinction of 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 Marine extinction intensity during 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 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 oxygen starved sulfurous oceans 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 Several other contributing factors have been proposed including the emission of carbon dioxide from the burning of oil and coal deposits ignited by the eruptions 27 28 emissions of methane from the gasification of methane clathrates 29 emissions of methane by novel methanogenic microorganisms nourished by minerals dispersed in the eruptions 30 31 32 and an extraterrestrial impact which created the Araguainha crater and caused seismic release of methane 33 34 35 and the destruction of the ozone layer with increased exposure to solar radiation 36 37 38 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 3 1 Synapsids 3 3 2 Sauropsids 3 3 3 Temnospondyls 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 1 3 Indochina South China subduction zone volcanic arc 4 2 Methane clathrate gasification 4 3 Hypercapnia and acidification 4 4 Anoxia and euxinia 4 5 Aridification 4 6 Ozone depletion 4 7 Asteroid impact 4 7 1 Possible impact sites 4 8 Methanogens 4 9 Supercontinent Pangaea 4 10 Interstellar dust 5 Comparison to present global warming 6 See also 7 References 8 Further reading 9 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 41 Full recovery of woody trees 42 Corals return 43 Scleractiniancorals amp calcified sponges return 41 First complex marine ecosystems Guiyang and Paris biota 39 40 End Capitanian extinctionMesozoicPaleozoicPermianTriassicAn 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 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 47 A large abrupt global decrease in d13C the ratio of the stable isotope 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 42 and an increase in CO2 levels to 2 500 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 fungi caused by the sharp increase in the amount of dead plants and animals fed upon by the fungi 56 This fungal spike has been used by some paleontologists to identify a lithological sequence as being on or very close to the Permian Triassic boundary in rocks that are unsuitable for radiometric dating or have a lack of suitable index fossils 57 However 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 56 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 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 Trondelag shows the terrestrial floral turnover occurred before the large negative d13C 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 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 84 Many of the extinctions once dated to the Permian Triassic boundary have more 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 This older theory still supported in some recent papers 15 85 proposes that there were two major extinction pulses 9 4 million years apart separated by a period of extinctions that were less extensive but still well above the background level and that the final extinction killed off only about 80 of marine species alive at that time whereas 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 86 15 87 For example all dinocephalian genera died out at the end of the Guadalupian 85 as did the Verbeekinidae a family of large size 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 editMarine 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 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 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 These organisms were susceptible to the effects of the ocean acidification that resulted from increased atmospheric CO2 93 Organisms that relied on haemocyanin or haemoglobin for transporting oxygen were more resistant to extinction than those utilising haemerythrin or oxygen diffusion 94 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 95 There was little latitudinal difference in the survival rates of taxa 96 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 97 98 The extinction rate of marine organisms was catastrophic 18 99 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 100 Surviving marine invertebrate groups included articulate brachiopods those with a hinge 101 which had undergone a slow decline in numbers since the P Tr extinction the Ceratitida order of ammonites 102 and crinoids sea lilies 102 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 103 104 In the case of the brachiopods at least surviving taxa were generally small rare members of a formerly diverse community 105 Conodonts were severely affected both in terms of taxonomic and morphological diversity although not as severely as during the Capitanian mass extinction 106 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 107 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 108 Ostracods experienced prolonged diversity perturbations during the Changhsingian before the PTME proper when immense proportions of them abruptly vanished 109 At least 74 of ostracods died out during the PTME itself 110 Bryozoans had been on a long term decline throughout the Late Permian epoch before they suffered even more catastrophic losses during the PTME 111 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 112 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 113 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 114 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 Early Cretaceous cladodontomorphs in deep outer shelf environments 115 Ichthyosaurs which are believed to have evolved immediately before the PTME were also PTME survivors 116 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 117 118 notably occurring in foraminifera 119 120 121 brachiopods 122 123 124 bivalves 125 126 127 and ostracods 128 129 Though gastropods that survived the cataclysm were smaller in size than those that did not 130 it remains debated whether the Lilliput effect truly took hold among gastropods 131 132 133 Some gastropod taxa termed Gulliver gastropods ballooned in size during and immediately following the mass extinction 134 exemplifying the Lilliput effect s opposite which has been dubbed the Brobdingnag effect 135 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 16 according to some sources it may well be the only mass extinction to significantly affect insect diversity 136 137 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 edit 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 138 Plants are relatively immune to mass extinction with the impact of all the major mass extinctions insignificant at a family level 42 dubious discuss Even the reduction observed in species diversity of 50 may be mostly due to taphonomic processes 139 42 However a massive rearrangement of ecosystems does occur with plant abundances and distributions changing profoundly and all the forests virtually disappearing 140 42 The dominant floral groups changed with many groups of land plants entering abrupt decline such as Cordaites gymnosperms and Glossopteris seed ferns 141 142 The severity of plant extinction has been disputed 143 139 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 144 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 61 The Cordaites flora which dominated the Angaran floristic realm corresponding to Siberia collapsed over the course of the extinction 145 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 146 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 147 In South China the subtropical Cathaysian gigantopterid dominated rainforests abruptly collapsed 148 149 150 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 151 Wildfires too likely played a role in the fall of Gigantopteris 152 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 153 Terrestrial vertebrates edit The terrestrial vertebrate extinction occurred rapidly taking 50 000 years or less 154 Aridification induced by global warming was the chief culprit behind terrestrial vertebrate extinctions 155 156 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 157 10 Gorgonopsians are traditionally thought to have gone extinct during the PTME but some tentative evidence suggests they may have survived into the Triassic 158 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 159 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 160 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 161 Another study of latest Permian vertebrates in the Karoo Basin found that 54 of them went extinct due to the PTME 162 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 Palaeozoic evolutionary fauna declined while the modern evolutionary fauna achieved greater dominance 163 the Permian Triassic mass extinction marked a key turning point in this ecological shift that began after the Capitanian mass extinction 164 and culminated in the Late Jurassic 165 Typical taxa of shelly benthic faunas were now bivalves snails sea urchins and Malacostraca whereas bony fishes 166 and marine reptiles 167 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 168 169 However recovery was also differential between taxa Some survivors