Jump to content

Permian–Triassic extinction event

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

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

The scientific consensus is that the main cause of the extinction was the flood basalt volcanic eruptions that created the Siberian Traps,[19][20] which released sulfur dioxide and carbon dioxide, resulting in euxinia (oxygen-starved, sulfurous oceans),[21][22] elevating global temperatures,[23][24][25] and acidifying the oceans.[26][27][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.[23] 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;[28][29] emissions of methane from the gasification of methane clathrates;[30] emissions of methane by novel methanogenic microorganisms nourished by minerals dispersed in the eruptions;[31][32][33] longer and more intense El Niño events;[34] and an extraterrestrial impact which created the Araguainha crater and caused seismic release of methane[35][36][37] and the destruction of the ozone layer with increased exposure to solar radiation.[38][39][40]

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.[46] 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.[47][48][49]

The extinction occurred between 251.941 ± 0.037 and 251.880 ± 0.031 million years ago, a duration of 60 ± 48 thousand years.[50] A large, abrupt global decrease in δ13C, the ratio of the stable isotope carbon-13 to that of carbon-12, coincides with this extinction,[51][52][53] and is sometimes used to identify the Permian–Triassic boundary and PTME in rocks that are unsuitable for radiometric dating.[54][55] The negative carbon isotope excursion's magnitude was 4-7% and lasted for approximately 500 kyr,[56] though estimating its exact value is challenging due to diagenetic alteration of many sedimentary facies spanning the boundary.[57][58]

Further evidence for environmental change around the Permian-Triassic boundary suggests an 8 °C (14 °F) rise in temperature,[44] and an increase in CO
2
levels to 2,500 ppm (for comparison, the concentration immediately before the Industrial Revolution was 280 ppm,[44] and the amount today is about 422 ppm[59]). There is also evidence of increased ultraviolet radiation reaching the earth, causing the mutation of plant spores.[44][40]

It has been suggested that the Permian–Triassic boundary is associated with a sharp increase in the abundance of marine and terrestrial fungi, caused by the sharp increase in the amount of dead plants and animals fed upon by the fungi.[60] 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.[61] 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.[60] 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;[44][62] the spike did not appear worldwide;[63][64][65] and in many places it did not fall on the Permian–Triassic boundary.[66] 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.[67] Newer chemical evidence agrees better with a fungal origin for Reduviasporonites, diluting these critiques.[68][69]

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[70][71][72] 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.[73][66][74] 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,[75] and a study of the Shangsi section showed two extinction pulses with different causes too.[76] 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.[77] 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.[78] 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.[79]

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.[80] Many sedimentary sequences from South China show synchronous terrestrial and marine extinctions.[81] Research in the Sydney Basin of the PTME's duration and course also supports a synchronous occurrence of the terrestrial and marine biotic collapses.[82] Other scientists believe the terrestrial mass extinction began between 60,000 and 370,000 years before the onset of the marine mass extinction.[20][83] Chemostratigraphic analysis from sections in Finnmark and Trøndelag shows the terrestrial floral turnover occurred before the large negative δ13C shift during the marine extinction.[84] 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.[85] The Sunjiagou Formation of South China also records a terrestrial ecosystem demise predating the marine crisis.[86] Other research still has found that the terrestrial extinction occurred after the marine extinction in the tropics.[87]

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][88] 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][89] 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.[90][15][91] For example, all dinocephalian genera died out at the end of the Guadalupian,[89] as did the Verbeekinidae, a family of large-size fusuline foraminifera.[92] 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.[93][94]

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%
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%[95]

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.[96]

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.[97] Organisms that relied on haemocyanin or haemoglobin for transporting oxygen were more resistant to extinction than those utilising haemerythrin or oxygen diffusion.[98] 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.[99] There was little latitudinal difference in the survival rates of taxa.[100] Organisms that inhabited refugia less affected by global warming experienced lesser or delayed extinctions.[101]

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).[102][103] The extinction rate of marine organisms was catastrophic.[18][104][80] Bioturbators were extremely severely affected, as evidenced by the loss of the sedimentary mixed layer in many marine facies during the end-Permian extinction.[105]

Surviving marine invertebrate groups included articulate brachiopods (those with a hinge),[106] which had undergone a slow decline in numbers since the P–Tr extinction; the Ceratitida order of ammonites;[107] and crinoids ("sea lilies"),[107] 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.[108][109] In the case of the brachiopods, at least, surviving taxa were generally small, rare members of a formerly diverse community.[110]

Conodonts were severely affected both in terms of taxonomic and morphological diversity, although not as severely as during the Capitanian mass extinction.[111]

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.[112] 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.[113]

Ostracods experienced prolonged diversity perturbations during the Changhsingian before the PTME proper, when immense proportions of them abruptly vanished.[114] At least 74% of ostracods died out during the PTME itself.[115]

Bryozoans had been on a long-term decline throughout the Late Permian epoch before they suffered even more catastrophic losses during the PTME,[116] being the most severely affected clade among the lophophorates.[117]

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.[118]

Foraminifera suffered a severe bottleneck in diversity.[5] Evidence from South China indicates the foraminiferal extinction had two pulses.[119] Foraminiferal biodiversity hotspots shifted into deeper waters during the PTME.[120] 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.[121] Foraminifera that were calcaerous suffered an extinction rate of 91%.[122] 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.[123]

Cladodontomorph sharks likely survived the extinction by surviving in refugia in the deep oceans, a hypothesis based on the discovery of Early Cretaceous cladodontomorphs in deep, outer shelf environments.[124] Ichthyosaurs, which evolved immediately before the PTME, were also PTME survivors.[125]

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,[126][127] notably occurring in foraminifera,[128][129][130] brachiopods,[131][132][133] bivalves,[134][135][136] and ostracods.[137][138] Though gastropods that survived the cataclysm were smaller in size than those that did not,[139] it remains debated whether the Lilliput effect truly took hold among gastropods.[140][141][142] Some gastropod taxa, termed "Gulliver gastropods", ballooned in size during and immediately following the mass extinction,[143] exemplifying the Lilliput effect's opposite, which has been dubbed the Brobdingnag effect.[144]