became extinct some million years after the extinction event without having rediversified dead clade walking 170 e g the snail family Bellerophontidae 171 whereas others rose to dominance over geologic times e g bivalves 172 173 Marine ecosystems edit nbsp Shell bed with the bivalve Claraia clarai a common early Triassic disaster taxon A cosmopolitanism event began immediately after the end Permian extinction event 174 Marine post extinction faunas were mostly species poor and were dominated by few disaster taxa such as the bivalves Claraia Unionites Eumorphotis and Promyalina 175 the conodonts Clarkina and Hindeodus 176 the inarticulate brachiopod Lingularia 175 and the foraminifera Earlandia and Rectocornuspira kalhori 177 the latter of which is sometimes classified under the genus Ammodiscus 178 Their guild diversity was also low 179 The speed of recovery from the extinction is disputed Some scientists estimate that it took 10 million years until the Middle Triassic 180 due to the severity of the extinction However studies in Bear Lake County near Paris Idaho 181 and nearby sites in Idaho and Nevada 182 showed a relatively quick rebound in a localized Early Triassic marine ecosystem Paris biota taking around 1 3 million years to recover 181 while an unusually diverse and complex ichnobiota is known from Italy less than a million years after the end Permian extinction 183 Additionally the complex Guiyang biota found near Guiyang China also indicates life thrived in some places just a million years after the mass extinction 184 185 as does a fossil assemblage known as the Shanggan fauna found in Shanggan China 186 the Wangmo biota from the Luolou Formation of Guizhou 187 and a gastropod fauna from the Al Jil Formation of Oman 188 Regional differences in the pace of biotic recovery existed 189 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 190 191 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 192 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 193 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 194 Epifaunal benthos took longer to recover than infaunal benthos 195 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 196 and conodonts which diversified considerably over the first two million years of the Early Triassic 197 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 198 That recovery was slow in the Early Triassic can be explained by low levels of biological competition due to the paucity of taxonomic diversity 199 and that biotic recovery explosively accelerated in the Anisian can be explained by niche crowding a phenomenon that would have drastically increased competition becoming prevalent by the Anisian 200 Biodiversity rise thus behaved as a positive feedback loop enhancing itself as it took off in the Spathian and Anisian 201 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 199 Other explanations state that life was delayed in its recovery because grim conditions returned periodically over the course of the Early Triassic 202 causing further extinction events such as the Smithian Spathian boundary extinction 11 203 204 Continual episodes of extremely hot climatic conditions during the Early Triassic have been held responsible for the delayed recovery of oceanic life 205 206 in particular skeletonised taxa that are most vulnerable to high carbon dioxide concentrations 207 The relative delay in the recovery of benthic organisms has been attributed to widespread anoxia 208 but high abundances of benthic species contradict this explanation 209 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 191 Recurrent Early Triassic environmental stresses also acted as a ceiling limiting the maximum ecological complexity of marine ecosystems until the Spathian 210 Recovery biotas appear to have been ecologically uneven and unstable into the Anisian making them vulnerable to environmental stresses 211 Whereas most marine communities were fully recovered by the Middle Triassic 212 213 global marine diversity reached pre extinction values no earlier than the Middle Jurassic approximately 75 million years after the extinction event 214 nbsp 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 215 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 215 and the increase in predation pressure and durophagy led to the Mesozoic Marine Revolution 216 Bivalves rapidly recolonised many marine environments in the wake of the catastrophe 217 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 218 Bivalves were once thought to have outcompeted brachiopods but this outdated hypothesis about the brachiopod bivalve transition has been disproven by Bayesian analysis 219 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 165 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 220 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 221 Infaunal habits in bivalves became more common after the PTME 222 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 223 The surviving brachiopod fauna was very low in diversity and exhibited no provincialism whatsoever 224 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 225 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 226 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 227 Ostracods were extremely rare during the basalmost Early Triassic 228 Taxa associated with microbialites were disproportionately represented among ostracod survivors 110 Ostracod recovery began in the Spathian 229 Despite high taxonomic turnover the ecological life modes of Early Triassic ostracods remained rather similar to those of pre PTME ostracods 230 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 231 Crinoids sea lilies suffered a selective extinction resulting in a decrease in the variety of their forms 232 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 233 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 234 Though their taxonomic diversity remained relatively low crinoids regained much of their ecological dominance by the Middle Triassic epoch 220 Stem group echinoids survived the PTME 235 The survival of miocidarid echinoids such as Eotiaris is likely attributable to their ability to thrive in a wide range of environmental conditions 236 Conodonts saw a rapid recovery during the Induan 237 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 238 Segminiplanate conodonts again experienced a brief period of domination in the early Spathian probably related to a transient oxygenation of deep waters 239 Neospathodid conodonts survived the crisis but underwent proteromorphosis 240 In the PTME s aftermath disaster taxa of benthic foraminifera filled many of their vacant niches The recovery of benthic foraminifera was very slow and frequently interrupted until the Spathian 241 Microbial reefs predominated across shallow seas for a short time during the earliest Triassic 242 occurring both in anoxic and oxic waters 243 Polybessurus like microfossils often dominated these earliest Triassic microbialites 244 Microbial metazoan reefs appeared very early in the Early Triassic 245 and they dominated many surviving communities across the recovery from the mass extinction 246 Microbialite deposits appear to have declined in the early Griesbachian synchronously with a significant sea level drop that occurred then 243 Metazoan built reefs reemerged during the Olenekian mainly being composed of sponge biostrome and bivalve builups 246 Keratose sponges were particularly noteworthy in their integral importance to Early Triassic microbial metazoan reef communities 247 248 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 246 Bryozoans after sponges were the most numerous organisms in Tethyan reefs during the Anisian 249 Metazoan reefs became common again during the Anisian because the oceans cooled down then from their overheated state during the Early Triassic 250 Biodiversity amongst metazoan reefs did not recover until well into the Anisian millions of years after non reef ecosystems recovered their diversity 251 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 252 Ichnocoenoses show that marine ecosystems recovered to pre extinction levels of ecological complexity by the late Olenekian 253 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 254 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 255 256 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 257 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 258 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 259 Furthermore severe extinction pulses continued to occur after the Permian Triassic boundary causing additional floral turnovers Gymnosperms recovered within a few thousand years after the Permian Triassic boundary but around 500 000 years after it the Dominant gymnosperm genera were replaced by lycophytes extant lycophytes are recolonizers of disturbed areas during an extinction pulse at the Griesbachian Dienerian boundary 260 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 145 The worldwide recovery of gymnosperm forests took approximately 4 5 million years 261 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 262 and floral recovery likely did not follow a congruent globally universal trend but instead varied by region according to local environmental conditions 150 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 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 262 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 263 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 264 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 150 Plant survivors of the PTME in South China experienced extremely high rates of mutagenesis induced by heavy metal poisoning 265 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 264 In Tibet the early Dienerian Endosporites papillatus Pinuspollenites 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 266 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 267 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 144 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 43 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 268 Approximately 95 of these peat producing plants went locally extinct at the P Tr boundary 269 coal deposits in Australia and Antarctica disappear significantly before the P Tr boundary 42 Terrestrial vertebrates edit 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 13 A tetrapod gap may have existed from the Induan until the early Spathian between 30 N and 40 S due to extreme heat making these low latitudes uninhabitable for these animals During the hottest phases of this interval the gap would have spanned an even greater latitudinal range 270 East central Pangaea with its relatively wet climate served as a dispersal corridor for PTME survivors during their Early Triassic recolonisation of the supercontinent 271 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 272 273 In Russia even after 15 Myr of recovery during which ecosystems were rebuilt and remodelled many terrestrial vertebrate guilds were absent including small insectivores small piscivores large herbivores and apex predators 274 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 275 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 276 Synapsids edit nbsp 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 although some recent evidence has called into question its status as a post PTME disaster taxon 277 The evolutionary success of Lystrosaurus in the aftermath of the PTME is believed to be attributable to the dicynodont taxon s grouping behaviour and tolerance for extreme and highly variable climatic conditions 278 Other likely factors behind the success of Lystrosaurus included extremely fast growth rate exhibited by the dicynodont genus 279 along with its early onset of sexual maturity 280 Antarctica may have served as a refuge for dicynodonts during the PTME from which surviving dicynodonts spread out of in its aftermath 281 Ichnological evidence from the earliest Triassic of the Karoo Basin shows dicynodonts were abundant in the immediate aftermath of the biotic crisis 282 Smaller carnivorous cynodont therapsids also survived a group that included the ancestors of mammals 283 As with dicynodonts selective pressures favoured endothermic epicynodonts 284 Therocephalians likewise survived burrowing may have been a key adaptation that helped them make it through the PTME 285 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 286 Early Triassic therocephalians were mostly survivors of the PTME rather than newly evolved taxa that originated during the evolutionary radiation in its aftermath 287 Both therocephalians and cynodonts known collectively as eutheriodonts decreased in body size from the Late Permian to the Early Triassic 283 This decrease in body size has been interpreted as an example of the Lilliput effect 288 Sauropsids edit Archosaurs which included the ancestors of dinosaurs and crocodilians were initially rarer than therapsids but they began to displace therapsids in the mid Triassic Olenekian tooth fossil assemblages from the Karoo Basin indicate that archosauromorphs were already highly diverse by this point in time though not very ecologically specialised 289 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 290 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 291 while losing part of the differential color sensitive retinal receptors reptilians and birds preserved Archosaurs also experienced an increase in metabolic rates over time during the Early Triassic 292 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 Temnospondyls edit 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 293 During the Induan tupilakosaurids in particular thrived as disaster taxa 294 including Tupilakosaurus itself 295 though they gave way to other temnospondyls as ecosystems recovered 294 Temnospondyls were reduced in size during the Induan but their body size rebounded to pre PTME levels during the Olenekian 296 Mastodonsaurus and trematosaurians were the main aquatic and semiaquatic predators during most of the Triassic some preying on tetrapods and others on fish 297 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 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 136 Microbially induced sedimentary structures MISS dominated North Chinese terrestrial fossil assemblages in the Early Triassic 298 299 In Arctic Canada as well MISS became a common occurrence following the Permian Triassic extinction 300 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 299 Hypotheses about cause editExplaining an event from 250 million years ago is inherently difficult with much of the evidence on land eroded or deeply buried while the spreading seafloor is completely recycled over 200 million years 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 but with much less current consensus 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 93 Volcanism edit Siberian Traps edit The flood basalt eruptions that produced the large igneous province of the Siberian Traps were among the largest known volcanic events extruding lava over 2 000 000 square kilometres 770 000 sq mi roughly the size of Saudi Arabia producing a catastrophic impact 301 302 303 304 305 The date of the Siberian Traps eruptions matches well with the extinction event 47 306 307 19 308 A study of the Norilsk and Maymecha Kotuy regions of the northern Siberian platform indicates that volcanic activity occurred during a few enormous pulses of magma as opposed to more regular flows 309 The Siberian Traps caused one of the most rapid rises of atmospheric carbon dioxide levels in the geologic record 310 with the rate of carbon dioxide emissions estimated as five times faster than during the preceeding catastrophic Capitanian mass extinction 311 during the eruption of the Emeishan Traps 312 313 314 Overwhelming inorganic carbon sinks carbon dioxide levels might have jumped from between 500 and 4 000 ppm prior to the extinction to around 8 000 ppm after according to one estimate 21 Another study estimated pre extinction carbon dioxide levels at 400 ppm which then rose to 2 500 ppm with 3 900 to 12 000 gigatonnes of carbon added to the ocean atmosphere system 22 Extreme temperature rise would have followed 315 though some evidence suggests a lag of 12 000 to 128 000 years between the rise in volcanic carbon dioxide emissions and global warming 316 During the latest Permian before the extinction global average surface temperatures were about 18 2 C 317 which shot up to as much as 35 C this hyperthermal condition lasting 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 surged about 8 C 24 In present day Iran tropical sea surface temperatures were between 27 and 33 C during the Changhsingian but jumped to over 35 C during the PTME 318 These extremely high atmospheric carbon dioxide concentrations persisted over a long period 319 The position and alignment of Pangaea at the time made the inorganic carbon cycle very inefficient at burying carbon 320 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 321 9 Evidence also points to volcanic combustion of undergroud fossil fuel deposits based on paired coronene mercury spikes 322 30 coinciding with geographically widespread mercury anomalies and the rise in isotopically light carbon 323 Te Th values increase twentyfold over the PTME further indicating it was concomitant with extreme volcanism 324 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 325 The Siberian Traps eruptions had unusual features that made them even more dangerous The Siberian lithosphere is rich in halogens extremely destructive to the ozone layer and evidence from subcontinental lithospheric xenoliths indicates that as much as 70 of their halogen content was released into the atmosphere 326 Around 18 teratonnes of hydrochloric acid were emitted 327 along with sulphur rich volatiles that caused dust clouds and acid aerosols which would have blocked out sunlight and disrupted photosynthesis 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 punctuating the broader trend of rapid global warming 328 with glacio eustatic sea level fall 326 329 The eruptions may also have caused acid rain as the aerosols washed out of the atmosphere 330 That may have killed land plants and mollusks and planktonic organisms with 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 