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][145] according to some sources, it may well be the only mass extinction to significantly affect insect diversity.[146][147] 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.[46] However, some observed entomofaunal declines in the PTME were biogeographic changes rather than outright extinctions.[148]

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.[149] Plants are relatively immune to mass extinction, with the impact of all the major mass extinctions "insignificant" at a family level.[44][dubiousdiscuss] Floral diversity losses were more superficial than those of marine animals.[150] Even the reduction observed in species diversity (of 50%) may be mostly due to taphonomic processes.[151][44] However, a massive rearrangement of ecosystems does occur, with plant abundances and distributions changing profoundly and all the forests virtually disappearing.[152][44] The dominant floral groups changed, with many groups of land plants entering abrupt decline, such as Cordaites (gymnosperms) and Glossopteris (seed ferns).[153][154] The severity of plant extinction has been disputed.[155][151]

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.[156] The collapse of this flora is indirectly marked by an abrupt change in river morphology from meandering to braided river systems, signifying the widespread demise of rooted plants.[157]

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.[65] Data from Kap Stosch suggest that floral species richness was not significantly affected during the PTME.[158]

The Cordaites flora, which dominated the Angaran floristic realm corresponding to Siberia, collapsed over the course of the extinction.[159] 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.[160]

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.[161]

In South China, the subtropical Cathaysian gigantopterid dominated rainforests abruptly collapsed.[162][163][164] 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.[165] Wildfires too likely played a role in the fall of Gigantopteris.[166]

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.[167]

Terrestrial vertebrates

[edit]

The tempo of the terrestrial vertebrate extinction is disputed. Some evidence from the Karoo Basin indicates a protracted extinction lasting a million years.[168] Other evidence from the Karoo deposits suggest it took 50,000 years or less,[169] while a study of coprolites in the Vyazniki fossil beds in Russia suggests it took only a few thousand years.[170] Aridification induced by global warming was the chief culprit behind terrestrial vertebrate extinctions.[171][172] 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).[173][10] Tangasaurids were largely unaffected.[174] Gorgonopsians are traditionally thought to have gone extinct during the PTME, but some tentative evidence suggests they may have survived into the Triassic.[175] Freshwater and euryhaline fishes, having experienced minimal diversity losses before the PTME, were unaffected during the PTME and actually appear to have increased in diversity across the Permian-Triassic boundary.[176] However, faunal turnovers in freshwater fish communities occurred in areas like the Kuznetsk Basin.[177]

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.[178]

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.[179] 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.[180] Another study of latest Permian vertebrates in the Karoo Basin found that 54% of them went extinct due to the PTME.[181]

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;[182] the Permian-Triassic mass extinction marked a key turning point in this ecological shift that began after the Capitanian mass extinction[183] and culminated in the Late Jurassic.[184] Typical taxa of shelly benthic faunas were now bivalves, snails, sea urchins and Malacostraca, whereas bony fishes[185] and marine reptiles[186] 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).[187][188] However, recovery was also differential between taxa. Some survivors became extinct some million years after the extinction event without having rediversified (dead clade walking,[189] e.g. the snail family Bellerophontidae),[190] whereas others rose to dominance over geologic times (e.g., bivalves).[191][192]

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.[193] Marine post-extinction faunas were mostly species-poor and were dominated by few disaster taxa such as the bivalves Claraia, Unionites, Eumorphotis, and Promyalina,[194] the conodonts Clarkina and Hindeodus,[195] the inarticulate brachiopod Lingularia,[194] and the foraminifera Earlandia and Rectocornuspira kalhori,[196] the latter of which is sometimes classified under the genus Ammodiscus.[197] Their guild diversity was also low.[198] Post-PTME faunas had a flat, insignificant latitudinal diversity gradient.[199]

The speed of recovery from the extinction is disputed. Some scientists estimate that it took 10 million years (until the Middle Triassic) due to the severity of the extinction.[200] However, studies in Bear Lake County, near Paris, Idaho,[201] and nearby sites in Idaho and Nevada[202] showed a relatively quick rebound in a localized Early Triassic marine ecosystem (Paris biota), taking around 1.3 million years to recover,[201] while an unusually diverse and complex ichnobiota is known from Italy less than a million years after the end-Permian extinction.[203] Additionally, the complex Guiyang biota found near Guiyang, China also indicates life thrived in some places just a million years after the mass extinction,[204][205] as does a fossil assemblage known as the Shanggan fauna found in Shanggan, China,[206] the Wangmo biota from the Luolou Formation of Guizhou,[207] and a gastropod fauna from the Al Jil Formation of Oman.[208] Regional differences in the pace of biotic recovery existed,[209][210] 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.[211][212] 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;[213] 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.[214]

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.[215] Epifaunal benthos took longer to recover than infaunal benthos.[216] 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,[217] and conodonts, which diversified considerably over the first two million years of the Early Triassic.[218]

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.[219] That recovery was slow in the Early Triassic can be explained by low levels of biological competition due to the paucity of taxonomic diversity,[220] 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.[221] Biodiversity rise thus behaved as a positive feedback loop enhancing itself as it took off in the Spathian and Anisian.[222] 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.[220] Other explanations state that life was delayed in its recovery because grim conditions returned periodically over the course of the Early Triassic,[223][224] causing further extinction events, such as the Smithian-Spathian boundary extinction.[11][225][226] Continual episodes of extremely hot climatic conditions during the Early Triassic have been held responsible for the delayed recovery of oceanic life,[227][228] in particular skeletonised taxa that are most vulnerable to high carbon dioxide concentrations.[229] The relative delay in the recovery of benthic organisms has been attributed to widespread anoxia,[230] but high abundances of benthic species contradict this explanation.[231] 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.[212] Recurrent Early Triassic environmental stresses also acted as a ceiling limiting the maximum ecological complexity of marine ecosystems until the Spathian.[232] Recovery biotas appear to have been ecologically uneven and unstable into the Anisian, making them vulnerable to environmental stresses.[233]