ash thrown high into the atmosphere increasing the short term cooling effect 331 When this had washed out of the atmosphere the excess carbon dioxide would have remained and global warming would have proceeded unchecked 315 Burning of hydrocarbon deposits may have exacerbated the extinction 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 may emit large amounts of greenhouse and toxic gases 332 The unique setting of the Siberian Traps over these deposits is likely the reason for the severity of the extinction 333 334 335 The basalt lava erupted or intruded into carbonate rocks and sediments in the process of forming large coal beds which would have emitted large amounts of carbon dioxide leading to stronger global warming after the dust and aerosols settled 315 The change of the eruptions from flood basalt to sill dominated emplacement liberating even more trapped hydrocarbon deposits coincides with the main onset of the extinction 27 and is linked to a major negative d 13C excursion 336 Venting of coal derived methane was accompanied by explosive combustion of coal and discharge of coal fly ash 28 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 They found ash deposits in deep rock layers near what is now the Buchanan Lake Formation 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 337 338 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 339 However some researchers propose that these supposed fly ashes were actually the result of wildfires not related to massive coal combustion by intrusive volcanism 340 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 341 The sill dominated emplacement of the Siberian Traps prolonged their warming effects 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 342 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 206 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 343 Mercury anomalies corresponding to the time of Siberian Traps activity have been found in many geographically disparate sites 344 indicating that these volcanic eruptions released significant quantities of toxic mercury into the atmosphere and ocean causing even larger terrestrial and marine die offs 345 346 347 A series of surges raised terrestrial and marine environmental mercury concentrations by orders of magnitude above normal background levels and caused mercury poisoning over periods of a thousand years each 348 Mutagenesis in surviving plants after the PTME coeval with mercury and copper loading confirms volcanically induced heavy metal toxicity 265 Increased bioproductivity may have sequestered mercury and party mitigated poisoning 349 Immense volumes of nickel aerosols and cobalt and arsenic emisions were also released 350 351 345 further contributing to metal poisoning 352 The devastation wrought by the Siberian Traps did not end following the Permian Triassic boundary Carbon isotope fluctuations suggest that massive Siberian Traps activity recurred many times during the Early Triassic 353 causing further extinction events during the epoch 354 Choiyoi Silicic Large Igneous Province edit A second flood basalt event that produced the Choiyoi Silicic Large Igneous Province in southwestern Gondwana between around 286 Ma and 247 Ma has also been suggested as a significant additional extinction mechanism 144 At 1 300 000 cubic kilometres in volume 355 and 1 680 000 square kilometres in area this event was 40 the size of the Siberian Traps 144 Specifically this flood basalt has been implicated in the regional demise of the Gondwanan Glossopteris flora 356 Indochina South China subduction zone volcanic arc edit 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 a subduction zone with a volcanic arc Hafnium isotopes from syndepositional magmatic zircons found in ash beds created by this volcanic pulse confirm its origin in subduction zone volcanism rather than large igneous province activity 357 The enrichment of copper samples from these deposits in isotopically light copper provide additional confirmation 358 This volcanism has been speculated to have caused local biotic stress among radiolarians sponges and brachiopods over the 60 000 years preceding the end Permian marine extinction as well as an ammonoid crisis with decreased morphological complexity and size and increased rate of turnover beginning in the lower C yini biozone around 200 000 years before the extinction 357 Methane clathrate gasification edit Main article Clathrate gun hypothesis Further information Arctic methane emissionsMethane clathrates also known as methane hydrates consist of molecules of methane trapped in the crystal lattice of ice This methane produced by methanogen microbes has a 13C 12C isotope ratio about 6 below normal d 13C 6 0 At the right combination of pressure and temperature clathrates form near the surface of permafrost and in large quantities on continental shelves and nearby seabed at water depths of at least 300 m 980 ft buried in sediments up to 2 000 m 6 600 ft below the sea floor 359 Massive release of methane from these clathrates may have contributed to the PTME as scientists have found worldwide evidence of a swift decrease of about 1 in the13C 12C ratio in carbonate rocks from the end Permian 75 360 This is the first largest and fastest of a series of excursions decreases and increases in the ratio until it abruptly stabilised in the middle Triassic followed soon afterwards by the recovery of calcifying shelled sealife 103 The seabed probably contained methane hydrate deposits and the lava caused the deposits to dissociate releasing vast quantities of methane 361 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 362 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 363 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 sulfidic events at the time with the exact mechanism compared to the 1986 Lake Nyos disaster 364 The clathrate hypothesis seemed the only proposed mechanism sufficient to cause a global 1 reduction in the 13C 12C ratio 365 44 While a variety of factors may have contributed to the ratio drop a 2002 review found most of them insufficient to account for the observed amount 29 Gases from volcanic eruptions have a13C 12C ratio about 0 5 to 0 8 below standard d 13C 0 5 to 0 8 but a 1995 assessment concluded that the observed 1 0 worldwide reduction would have required eruptions massively larger than any found 366 However this analysis addressed only CO2 produced by the magma itself not from interactions with carbon bearing sediments as described below 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 the13C 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 366 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 gigatonnes 1 gigatonne 1012 kg 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 44 Moreover sea levels were rising rather than falling at the time of the extinction 315 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 103 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 367 That or another organic based reason may have been responsible for both that and a late Proterozoic Cambrian pattern of fluctuating 13C 12C ratios 103 However the clathrate hypothesis has also been criticized Carbon cycle models which include consideration of roasting carbonate sediments by volcanism confirm that it would have had enough effect to produce the observed reduction 29 368 Also 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 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 103 The latest research suggests that greenhouse gas release during the extinction event was dominated by volcanic carbon dioxide 369 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 22 Adding to the evidence against methane clathrate release as the central driver of warming the main rapid warming event is also associated with marine transgression rather than regression the former would not normally have initiated methane release which would have instead required a decrease in pressure something that would be generated by a retreat of shallow seas 370 The configuration of the world s landmasses into one supercontinent would also mean that the global gas hydrate reservoir was lower than today further damaging the case for methane clathrate dissolution as a major cause of the carbon cycle disruption 371 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 160 Higher concentrations of CO2 also result in decreased activity levels in many active marine animals hindering their ability to obtain food 372 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 160 In addition an increase in CO2 concentration is inevitably linked to ocean acidification 373 consistent with the preferential extinction of heavily calcified taxa and other signals in the rock record that suggest a more acidic ocean 26 The decrease in ocean pH is calculated to be up to 0 7 units 25 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 3 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 374 One study published in Scientific Reports concluded that widespread ocean acidification if it did occur was not intense enough to impede calcification and only occurred during the beginning of the extinction event 375 The relative success of many marine organisms that were very vulnerable to acidification has further been used to argue that acidification was not a major extinction contributor 376 The persistence of highly elevated carbon dioxide concentrations in the atmosphere during the Early Triassic would have impeded the recovery of biocalcifying organisms after the PTME 377 Acidity generated by increased carbon dioxide concentrations in soil and sulphur dioxide dissolution in rainwater was also a kill mechanism on land 378 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 379 Further evidence of an increase in soil acidity comes from elevated Ba Sr ratios in earliest Triassic soils 380 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 381 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 382 Hopane anomalies have also been interpreted as evidence of acidic soils and peats 383 As with many other environmental stressors acidity on land episodically persisted well into the Triassic stunting the recovery of terrestrial ecosystems 384 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 385 386 387 Throughout most of the Tethys and Panthalassic Oceans evidence for anoxia appears at the extinction event including small pyrite framboids 388 negative d238U excursions 389 390 negative d15N excursions 391 positive d82 78Se isotope excursions 392 relatively positive d13C ratios in polycyclic aromatic hydrocarbons 393 high Th U ratios 389 positive Ce Ce anomalies 394 and fine laminations in sediments 388 However evidence for anoxia precedes the extinction at some other sites including Spiti India 395 Shangsi China 396 Meishan China 397 Opal Creek Alberta 398 and Kap Stosch Greenland 399 Biogeochemical evidence also points to the presence of euxinia during the PTME 400 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 401 Negative mercury isotope excursions further bolster evidence for extensive euxinia during the PTME 402 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 403 Evidence of a lag between volcanic mercury inputs and biotic turnovers provides further support for anoxia and euxinia as the key killing mechanism because extinctions would be expected to be synchronous with volcanic mercury discharge if volcanism and hypercapnia was the primary driver of extinction 404 The sequence of events leading to anoxic oceans may have been triggered by carbon dioxide emissions from the eruption of the Siberian Traps 405 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 coastal evaporation would have caused the formation of warm saline bottom water WSBW depleted in oxygen and nutrients which spread across the world through the deep oceans The influx of WSBW caused thermal expansion of water that raised sea levels bringing anoxic waters onto shallow shelfs and enhancing the formation of WSBW in a positive feedback loop 406 The flux of terrigeneous material into the oceans increased as a result of soil erosion which would have facilitated increased eutrophication 407 Increased chemical weathering of the continents due to warming and the acceleration of the water cycle would increase the riverine flux of nutrients to the ocean 408 Additionally the Siberian Traps directly fertilised the oceans with iron and phosphorus as well triggering bioblooms and marine snowstorms Increased phosphate levels would have supported greater primary productivity in the surface oceans 409 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 410 Along the Panthalassan coast of South China oxygen decline was also driven by large scale upwelling of deep water enriched in various nutrients causing this region of the ocean to be hit by especially severe anoxia 411 Convective overturn helped facilitate the expansion of anoxia throughout the water column 412 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 405 In some regions anoxia briefly disappeared when transient cold snaps resulting from volcanic sulphur emissions occurred 413 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 414 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 415 416 417 particularly that of benthic organisms 418 208 Anoxia disappeared from shallow waters more rapidly than the deep ocean 419 Reexpansions of oxygen minimum zones did not cease until the late Spathian periodically setting back and restarting the biotic recovery process 420 The decline in continental weathering towards the end of the Spathian at last began ameliorating marine life from recurrent anoxia 421 In some regions of Panthalassa pelagic zone anoxia continued to recur as late as the Anisian 422 probably due to increased productivity and a return of aeolian upwelling 423 Some sections show a rather quick return to oxic water column conditions however so for how long anoxia persisted remains debated 424 The volatility of the Early Triassic sulphur cycle suggests marine life continued to face returns of euxinia as well 425 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 426 Oxygen depletion via extensive microbial blooms also played a role in the biological collapse of not just marine ecosystems but freshwater ones as well Persistent lack of oxygen after the extinction event itself helped delay biotic recovery for much of the Early Triassic epoch 427 Aridification edit Increasing continental aridity a trend well underway even before the PTME as a result of the coalescence of the supercontinent Pangaea was drastically exacerbated by terminal Permian volcanism and global warming 428 The combination of global warming and drying generated an increased incidence of wildfires 429 Tropical coastal swamp floras such as those in South China may have been very detrimentally impacted by the increase in wildfires 152 though it is ultimately unclear if an increase in wildfires played a role in driving taxa to extinction 430 Aridification trends varied widely in their tempo and regional impact Analysis of the fossil river deposits of the floodplains of the Karoo Basin indicate a shift from meandering to braided river patterns indicating a very abrupt drying of the climate 431 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 432 A pattern of aridity induced extinctions that progressively ascended up the food chain has been deduced from Karoo Basin biostratigraphy 155 Evidence for aridification in the Karoo across the Permian Triassic boundary is not however universal as some palaeosol evidence indicates a wettening of the local climate during the transition between the two geologic periods 433 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 434 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 435 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 Drying of this basin 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 146 The persistence of hyperaridity varied regionally as well 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 273 272 Elsewhere such as in the Karoo Basin episodes of dry climate recurred regularly in the Early Triassic with profound effects on terrestrial tetrapods 428 The types and diversity of ichnofossils in a locality has been used as an indicator measuring aridity Nurra an ichnofossil site on the island of Sardinia shows evidence of major drought related stress among crustaceans Whereas the Permian subnetwork at Nurra displays extensive horizontal backfilled traces and high ichnodiversity the Early Triassic subnetwork is characterised by an absence of backfilled traces an ichnological sign of aridification 436 Ozone depletion edit A collapse of the atmospheric ozone shield has been invoked as an explanation for the mass extinction 437 particularly that of terrestrial plants 37 Ozone production may have been reduced by 60 70 increasing the flux of ultraviolet radiation by 400 at equatorial latitudes and 5 000 at polar latitudes 438 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 439 38 Malformed plant spores from the time of the extinction event show an increase in ultraviolet B absorbing compounds confirming that increased ultraviolet radiation played a role in the environmental catastrophe and excluding other possible causes of mutagenesis such as heavy metal toxicity in these mutated spores 36 Extremely positive D33S anomalies provide evidence of photolysis of volcanic SO2 indicating increased ultraviolet radiation flux 440 Multiple mechanisms could have reduced the ozone shield and rendered it ineffective Computer modelling shows high atmospheric methane levels are associated with ozone shield decline and may have contributed to its reduction during the PTME 441 Volcanic emissions of sulphate aerosols into the stratosphere would have dealt significant destruction to the ozone layer 439 As mentioned previously the rocks in the region where the Siberian Traps were emplaced are extremely rich in halogens 326 The intrusion of Siberian Traps volcanism into deposits rich in organohalogens such as methyl bromide and methyl chloride would have been another source of ozone destruction 38 An uptick in wildfires a natural source of methyl chloride would have had further deleterious effects still on the atmospheric ozone shield 