Whereas most marine communities were fully recovered by the Middle Triassic,[234][235] global marine diversity reached pre-extinction values no earlier than the Middle Jurassic, approximately 75 million years after the extinction event.[236]

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.[237] 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,[237] and the increase in predation pressure and durophagy led to the Mesozoic Marine Revolution.[238]

Marine vertebrates recovered relatively quickly, with complex predator-prey interactions with vertebrates at the top of the food web being known from coprolites five million years after the PTME.[239] Post-PTME hybodonts exhibited extremely rapid tooth replacement.[240] Ichthyopterygians appear to have ballooned in size extremely rapidly following the PTME.[241]

Bivalves rapidly recolonised many marine environments in the wake of the catastrophe.[242] 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.[243] Bivalves were once thought to have outcompeted brachiopods, but this outdated hypothesis about the brachiopod-bivalve transition has been disproven by Bayesian analysis.[244] 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,[184] whilst other studies have emphasised the greater niche breadth of the former.[245] 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.[246] 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.[247] Infaunal habits in bivalves became more common after the PTME.[248]

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.[249] The surviving brachiopod fauna was very low in diversity and exhibited no provincialism whatsoever.[250] 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.[251] 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.[252] 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.[253]

Ostracods were extremely rare during the basalmost Early Triassic.[254] Taxa associated with microbialites were disproportionately represented among ostracod survivors.[115] Ostracod recovery began in the Spathian.[255] Despite high taxonomic turnover, the ecological life modes of Early Triassic ostracods remained rather similar to those of pre-PTME ostracods.[256]

Bryozoans in the Early Triassic were restricted to the Boreal realm.[117] They were also not diverse, represented mainly by members of Trepostomatida. During the Middle Triassic, there was a rise in bryozoan diversity, which peaked in the Carnian.[257] However, bryozoans took until the Late Cretaceous to recover their full diversity.[258]

Crinoids ("sea lilies") suffered a selective extinction, resulting in a decrease in the variety of their forms.[259] 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.[260] 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.[261] Though their taxonomic diversity remained relatively low, crinoids regained much of their ecological dominance by the Middle Triassic epoch.[246] Stem-group echinoids survived the PTME.[262] The survival of miocidarid echinoids such as Eotiaris is likely attributable to their ability to thrive in a wide range of environmental conditions.[263]

Conodonts saw a rapid recovery during the Induan,[264] 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.[265] Segminiplanate conodonts again experienced a brief period of domination in the early Spathian, probably related to a transient oxygenation of deep waters.[266] Neospathodid conodonts survived the crisis but underwent proteromorphosis.[267]

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.[268] In the Tethys, foraminiferal communities remained low in diversity into the Middle Triassic, with the exception of a notable Ladinian fauna from the Catalonian Basin.[269]

Microbial reefs were common across shallow seas for a short time during the earliest Triassic,[270][271] predominating in low latitudes while being rarer in higher latitudes,[272] occurring both in anoxic and oxic waters.[273] Polybessurus-like microfossils often dominated these earliest Triassic microbialites.[274] Microbial-metazoan reefs appeared very early in the Early Triassic;[275] and they dominated many surviving communities across the recovery from the mass extinction.[276] Microbialite deposits appear to have declined in the early Griesbachian synchronously with a significant sea level drop that occurred then.[273] Metazoan-built reefs reemerged during the Olenekian, mainly being composed of sponge biostrome and bivalve builups.[276] Keratose sponges were particularly noteworthy in their integral importance to Early Triassic microbial-metazoan reef communities,[277][278] and they helped to create stability in heavily damaged ecosystems during early phases of biotic recovery.[279] "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.[276] Bryozoans, after sponges, were the most numerous organisms in Tethyan reefs during the Anisian.[280] Metazoan reefs became common again during the Anisian because the oceans cooled down then from their overheated state during the Early Triassic.[281] Biodiversity amongst metazoan reefs did not recover until well into the Anisian, millions of years after non-reef ecosystems recovered their diversity.[282] 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.[283] The microbialite refuge hypothesis has been criticised as reflecting a taphonomic bias due to the greater preservation potential of microbialite deposits, however, rather than a genuine phenomenon.[284]

Ichnocoenoses show that marine ecosystems recovered to pre-extinction levels of ecological complexity by the late Olenekian.[285] 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.[286]

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.[287][288] 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.[289] 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.[290]

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.[291] Furthermore, severe extinction pulses continued to occur after the Permian-Triassic boundary, causing additional floral turnovers.[292] 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.[293] 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.[159] The worldwide recovery of gymnosperm forests took approximately 4–5 million years.[294][44] 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,[295] and floral recovery likely did not follow a congruent, globally universal trend but instead varied by region according to local environmental conditions.[164]

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.[65]

In what is now the Barents Sea of the coast of Norway, the post-extinction flora 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.[295]

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.[296]

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.[297] 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.[164] Plant survivors of the PTME in South China experienced extremely high rates of mutagenesis induced by heavy metal poisoning.[298] 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.[297]

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.[299]

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.[300]

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.[156]

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.[45] Abiotic factors (factors not caused by organisms), such as decreased rainfall or increased input of clastic sediments, may also be to blame.[44]

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.[44] 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.[301] Approximately 95% of these peat-producing plants went locally extinct at the P–Tr boundary;[302] coal deposits in Australia and Antarctica disappear significantly before the P–Tr boundary.[44]

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.[303] East-central Pangaea, with its relatively wet climate, served as a dispersal corridor for PTME survivors during their Early Triassic recolonisation of the supercontinent.[304] 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.[305][306] 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.[307] 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.[308] 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.[309]