442 Upwelling of euxinic water may have released massive hydrogen sulphide 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 443 Indeed biomarker evidence for anaerobic photosynthesis by Chlorobiaceae green sulfur bacteria from the Late Permian into the Early Triassic indicates that hydrogen sulphide did upwell into shallow waters because these bacteria are restricted to the photic zone and use sulfide as an electron donor 444 Asteroid impact edit nbsp 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 445 However suggestions that an asteroid impact was the trigger of the Permian Triassic extinction are now largely rejected 446 Reported evidence for an impact event from the P Tr boundary level includes rare grains of shocked quartz in Australia and Antarctica 447 448 fullerenes trapping extraterrestrial noble gases 449 meteorite fragments in Antarctica 450 and grains rich in iron nickel and silicon which may have been created by an impact 451 However the accuracy of most of these claims has been challenged 452 453 454 455 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 456 Iridium levels in many sites straddling the Permian Triassic boundaries are not anomalous providing evidence against an extraterrestrial impact as the PTME s cause 457 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 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 458 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 315 such as the Siberian Traps eruptions at either an impact site 459 or the antipode of an impact site 315 460 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 461 Additionally many sites spanning the Permian Triassic boundary display a complete lack of evidence of an impact event 462 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 448 and the hypothesized 480 km 300 mi Wilkes Land crater of East Antarctica 463 464 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 33 precipitated the mass extinction 465 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 34 The large earthquakes generated by the impact would have additionally generated massive tsunamis across much of the globe 465 35 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 34 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 466 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 467 Methanogens edit A hypothesis published in 2014 posits that a genus of anaerobic methanogenic archaea known as Methanosarcina was responsible for the event 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 marine sediment The result would have been a sharp buildup of methane and carbon dioxide in the 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 31 Chemostratigraphic analysis of Permian Triassic boundary sediments in Chaotian demonstrates a methanogenic burst could be responsible for some percentage of the carbon isotopic fluctuations 32 On the other hand in the canonical Meishan sections the nickel concentration increases somewhat after the d 13C concentrations have begun to fall 468 Supercontinent Pangaea edit nbsp 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 469 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 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 470 Comparison to present global warming editThe PTME has been compared to the current anthropogenic global warming situation and Holocene extinction due to sharing the common characteristic of rapid rates of carbon dioxide release Though the current rate of greenhouse gas emissions is more than an order of magnitude greater than the rate measured over the course of the PTME the discharge of greenhouse gases during the PTME is poorly constrained geo chronologically and was most likely pulsed and constrained to a few key short intervals rather than continuously occurring at a constant rate for the whole extinction interval the rate of carbon release within these intervals was likely to have been similar in timing to modern anthropogenic emissions 310 As they did during the PTME oceans in the present day are experiencing drops in pH and in oxygen levels prompting further comparisons between modern anthropogenic ecological conditions and the PTME 471 Another biocalcification event similar in its effects on modern marine ecosystems is predicted to occur if carbon dioxide levels continue to rise 377 The similarities between the two extinction events have led to warnings from geologists about the urgent need for reducing carbon dioxide emissions if an event similar to the PTME is to be prevented from occurring 310 See also edit nbsp Evolutionary biology portal nbsp Paleontology portal Carbon dioxide Extinction event Climate change List of possible impact structures on Earth Silurian hypothesisReferences edit a b Rohde RA Muller RA 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 a b c 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 2024 03 26 Jouault Corentin Nel Andre Perrichot Vincent Legendre Frederic Condamine Fabien L 6 December 2011 Multiple drivers and lineage specific insect extinctions during the Permo Triassic Nature Communications 13 1 7512 doi 10 1038 s41467 022 35284 4 PMC 9726944 PMID 36473862 a b Delfini Massimo Kustatscher Evelyn Lavezzi Fabrizio Bernardi Massimo 29 July 2021 The End Permian Mass Extinction Nature s Revolution In Martinetto Edoardo Tschopp Emanuel Gastaldo Robert A eds Nature through Time Springer Textbooks in Earth Sciences Geography and Environment Springer Cham pp 253 267 doi 10 1007 978 3 030 35058 1 10 ISBN 978 3 030 35060 4 S2CID 226405085 Great Dying lasted 200 000 years National Geographic 23 November 2011 Archived from the original on November 24 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 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 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 Evolution 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 18 October 2016 Estimates of the magnitudes of major marine mass extinctions in earth history Proceedings of the National Academy of Sciences of the United States of America 113 42 E6325 E6334 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 a b 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 Marshall Charles R 5 January 2023 Forty years later The status of the Big Five mass extinctions Cambridge Prisms Extinction 1 1 13 doi 10 1017 ext 2022 4 S2CID 255710815 a b c d Jin Y G Wang Y Wang W Shang Q H Cao C Q Erwin D H 21 July 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 Retrieved 5 March 2023 a b Burgess Seth D Bowring Samuel A 1 August 2015 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 Hulse D Lau K V Sebastiaan J V Arndt S Meyer K M Ridgwell A 28 Oct 2021 End Permian marine extinction due to temperature driven nutrient recycling and euxinia Nat Geosci 14 11 862 867 Bibcode 2021NatGe 14 862H doi 10 1038 s41561 021 00829 7 S2CID 240076553 a b 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 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 Retrieved 2024 03 26 a b Frank T D Fielding Christopher R Winguth A M E Savatic K Tevyaw A Winguth C McLoughlin Stephen Vajda Vivi Mays C Nicoll R Bocking M Crowley J L 19 May 2021 Pace magnitude and nature of terrestrial climate change through the end Permian extinction in southeastern Gondwana Geology 49 9 1089 1095 Bibcode 2021Geo 49 1089F doi 10 1130 G48795 1 S2CID 236381390 Retrieved 2024 03 26 a b Joachimski Michael M Lai Xulong Shen Shuzhong Jiang Haishui Luo Genming Chen Bo Chen Jun Sun Yadong 1 March 2012 Climate warming in the latest Permian and the Permian Triassic mass extinction Geology 40 3 195 198 Bibcode 2012Geo 40 195J doi 10 1130 G32707 1 Retrieved 2024 03 26 a b 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 a b 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 a b 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 a b 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 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 of the United States of America 99 7 4172 4177 Bibcode 2002PNAS 99 4172B doi 10 1073 pnas 032095199 PMC 123621 PMID 11917102 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 nbsp Available under CC BY 4 0 a b Rothman D H Fournier G P French K L Alm E J Boyle E A Cao C Summons R E 31 March 2014 Methanogenic burst in the end Permian carbon cycle Proceedings of the National Academy of Sciences of the United States of America 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 a b Saitoh Masafumi Isozaki Yukio 5 February 2021 Carbon Isotope Chemostratigraphy Across the Permian Triassic Boundary at Chaotian China Implications for the Global Methane Cycle in the Aftermath of the Extinction Frontiers in Earth Science 8 665 Bibcode 2021FrEaS 8 665S doi 10 3389 feart 2020 596178 a b Tohver Eric Lana Cris Cawood P A Fletcher I R Jourdan F Sherlock S et al 1 June 2012 Geochronological constraints on the age of a Permo Triassic impact event U Pb and 40Ar 39Ar results for the 40 km Araguainha structure of central Brazil Geochimica et Cosmochimica Acta 86 214 227 Bibcode 2012GeCoA 86 214T doi 10 1016 j gca 2012 03 005 a b c Tohver Eric Cawood P A Riccomini Claudio Lana Cris Trindade R I F 1 October 2013 Shaking a methane fizz Seismicity from the Araguainha impact event and the Permian Triassic global carbon isotope record Palaeogeography Palaeoclimatology Palaeoecology 387 66 75 Bibcode 2013PPP 387 66T doi 10 1016 j palaeo 2013 07 010 Retrieved 2024 03 26 a b Tohver Eric Schmieder Martin Lana Cris Mendes Pedro S T Jourdan Fred Warren Lucas Riccomini Claudio 2 January 2018 End Permian impactogenic earthquake and tsunami deposits in the intracratonic Parana Basin of Brazil Geological Society of America Bulletin 130 7 8 1099 1120 Bibcode 2018GSAB 130 1099T doi 10 1130 B31626 1 Retrieved 2024 03 26 a b Liu Feng Peng Huiping Marshall John E A Lomax Barry H Bomfleur Benjamin Kent Matthew S Fraser Wesley T Jardine Phillip E 6 January 2023 Dying in the Sun Direct evidence for elevated UV B radiation at the end Permian mass extinction Science Advances 9 1 eabo6102 Bibcode 2023SciA 9O6102L doi 10 1126 sciadv abo6102 PMC 9821938 PMID 36608140 a b Benca Jeffrey P Duijnstee Ivo A P Looy Cindy V 7 February 2018 UV B induced forest sterility Implications of ozone shield failure in Earth s largest extinction Science Advances 4 2 e1700618 Bibcode 2018SciA 4 618B doi 10 1126 sciadv 1700618 PMC 5810612 PMID 29441357 a b c d Visscher Henk Looy Cindy V Collinson Margaret E Brinkhuis Henk Cittert Johanna H A van Konijnenburg Kurschner Wolfram M Sephton Mark A 31 August 2004 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 Dai X Davies J H F L Yuan Z Brayard A Ovtcharova M Xu G Liu X Smith C P A Schweitzer C E Li M Perrot M G Jiang S Miao L Cao Y Yan J Bai R Wang F Guo W Song H Tian L Dal Corso J Liu Y Chu D Song H 2023 A Mesozoic fossil lagerstatte from 250 8 million years ago shows a modern type marine ecosystem Science 379 6632 567 572 doi 10 1126 science adf1622 PMID 36758082 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 2017 Unexpected Early Triassic marine ecosystem and the rise of the Modern evolutionary fauna Science Advances 3 2 e1602159 Bibcode 2017SciA 3E2159B doi 10 1126 sciadv 1602159 PMC 5310825 PMID 28246643 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 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 Erwin D H 1993 The great Paleozoic crisis Life and death in the Permian Columbia University Press ISBN 978 0 231 07467 4 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 World and Regional Geology Cambridge Cambridge University Press pp 1 7 doi 10 1017 CBO9780511529498 002 ISBN 978 0 521 54573 0 Retrieved 3 July 2023 a b Burgess Seth D Bowring Samuel Shen Shu zhong 10 February 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 Yuan Dong xun Shen Shu zhong Henderson Charles M Chen Jun Zhang Hua Feng Hong zhen 1 September 2014 Revised conodont based integrated high resolution timescale for the Changhsingian Stage and end Permian extinction interval at the Meishan sections South China Lithos 204 220 245 Bibcode 2014Litho 204 220Y doi 10 1016 j lithos 2014 03 026 Retrieved 10 August 2023 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 Musashi Masaaki Isozaki Yukio Koike Toshio Kreulen Rob 30 August 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 Retrieved 3 July 2023 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 Korte Christoph Kozur Heinz W 9 September 2010 Carbon isotope stratigraphy across the Permian Triassic boundary A review Journal of Asian Earth Sciences 39 4 215 235 Bibcode 2010JAESc 39 215K doi 10 1016 j jseaes 2010 01 005 Retrieved 26 June 2023 Li Rong Jones Brian 15 February 2017 Diagenetic overprint on negative d13C excursions across the Permian Triassic boundary A case study from Meishan section China Palaeogeography Palaeoclimatology Palaeoecology 468 18 33 Bibcode 2017PPP 468 18L doi 10 1016 j palaeo 2016 11 044 ISSN 0031 0182 Retrieved 20 September 2023 Schobben Martin Heuer Franziska Tietje Melanie Ghaderi Abbas Korn Dieter Korte Christoph Wignall Paul B 19 November 2018 Chemostratigraphy Across the Permian Triassic Boundary The Effect of Sampling Strategies on Carbonate Carbon Isotope Stratigraphic Markers In Sial Alcides N Gaucher Claudio Ramkumar Muthuvairavasamy Pinto Ferreira Valderez eds Chemostratigraphy Across Major Chronological Boundaries American Geophysical Union pp 159 181 doi 10 1002 9781119382508 ch9 ISBN 9781119382508 S2CID 134060610 Retrieved 20 September 2023 Daily CO2 Mauna Loa Observatory 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 of the United States of America 93 5 2155 2158 Bibcode 1996PNAS 93 2155V doi 10 1073 pnas 93 5 2155 PMC 39926 PMID 11607638 Tewari Rajni Awatar Ram Pandita Sundeep K McLoughlin Stephen D Agnihotri Deepa Pillai Suresh S K et al October 2015 The Permian Triassic palynological transition in the Guryul Ravine section Kashmir India implications for Tethyan Gondwanan correlations Earth Science Reviews 149 53 66 Bibcode 2015ESRv 149 53T doi 10 1016 j earscirev 2014 08 018 Retrieved 26 May 2023 Foster CB Stephenson MH Marshall C Logan GA Greenwood PF 2002 A revision of Reduviasporonites Wilson 1962 Description illustration comparison and biological affinities Palynology 26 1 35 58 Bibcode 2002Paly 26 35F 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 Bibcode 2010RPaPa 162 522D doi 10 1016 j revpalbo 2010 09 003 Retrieved 8 January 2023 Lopez Gomez J Taylor EL 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 of the United States of America 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 GJ Smith RMH Ward PD 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 MA Visscher H Looy CV Verchovsky AB Watson JS 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 Bibcode 2018GeoFr 9 147R doi 10 1016 j gsf 2017 06 005 Retrieved 24 December 2022 Yin Hongfu Jiang Haishui Xia Wenchen Feng Qinglai Zhang Ning Shen Jun October 2014 The end Permian regression in South China and its implication on mass extinction Earth Science Reviews 137 19 33 Bibcode 2014ESRv 137 19Y doi 10 1016 j earscirev 2013 06 003 Retrieved 20 September 2023 de Wit Maarten J Ghosh Joy G de Villiers Stephanie Rakotosolofo Nicolas Alexander James Tripathi Archana Looy Cindy March 2002 Multiple Organic Carbon Isotope Reversals across the Permo Triassic Boundary of Terrestrial Gondwana Sequences Clues to Extinction Patterns and Delayed Ecosystem Recovery The Journal of Geology 110 2 227 240 Bibcode 2002JG 110 227D doi 10 1086 338411 ISSN 0022 1376 S2CID 129653925 Retrieved 20 September 2023 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 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 Li Guoshan Liao Wei Li Sheng Wang Yongbiao Lai Zhongping 23 March 2021 Different triggers for the two pulses of mass extinction across the Permian and Triassic boundary Scientific Reports 11 1 6686 Bibcode 2021NatSR 11 6686L doi 10 1038 s41598 021 86111 7 PMC 7988102 PMID 33758284 Shen Jiaheng Zhang Yi Ge Yang Huan Xie Shucheng Pearson Ann 3 October 2022 Early and late phases of the Permian Triassic mass extinction marked by different atmospheric CO2 regimes Nature Geoscience 15 1 839 844 Bibcode 2022NatGe 15 839S doi 10 1038 s41561 022 01034 w S2CID 252697822 Retrieved 20 April 2023 Wang SC Everson PJ 2007 Confidence intervals for pulsed mass extinction events Paleobiology 33 2 324 336 Bibcode 2007Pbio 33 324W doi 10 1666 06056 1 S2CID 2729020 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 59410936F doi 10 1016 j palaeo 2022 110936 S2CID 247514266 Huang Yuangeng Chen Zhong Qiang Roopnarine Peter D Benton Michael James Zhao Laishi Feng Xueqian Li Zhenhua 24 February 2023 The stability and collapse of marine ecosystems during the Permian Triassic mass extinction Current Biology 33 6 1059 1070 e4 doi 10 1016 j cub 2023 02 007 PMID 36841237 S2CID 257186215 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 Shen Shu Zhong Crowley James L Wang Yue Bowring Samuel R Erwin Douglas H Sadler Peter M et al 17 November 2011 Calibrating the End Permian Mass Extinction Science 334 6061 1367 1372 Bibcode 2011Sci 334 1367S doi 10 1126 science 1213454 PMID 22096103 S2CID 970244 Retrieved 26 May 2023 Metcalfe I Crowley J L Nicoll Robert S Schmitz M August 2015 High precision U Pb CA TIMS calibration of Middle Permian to Lower Triassic sequences mass extinction and extreme climate change in eastern Australian Gondwana Gondwana Research 28 1 61 81 Bibcode 2015GondR 28 61M doi 10 1016 j gr 2014 09 002 Retrieved 31 May 2023 Dal Corso Jacopo Song Haijun Callegaro Sara Chu Daoliang Sun Yadong Hilton Jason et al 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 hdl 10852 100010 S2CID 247013868 Retrieved 20 December 2022 Chen Zhong Qiang Harper David A T Grasby Stephen Zhang Lei 1 August 2022 Catastrophic event sequences across the Permian Triassic boundary in the ocean and on land Global and Planetary Change 215 103890 Bibcode 2022GPC 21503890C doi 10 1016 j gloplacha 2022 103890 ISSN 0921 8181 S2CID 250417358 Retrieved 24 November 2023 Hermann Elke Hochuli Peter A Bucher Hugo Vigran Jorunn O Weissert Helmut Bernasconi Stefano M December 2010 A close up view of the Permian Triassic boundary based on expanded organic carbon isotope records from Norway Trondelag and Finnmark Platform Global and Planetary Change 74 3 4 156 167 Bibcode 2010GPC 74 156H doi 10 1016 j gloplacha 2010 10 007 