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.[310] The dicynodont genus is often used as a biostratigraphic marker for the PTME.[311] 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.[312] Other likely factors behind the success of Lystrosaurus included extremely fast growth rate exhibited by the dicynodont genus,[313] along with its early onset of sexual maturity.[314] Antarctica may have served as a refuge for dicynodonts during the PTME from which surviving dicynodonts spread out of in its aftermath.[315] Ichnological evidence from the earliest Triassic of the Karoo Basin shows dicynodonts were abundant in the immediate aftermath of the biotic crisis.[316] Smaller carnivorous cynodont therapsids also survived, a group that included the ancestors of mammals.[317] As with dicynodonts, selective pressures favoured endothermic epicynodonts.[318] Therocephalians likewise survived; burrowing may have been a key adaptation that helped them make it through the PTME.[319] 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.[320] Early Triassic therocephalians were mostly survivors of the PTME rather than newly evolved taxa that originated during the evolutionary radiation in its aftermath.[321] Both therocephalians and cynodonts, known collectively as eutheriodonts, decreased in body size from the Late Permian to the Early Triassic.[317] This decrease in body size has been interpreted as an example of the Lilliput effect.[322]

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.[323] 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.[324] 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,[325] 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.[326] 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.[327] During the Induan, tupilakosaurids in particular thrived as disaster taxa,[328] including Tupilakosaurus itself,[329] though they gave way to other temnospondyls as ecosystems recovered.[328] Temnospondyls were reduced in size during the Induan, but their body size rebounded to pre-PTME levels during the Olenekian.[330] Mastodonsaurus and trematosaurians were the main aquatic and semiaquatic predators during most of the Triassic, some preying on tetrapods and others on fish.[331]

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.[146]

Microbially induced sedimentary structures (MISS) dominated North Chinese terrestrial fossil assemblages in the Early Triassic.[332][333] In Arctic Canada as well, MISS became a common occurrence following the Permian-Triassic extinction.[334] 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.[333]

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.[97]

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.[335][336][337][338][339] The date of the Siberian Traps eruptions matches well with the extinction event.[50][340][341][19][342] 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.[343]

The Siberian Traps caused one of the most rapid rises of atmospheric carbon dioxide levels in the geologic record,[344] with the rate of carbon dioxide emissions estimated as five times faster than during the preceding catastrophic Capitanian mass extinction[345] during the eruption of the Emeishan Traps.[346][347][348] 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.[22] 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.[23] Extreme temperature rise would have followed,[349] though some evidence suggests a lag of 12,000 to 128,000 years between the rise in volcanic carbon dioxide emissions and global warming.[350] During the latest Permian before the extinction, global average surface temperatures were about 18.2 °C,[351] which shot up to as much as 35 °C, this hyperthermal condition lasting as long as 500,000 years.[23] Air temperatures at Gondwana's high southern latitudes experienced a warming of ~10–14 °C.[24] According to oxygen isotope shifts from conodont apatite in South China, low latitude surface water temperatures surged about 8 °C.[25] 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.[352] The increased mean state temperatures also brought stronger El Nino events, heightening short-term climate variability.[34]

These extremely high atmospheric carbon dioxide concentrations persisted over a long period.[353] The position and alignment of Pangaea at the time made the inorganic carbon cycle very inefficient at burying carbon.[354] 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.[355][9] Evidence also points to volcanic combustion of underground fossil fuel deposits, based on paired coronene-mercury spikes[356][31] coinciding with geographically widespread mercury anomalies and the rise in isotopically light carbon.[357] Te/Th values increase twentyfold over the PTME, further indicating it was concomitant with extreme volcanism.[358] 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.[359]

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.[360] Around 18 teratonnes of hydrochloric acid were emitted,[361] 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,[362] with glacio-eustatic sea level fall.[360][363] However, the briefness of these cold events makes them unlikely to have been a significant kill mechanism.[364]

The eruptions may also have caused acid rain as the aerosols washed out of the atmosphere.[365] 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.[366] When this had washed out of the atmosphere, the excess carbon dioxide would have remained and global warming would have proceeded unchecked.[349]

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.[367] The unique setting of the Siberian Traps over these deposits is likely the reason for the severity of the extinction.[368][369][370] 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.[349] 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[28] and is linked to a major negative δ13C excursion.[371] The intermediate temperature of the Siberian Traps magmas optimised the extremely voluminous release of CO2 by way of heating of evaporites and carbonates.[372]

Venting of coal-derived methane was accompanied by explosive combustion of coal and discharge of coal-fly ash.[29] 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."[373][374] 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."[375] However, some researchers propose that these supposed fly ashes were actually the result of wildfires not related to massive coal combustion by intrusive magmatism.[376] 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.[377]

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.[378] 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.[228]

The reduction in marine primary productivity diminished emissions of dimethyl sulphate and dimethylsulphoniopropionate, enhancing warming.[379] Also, 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 by depleting yet another siliceous carbon sink.[380]

Mercury anomalies corresponding to the time of Siberian Traps activity have been found in many geographically disparate sites,[381][382][383] indicating that these volcanic eruptions released significant quantities of toxic mercury into the atmosphere and ocean, causing even larger terrestrial and marine die-offs.[384][385][386] 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.[387][388] Mutagenesis in surviving plants after the PTME coeval with mercury and copper loading confirms volcanically induced heavy metal toxicity.[298] Increased bioproductivity may have sequestered mercury and party mitigated poisoning.[389] Immense volumes of nickel aerosols and cobalt and arsenic emisions, were also released,[390][391][384] further contributing to metal poisoning.[392]

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,[393][394] a finding corroborated by mercury spikes,[395] causing further extinction events during the epoch.[396]

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.[156] At 1,300,000 cubic kilometres in volume[397] and 1,680,000 square kilometres in area, this event was 40% the size of the Siberian Traps.[156] Specifically, this flood basalt has been implicated in the regional demise of the Gondwanan Glossopteris flora.[398]

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.[399] The enrichment of copper samples from these deposits in isotopically light copper provide additional confirmation.[400] 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.[399]

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.[401]

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.[80][402] 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.[108] The seabed probably contained methane hydrate deposits, and the lava caused the deposits to dissociate, releasing vast quantities of methane.[403] 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);[404] 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.[405] 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.[406]

The clathrate hypothesis seemed the only proposed mechanism sufficient to cause a global 1% reduction in the 13C 12C ratio .[407][46] 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:[30]