Retrieved 10 August 2023 Gastaldo Robert A Kamo Sandra L Neveling Johann Geissman John W Bamford Marion Looy Cindy V 1 October 2015 Is the vertebrate defined Permian Triassic boundary in the Karoo Basin South Africa the terrestrial expression of the end Permian marine event Geology 43 10 939 942 Bibcode 2015Geo 43 939G doi 10 1130 G37040 1 S2CID 129258297 Zhang Peixin Yang Minfang Lu Jing Bond David P G Zhou Kai Xu Xiaotao et al March 2023 End Permian terrestrial ecosystem collapse in North China Evidence from palynology and geochemistry Global and Planetary Change 222 104070 Bibcode 2023GPC 22204070Z doi 10 1016 j gloplacha 2023 104070 S2CID 256920630 Wu Qiong Zhang Hua Ramezani Jahandar Zhang Fei fei Erwin Douglas H Feng Zhuo Shao Long yi Cai Yao feng Zhang Shu han Xu Yi gang Shen Shu zhong 2 February 2024 The terrestrial end Permian mass extinction in the paleotropics postdates the marine extinction Science Advances 10 5 eadi7284 doi 10 1126 sciadv adi7284 ISSN 2375 2548 PMC 10830061 PMID 38295161 Bond DPG Wignall PB Wang W Izon G Jiang HS Lai XL 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 a b Retallack GJ Metzger CA Greaver T Jahren AH Smith RMH Sheldon ND 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 Li Dirson Jian 18 December 2012 Tectonic cause of mass extinctions and the genomic contribution to biodiversification arXiv 1212 4229 q bio PE 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 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 Shi GR 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 Sugiyama T December 2000 Diversity and extinction patterns of Permian coral faunas of China Lethaia 33 4 285 294 Bibcode 2000Letha 33 285W doi 10 1080 002411600750053853 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 Song Haijun Wu Yuyang Dai Xu Dal Corso Jacopo Wang Fengyu Feng Yan Chu Daoliang Tian Li Song Huyue Foster William J 6 May 2024 Respiratory protein driven selectivity during the Permian Triassic mass extinction The Innovation 5 3 100618 doi 10 1016 j xinn 2024 100618 Retrieved 21 May 2024 Yan Jia Song Haijun Dai Xu 1 February 2023 Increased bivalve cosmopolitanism during the mid Phanerozoic mass extinctions Palaeogeography Palaeoclimatology Palaeoecology 611 111362 Bibcode 2023PPP 61111362Y doi 10 1016 j palaeo 2022 111362 Retrieved 20 February 2023 Allen Bethany J Clapham Matthew E Saupe Erin E Wignall Paul B Hill Daniel J Dunhill Alexander M 10 February 2023 Estimating spatial variation in origination and extinction in deep time a case study using the Permian Triassic marine invertebrate fossil record Paleobiology 49 3 509 526 Bibcode 2023Pbio 49 509A doi 10 1017 pab 2023 1 hdl 20 500 11850 598054 S2CID 256801383 Stanley S M 2008 Predation defeats competition on the seafloor Paleobiology 34 1 1 21 Bibcode 2008Pbio 34 1S 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 Bibcode 2007Pbio 33R 1S 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 Bibcode 2015PPP 428 1H 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 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 Xue Chunling Yuan Dong xun Chen Yanlong Stubbs Thomas L Zhao Yueli Zhang Zhifei 15 May 2024 Morphological innovation after mass extinction events in Permian and Early Triassic conodonts based on Polygnathacea Palaeogeography Palaeoclimatology Palaeoecology 642 112149 doi 10 1016 j palaeo 2024 112149 Retrieved 21 May 2024 via Elsevier Science Direct 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 Bibcode 2008Pbio 34 128S doi 10 1666 07053 1 S2CID 83650272 Crasquin Sylvie Forel Marie Beatrice Qinglai Feng Aihua Yuan Baudin Francois Collin Pierre Yves 30 July 2010 Ostracods Crustacea through the Permian Triassic boundary in South China the Meishan stratotype Zhejiang Province Journal of Systematic Palaeontology 8 3 331 370 Bibcode 2010JSPal 8 331C doi 10 1080 14772011003784992 S2CID 85986762 Retrieved 3 July 2023 a b Forel Marie Beatrice 1 August 2012 Ostracods Crustacea associated with microbialites across the Permian Triassic boundary in Dajiang Guizhou Province South China European Journal of Taxonomy 19 1 34 doi 10 5852 ejt 2012 19 Retrieved 3 July 2023 Powers Catherine M Bottjer David J 1 November 2007 Bryozoan paleoecology indicates mid Phanerozoic extinctions were the product of long term environmental stress Geology 35 11 995 Bibcode 2007Geo 35 995P doi 10 1130 G23858A 1 ISSN 0091 7613 Retrieved 30 December 2023 Liu Guichun Feng Qinglai Shen Jun Yu Jianxin He Weihong Algeo Thomas J 1 August 2013 Decline of Siliceous Sponges and Spicule Miniaturization Induced by Marine Productivity Collapse and Expanding Anoxia During the Permian Triassic Crisis in South China PALAIOS 28 8 664 679 doi 10 2110 palo 2013 p13 035r S2CID 128751510 Retrieved 31 May 2023 Song Haijun Tong Jinnan Chen Zhong Qiang Yang Hao Wang Yongbiao 14 July 2015 End Permian mass extinction of foraminifers in the Nanpanjiang basin South China Journal of Paleontology 83 5 718 738 doi 10 1666 08 175 1 S2CID 130765890 Retrieved 23 March 2023 Groves John R Altiner Demir Rettori Roberto 11 August 2017 Extinction Survival and Recovery of Lagenide Foraminifers in the Permian Triassic Boundary Interval Central Taurides Turkey Journal of Paleontology 79 S62 1 38 doi 10 1666 0022 3360 2005 79 1 ESAROL 2 0 CO 2 S2CID 130620805 Retrieved 23 March 2023 Guinot Guillaume Adnet Sylvain Cavin Lionel Cappetta Henri 29 October 2013 Cretaceous stem chondrichthyans survived the end Permian mass extinction Nature Communications 4 2669 Bibcode 2013NatCo 4 2669G doi 10 1038 ncomms3669 PMID 24169620 S2CID 205320689 Kear Benjamin P Engelschion Victoria S Hammer Oyvind Roberts Aubrey J Hurum Jorn H 13 March 2023 Earliest Triassic ichthyosaur fossils push back oceanic reptile origins Current Biology 33 5 R178 R179 doi 10 1016 j cub 2022 12 053 PMID 36917937 S2CID 257498390 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 Feng Yan Song Haijun Bond David P G 1 November 2020 Size variations in foraminifers from the early Permian to the Late Triassic implications for the Guadalupian Lopingian and the Permian Triassic mass extinctions Paleobiology 46 4 511 532 Bibcode 2020Pbio 46 511F doi 10 1017 pab 2020 37 S2CID 224855811 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 Rego Brianna L Wang Steve C Altiner Demir Payne Jonathan L 8 February 2016 Within and among genus components of size evolution during mass extinction recovery and background intervals a case study of Late Permian through Late Triassic foraminifera Paleobiology 38 4 627 643 doi 10 1666 11040 1 S2CID 17284750 Retrieved 23 March 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 Bibcode 2010Gbio 8 391H 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 Shi Guang R Zhang Yi chun Shen Shu zhong He Wei hong 15 April 2016 Nearshore offshore basin species diversity and body size variation patterns in Late Permian Changhsingian brachiopods Palaeogeography Palaeoclimatology Palaeoecology 448 96 107 Bibcode 2016PPP 448 96S doi 10 1016 j palaeo 2015 07 046 hdl 10536 DRO DU 30080099 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 60911307H 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 Forel Marie Beatrice Crasquin Sylvie Chitnarin Anisong Angiolini Lucia Gaetani Maurizio 25 February 2015 Precocious sexual dimorphism and the Lilliput effect in Neo Tethyan Ostracoda Crustacea through the Permian Triassic boundary Palaeontology 58 3 409 454 Bibcode 2015Palgy 58 409F doi 10 1111 pala 12151 hdl 2434 354270 S2CID 140198533 Retrieved 23 May 2023 Payne Jonathan L 8 April 2016 Evolutionary dynamics of gastropod size across the end Permian extinction and through the Triassic recovery interval Paleobiology 31 2 269 290 doi 10 1666 0094 8373 2005 031 0269 EDOGSA 2 0 CO 2 S2CID 7367192 Retrieved 23 March 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 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 Brayard Arnaud Meier Maximiliano Escarguel Gilles Fara Emmanuel Nutzel Alexander Olivier Nicolas Bylund Kevin G Jenks James F Stephen Daniel A Hautmann Michael Vennin Emmanuelle Bucher Hugo July 2015 Early Triassic Gulliver gastropods Spatio temporal distribution and significance for biotic recovery after the end Permian mass extinction Earth Science Reviews 146 31 64 Bibcode 2015ESRv 146 31B doi 10 1016 j earscirev 2015 03 005 Retrieved 20 January 2023 Atkinson Jed W Wignall Paul B Morton Jacob D Aze Tracy 9 January 2019 Body size changes in bivalves of the family Limidae in the aftermath of the end Triassic mass extinction the Brobdingnag effect Palaeontology 62 4 561 582 Bibcode 2019Palgy 62 561A doi 10 1111 pala 12415 S2CID 134070316 Retrieved 20 January 2023 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 R V Newman M 2003 Extinctions and Biodiversity in the Fossil Record In Canadell J G Mooney H A 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 Rees P McAllister 1 September 2002 Land plant diversity and the end Permian mass extinction Geology 30 9 827 830 Bibcode 2002Geo 30 827M doi 10 1130 0091 7613 2002 030 lt 0827 LPDATE gt 2 0 CO 2 Retrieved 31 May 2023 a b 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 a, 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.