  • 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.[408] (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.[408]
  • 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.[46] Moreover, sea levels were rising rather than falling at the time of the extinction.[349]
  • 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;[108] 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.[409] That or another organic-based reason may have been responsible for both that and a late Proterozoic/Cambrian pattern of fluctuating 13C 12C ratios.[108]

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.[30][410] 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.[108] The latest research suggests that greenhouse gas release during the extinction event was dominated by volcanic carbon dioxide,[411] 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.[23]

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.[412] 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.[413]

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.[179] Higher concentrations of CO2 also result in decreased activity levels in many active marine animals, hindering their ability to obtain food.[414] 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.[179]

In addition, an increase in CO2 concentration is inevitably linked to ocean acidification,[415] consistent with the preferential extinction of heavily calcified taxa and other signals in the rock record that suggest a more acidic ocean,[27] such as a carbonate production crisis that occurred a few thousand years after volcanic greenhouse gas emissions began.[416] The decrease in ocean pH is calculated to be up to 0.7 units.[26] An extreme aragonite sea formed.[417] 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.[418] 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.[419] 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.[420] The persistence of highly elevated carbon dioxide concentrations in the atmosphere and ocean during the Early Triassic would have impeded the recovery of biocalcifying organisms after the PTME.[421]

Acidity generated by increased carbon dioxide concentrations in soil and sulphur dioxide dissolution in rainwater was also a kill mechanism on land.[422] 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.[423] Further evidence of an increase in soil acidity comes from elevated Ba/Sr ratios in earliest Triassic soils.[424] 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.[425] 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.[426] Hopane anomalies have also been interpreted as evidence of acidic soils and peats.[427] As with many other environmental stressors, acidity on land episodically persisted well into the Triassic, stunting the recovery of terrestrial ecosystems.[428]

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.[429][430][431] Throughout most of the Tethys and Panthalassic Oceans, evidence for anoxia appears at the extinction event, including small pyrite framboids,[432][433] negative δ238U excursions,[434][435] negative δ15N excursions,[436] positive δ82/78Se isotope excursions,[437] relatively positive δ13C ratios in polycyclic aromatic hydrocarbons,[438] high Th/U ratios,[439][434] positive Ce/Ce* anomalies,[440] depletions of molybdenum, uranium, and vanadium from seawater,[441] and fine laminations in sediments.[432] However, evidence for anoxia precedes the extinction at some other sites, including Spiti, India,[442] Shangsi, China,[443] Meishan, China,[444] Opal Creek, Alberta,[445] and Kap Stosch, Greenland.[446] Biogeochemical evidence also points to the presence of euxinia during the PTME.[447] 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.[448][449] Negative mercury isotope excursions further bolster evidence for extensive euxinia during the PTME.[450] 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.[451] 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.[452] The sequence of extinctions in some sections, with deep water organisms being affected first followed by shallow water and then by bottom water organisms, is believed to reflect the migration of oxygen minimum zones.[453] Models of ocean chemistry suggest that anoxia and euxinia were closely associated with hypercapnia. 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.[454]

The sequence of events leading to anoxic oceans may have been triggered by carbon dioxide emissions from the eruption of the Siberian Traps.[455] 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.[456] The flux of terrigeneous material into the oceans increased as a result of soil erosion, which would have facilitated increased eutrophication;[457] marine regression likewise enhanced terrigeneous material inputs.[458] 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.[459] 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.[460] 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.[461] 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.[462] Convective overturn helped facilitate the expansion of anoxia throughout the water column.[463] 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.[455] In some regions, anoxia briefly disappeared when transient cold snaps resulting from volcanic sulphur emissions occurred.[464]

The persistence of anoxia through the Early Triassic may explain the slow recovery of marine life and low levels of biodiversity after the extinction,[465][466][467] particularly that of benthic organisms.[468][230] Anoxia disappeared from shallow waters more rapidly than the deep ocean.[469] Reexpansions of oxygen-minimum zones did not cease until the late Spathian, periodically setting back and restarting the biotic recovery process.[470] The decline in continental weathering towards the end of the Spathian at last began ameliorating marine life from recurrent anoxia.[471] In some regions of Panthalassa, pelagic zone anoxia continued to recur as late as the Anisian,[472] probably due to increased productivity and a return of aeolian upwelling.[473] Some sections show a rather quick return to oxic water column conditions, however, so for how long anoxia persisted remains debated.[474] The volatility of the Early Triassic sulphur cycle suggests marine life continued to face returns of euxinia as well.[475]

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.[476]

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.[477]

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.[478] The combination of global warming and drying generated an increased incidence of wildfires.[479] Tropical coastal swamp floras such as those in South China may have been very detrimentally impacted by the increase in wildfires,[166] though it is ultimately unclear if an increase in wildfires played a role in driving taxa to extinction.[480]

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.[481] 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.[482] A pattern of aridity-induced extinctions that progressively ascended up the food chain has been deduced from Karoo Basin biostratigraphy.[171] 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.[483] 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.[484] 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.[485] 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.[160]

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.[306][305] Elsewhere, such as in the Karoo Basin, episodes of dry climate recurred regularly in the Early Triassic, with profound effects on terrestrial tetrapods.[478]

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.[486]

Ozone depletion

[edit]

A collapse of the atmospheric ozone shield has been invoked as an explanation for the mass extinction,[487][488] particularly that of terrestrial plants.[39] 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.[489] 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.[490][40] 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.[38] Extremely positive Δ33S anomalies provide evidence of photolysis of volcanic SO2, indicating increased ultraviolet radiation flux.[491] Sulphur isotope data from North China are inconsistent with a total collapse of the ozone layer, however, suggesting it may have not been as major a kill mechanism as others.[492]

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.[493] Volcanic emissions of sulphate aerosols into the stratosphere would have dealt significant destruction to the ozone layer.[490] As mentioned previously, the rocks in the region where the Siberian Traps were emplaced are extremely rich in halogens.[360] 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.[494][40] An uptick in wildfires, a natural source of methyl chloride, would have had further deleterious effects still on the atmospheric ozone shield.[495] 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,[496] although other modelling work has found that the release of this gas would not have significantly damaged the ozone layer.[497] 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.[498]

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.[499] However, suggestions that an asteroid impact was the trigger of the Permian-Triassic extinction are now largely rejected.[500][20]

Reported evidence for an impact event from the P–Tr boundary level includes rare grains of shocked quartz in Australia and Antarctica;[501][502] fullerenes trapping extraterrestrial noble gases;[503] meteorite fragments in Antarctica;[504] and grains rich in iron, nickel, and silicon, which may have been created by an impact.[505] However, the accuracy of most of these claims has been challenged.[506][507][508][509] 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.[510] Iridium levels in many sites straddling the Permian-Triassic boundaries are not anomalous, providing evidence against an extraterrestrial impact as the PTME's cause.[511]

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.[512] 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,[349] such as the Siberian Traps eruptions at either an impact site[513] or the antipode of an impact site.[349][514] 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.[515] Additionally, many sites spanning the Permian-Triassic boundary display a complete lack of evidence of an impact event.[516]

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[502] and the hypothesized 480 km (300 mi) Wilkes Land crater of East Antarctica.[517][518] 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,[35] precipitated the mass extinction.[519] 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.[36] The large earthquakes generated by the impact would have additionally generated massive tsunamis across much of the globe.[519][37] 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.[36]

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),[520] 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.[521]

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.[32] Chemostratigraphic analysis of Permian-Triassic boundary sediments in Chaotian demonstrates a methanogenic burst could be responsible for some percentage of the carbon isotopic fluctuations.[33] On the other hand, in the canonical Meishan sections, the nickel concentration increases somewhat after the δ13C concentrations have begun to fall.[522]

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.[523]

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.[344] 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.[524] Another biocalcification event similar in its effects on modern marine ecosystems is predicted to occur if carbon dioxide levels continue to rise.[421] The changes in plant-insect interactions resulting from the PTME have also been invoked as possible indicators of the world's future ecology.[525] 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.[344]

Just as during the PTME, contemporary oceans experience their extreme change-change in the form of a decline in pH and oxygen levels, which further strengthens the pull between the two events. This is emphasised by geologist Lee Kump:

“The Permian-Triassic mass extinction provides a stark reminder of the consequences of rapid carbon dioxide emissions. During the PTME, volcanic activity unleashed massive amounts of CO₂, leading to ocean acidification, deoxygenation, and widespread ecological collapse. Today, we see human activities driving similar processes at an even faster rate. The geological record shows that once these tipping points are reached, the cascading effects on ecosystems can last for millions of years.”[526] [527]

If it continues to rise, the consequence could be another bio-calcification crisis, as seems to have occurred in the fossil record, which would have disastrous consequences for modern marine ecosystems.

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. ^ ""Great Dying" lasted 200,000 years". 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). "The P–T Extinction was a Slow Death". 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. hdl:11573/1707839. 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 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. ^ a b c 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. overwhelming data support that the [Permian-Triassic mass extinction] was triggered by the eruption of the [Siberian Traps Large Igneous Province].
  21. ^ 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.
  22. ^ 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.
  23. ^ a b c d e Wu, Yuyang; Chu, Daoliang; Tong, Jinnan; Song, Haijun; Dal Corso, Jacopo; Wignall, Paul Barry; 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.
  24. ^ 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.
  25. ^ 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.
  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.
  27. ^ 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.
  28. ^ 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.
  29. ^ 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.
  30. ^ 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.
  31. ^ 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.
  32. ^ 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.
  33. ^ 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.
  34. ^ a b Yadong Sun; Alexander Farnsworth; Michael M. Joachimski; Paul Barry Wignall; Leopold Krystyn; David P. G. Bond; Domenico C. G. Ravidà; Paul J. Valdes (September 12, 2024). "Mega El Niño instigated the end-Permian mass extinction". Science. 385 (6714): 1189–1195. doi:10.1126/science.ado2030. PMID 39265011.
  35. ^ 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.
  36. ^ 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.
  37. ^ 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.
  38. ^ 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.
  39. ^ 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.
  40. ^ 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.
  41. ^ 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.
  42. ^ 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.
  43. ^ 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.
  44. ^ 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.
  45. ^ 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.
  46. ^ 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.
  47. ^ 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.
  48. ^ 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.
  49. ^ Kershaw, Stephen (1 April 2017). "Palaeogeographic variation in the Permian–Triassic boundary microbialites: A discussion of microbial and ocean processes after the end-Permian mass extinction". Journal of Palaeogeography. 6 (2): 97–107. doi:10.1016/j.jop.2016.12.002. ISSN 2095-3836.
  50. ^ 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.
  51. ^ 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.
  52. ^ 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.
  53. ^ 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.
  54. ^ 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.
  55. ^ Yuan, Dong-xun; Chen, Jun; Zhang, Yi-chun; Zheng, Quan-feng; Shen, Shu-zhong (1 June 2015). "Changhsingian conodont succession and the end-Permian mass extinction event at the Daijiagou section in Chongqing, Southwest China". Journal of Asian Earth Sciences. 105: 234–251. doi:10.1016/j.jseaes.2015.04.002. Retrieved 18 June 2024 – via Elsevier Science Direct.
  56. ^ 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.
  57. ^ 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.
  58. ^ Schobben, Martin; Heuer, Franziska; Tietje, Melanie; Ghaderi, Abbas; Korn, Dieter; Korte, Christoph; Wignall, Paul Barry (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.
  59. ^ "Daily CO2". Mauna Loa Observatory.
  60. ^ 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.
  61. ^ 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.
  62. ^ 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.
  63. ^ 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.
  64. ^ 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.
  65. ^ 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
  66. ^ 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.
  67. ^ 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.
  68. ^ 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.
  69. ^ 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.
  70. ^ Shen, Jun; Feng, Qinglai; Algeo, Thomas J.; Li, Chao; Planavsky, Noah J.; Zhou, Lian; Zhang, Mingliang (1 June 2016). "Two pulses of oceanic environmental disturbance during the Permian–Triassic boundary crisis". Earth and Planetary Science Letters. 443: 139–152. doi:10.1016/j.epsl.2016.03.030. Retrieved 18 June 2024 – via Elsevier Science Direct.
  71. ^ 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.
  72. ^ 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.
  73. ^ 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.
  74. ^ 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.
  75. ^ 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.
  76. ^ 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.
  77. ^ 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.
  78. ^ 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.
  79. ^ 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.
  80. ^ 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.
  81. ^ 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.
  82. ^ 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.
  83. ^ 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.
  84. ^ 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.
  85. ^ 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.
  86. ^ 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.
  87. ^ 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.
  88. ^ 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.
  89. ^ 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.
  90. ^ Li, Dirson Jian (18 December 2012). "Tectonic cause of mass extinctions and the genomic contribution to biodiversification". arXiv:1212.4229 [q-bio.PE].
  91. ^ 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.
  92. ^ 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.
  93. ^ 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.
  94. ^ 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.
  95. ^ 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.
  96. ^ Bambach, R. K.; Knoll, A. H.; Wang, S. C. (December 2004). "Origination, extinction, and mass depletions of marine diversity". Paleobiology. 30 (4): 522–542. Bibcode:2004Pbio...30..522B. doi:10.1666/0094-8373(2004)030<0522:OEAMDO>2.0.CO;2. ISSN 0094-8373. S2CID 17279135.
  97. ^ 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.
  98. ^ 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. PMC 11025005. PMID 38638583.
  99. ^ 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.
  100. ^ Allen, Bethany J.; Clapham, Matthew E.; Saupe, Erin E.; Wignall, Paul Barry; 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.
  101. ^ Jiang, Haishui; Joachimski, Michael M.; Wignall, Paul Barry; Zhang, Muhui; Lai, Xulong (15 December 2015). "A delayed end-Permian extinction in deep-water locations and its relationship to temperature trends (Bianyang, Guizhou Province, South China)". Palaeogeography, Palaeoclimatology, Palaeoecology. 440: 690–695. doi:10.1016/j.palaeo.2015.10.002. Retrieved 13 October 2024 – via Elsevier Science Direct.
  102. ^ 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.
  103. ^ 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.
  104. ^ 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.
  105. ^ 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.
  106. ^ "Permian: The Marine Realm and The End-Permian Extinction". paleobiology.si.edu. Retrieved 2016-01-26.
  107. ^ a b "Permian extinction". Encyclopædia Britannica. Retrieved 2016-01-26.
  108. ^ 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.
  109. ^ 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.
  110. ^ 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.
  111. ^ 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.
  112. ^ 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.
  113. ^ 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.
  114. ^ 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.
  115. ^ 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.
  116. ^ 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.
  117. ^ a b Powers, Catherine M.; Bottjer, David J. (2 November 2009). "Behavior of lophophorates during the end-Permian mass extinction and recovery". Journal of Asian Earth Sciences. 36 (6): 413–419. doi:10.1016/j.jseaes.2008.05.002. Retrieved 28 October 2024 – via Elsevier Science Direct.
  118. ^ 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.
  119. ^ Song, Haijun; Tong, Jinnan; Chen, Zhong-Qiang Chen (13 July 2008). "Two episodes of foraminiferal extinction near the Permian–Triassic boundary at the Meishan section, South China". Australian Journal of Earth Sciences. 56 (6): 765–773. doi:10.1080/08120090903002599. ISSN 0812-0099. Retrieved 28 October 2024 – via Taylor and Francis Online.
  120. ^ Liu, Xiaokang; Song, Haijun; Bond, David P.G.; Tong, Jinnan; Benton, Michael James (October 2020). "Migration controls extinction and survival patterns of foraminifers during the Permian-Triassic crisis in South China". Earth-Science Reviews. 209: 103329. doi:10.1016/j.earscirev.2020.103329. Retrieved 28 October 2024 – via Elsevier Science Direct.
  121. ^ 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.
  122. ^ Groves, John R.; Altiner, Demír (September–October 2005). "Survival and recovery of calcareous foraminifera pursuant to the end-Permian mass extinction". Comptes Rendus Palevol. 4 (6–7): 487–500. doi:10.1016/j.crpv.2004.12.007. Retrieved 28 October 2024 – via Elsevier Science Direct.
  123. ^ 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.
  124. ^ 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.
  125. ^ 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.
  126. ^ 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.
  127. ^ 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.
  128. ^ 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.
  129. ^ 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.
  130. ^ 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.
  131. ^ 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.
  132. ^ 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.
  133. ^ 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.
  134. ^ 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.
  135. ^ 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.
  136. ^ 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.
  137. ^ 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.
  138. ^ 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.
  139. ^ 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.
  140. ^ 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.
  141. ^ 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.
  142. ^ 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.
  143. ^ 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.
  144. ^ Atkinson, Jed W.; Wignall, Paul Barry; 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.
  145. ^ Ponomarenko, A. G. (13 May 2016). "Insects during the time around the Permian—Triassic crisis". Paleontological Journal. 50 (2): 174–186. doi:10.1134/S0031030116020052. ISSN 0031-0301. Retrieved 13 October 2024 – via Springer Link.
  146. ^ 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.
  147. ^ 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.
  148. ^ Shcherbakov, D. E. (30 January 2008). "On Permian and Triassic insect faunas in relation to biogeography and the Permian-Triassic crisis". Paleontological Journal. 42 (1): 15–31. doi:10.1134/S0031030108010036. ISSN 0031-0301. Retrieved 18 June 2024 – via Springer Link.
  149. ^ 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.
  150. ^ Ponomarenko, A. G. (August 2006). "Changes in terrestrial biota before the Permian-Triassic ecological crisis". Paleontological Journal. 40 (4): S468–S474. doi:10.1134/S0031030106100066. ISSN 0031-0301. Retrieved 13 October 2024 – via Springer Link.
  151. ^ 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:10.1038/s41467-018-07945-w. PMC 6344494. PMID 30674875.
  152. ^ "The Dino Directory – Natural History Museum".
  153. ^ Gulbranson, Erik L.; Mellum, Morgan M.; Corti, Valentina; Dahlseid, Aidan; Atkinson, Brian A.; Ryberg, Patricia E.; Cornamusini, Giancula (24 May 2022). "Paleoclimate-induced stress on polar forested ecosystems prior to the Permian–Triassic mass extinction". Scientific Reports. 12 (1): 8702. Bibcode:2022NatSR..12.8702G. doi:10.1038/s41598-022-12842-w. PMC 9130125. PMID 35610472.
  154. ^ Retallack, G. J. (1995). "Permian–Triassic life crisis on land". Science. 267 (5194): 77–80. Bibcode:1995Sci...267...77R. doi:10.1126/science.267.5194.77. PMID 17840061. S2CID 42308183.
  155. ^ Cascales-Miñana, B.; Cleal, C. J. (2011). "Plant fossil record and survival analyses". Lethaia. 45: 71–82. doi:10.1111/j.1502-3931.2011.00262.x.
  156. ^ a b c d Bodnar, Josefina; Coturel, Eliana P.; Falco, Juan Ignacio; Beltrána, Marisol (4 March 2021). "An updated scenario for the end-Permian crisis and the recovery of Triassic land flora in Argentina". Historical Biology. 33 (12): 3654–3672. Bibcode:2021HBio...33.3654B. doi:10.1080/08912963.2021.1884245. S2CID 233810158. Retrieved 24 December 2022.
  157. ^ Ward, Peter Douglas; Montgomery, David R.; Smith, Roger (8 September 2000). "Altered River Morphology in South Africa Related to the Permian-Triassic Extinction". Science. 289 (5485): 1740–1743. doi:10.1126/science.289.5485.1740. ISSN 0036-8075. PMID 10976065. Retrieved 13 October 2024.
  158. ^ Schneebeli-Hermann, Elke; Hochuli, Peter A.; Bucher, Hugo (August 2017). "Palynofloral associations before and after the Permian–Triassic mass extinction, Kap Stosch, East Greenland". Global and Planetary Change. 155: 178–195. doi:10.1016/j.gloplacha.2017.06.009. Retrieved 11 September 2024 – via Elsevier Science Direct.
  159. ^ a b Vajda, Vivi; McLoughlin, Stephen (April 2007). "Extinction and recovery patterns of the vegetation across the Cretaceous–Palaeogene boundary — a tool for unravelling the causes of the end-Permian mass-extinction". Review of Palaeobotany and Palynology. 144 (1–2): 99–112. Bibcode:2007RPaPa.144...99V. doi:10.1016/j.revpalbo.2005.09.007. Retrieved 24 December 2022.
  160. ^ a b Davydov, V. I.; Karasev, E. V.; Nurgalieva, N. G.; Schmitz, M. D.; Budnikov, I. V.; Biakov, A. S.; Kuzina, D. M.; Silantiev, V. V.; Urazaeva, M. N.; Zharinova, V. V.; Zorina, S. O.; Gareev, B.; Vasilenko, D. V. (1 July 2021). "Climate and biotic evolution during the Permian-Triassic transition in the temperate Northern Hemisphere, Kuznetsk Basin, Siberia, Russia". Palaeogeography, Palaeoclimatology, Palaeoecology. 573: 110432. Bibcode:2021PPP...57310432D. doi:10.1016/j.palaeo.2021.110432. S2CID 235530804. Retrieved 19 December 2022.
  161. ^ Xiong, Conghui; Wang, Jiashu; Huang, Pu; Cascales-Miñana, Borja; Cleal, Christopher J.; Benton, Michael James; Xue, Jinzhuang (December 2021). "Plant resilience and extinctions through the Permian to Middle Triassic on the North China Block: A multilevel diversity analysis of macrofossil records". Earth-Science Reviews. 223: 103846. Bibcode:2021ESRv..22303846X. doi:10.1016/j.earscirev.2021.103846. hdl:20.500.12210/76649. S2CID 240118558. Retrieved 28 March 2023.
  162. ^ Zhang, Hua; Cao, Chang-qun; Liu, Xiao-lei; Mu, Lin; Zheng, Quan-feng; Liu, Feng; Xiang, Lei; Liu, Lu-jun; Shen, Shu-zhong (15 April 2016). "The terrestrial end-Permian mass extinction in South China". Palaeogeography, Palaeoclimatology, Palaeoecology. 448: 108–124. Bibcode:2016PPP...448..108Z. doi:10.1016/j.palaeo.2015.07.002. Retrieved 20 January 2023.
  163. ^ Chu, Daoliang; Yu, Jianxin; Tong, Jinnan; Benton, Michael James; Song, Haijun; Huang, Yunfei; Song, Ting; Tian, Li (November 2016). "Biostratigraphic correlation and mass extinction during the Permian-Triassic transition in terrestrial-marine siliciclastic settings of South China". Global and Planetary Change. 146: 67–88. Bibcode:2016GPC...146...67C. doi:10.1016/j.gloplacha.2016.09.009. hdl:1983/2d475883-42ea-4988-b01c-0a56912e6aec. Retrieved 12 November 2022.
  164. ^ a b c Feng, Zhuo; Wei, Hai-Bo; Guo, Yun; He, Xiao-Yuan; Sui, Qun; Zhou, Yu; Liu, Hang-Yu; Gou, Xu-Dong; Lv, Yong (May 2020). "From rainforest to herbland: New insights into land plant responses to the end-Permian mass extinction". Earth-Science Reviews. 204: 103153. Bibcode:2020ESRv..20403153F. doi:10.1016/j.earscirev.2020.103153. S2CID 216433847.
  165. ^ Biswas, Raman Kumar; Kaiho, Kunio; Saito, Ryosuke; Tian, Li; Shi, Zhiqiang (December 2020). "Terrestrial ecosystem collapse and soil erosion before the end-Permian marine extinction: Organic geochemical evidence from marine and non-marine records". Global and Planetary Change. 195: 103327. Bibcode:2020GPC...19503327B. doi:10.1016/j.gloplacha.2020.103327. S2CID 224900905. Retrieved 21 December 2022.
  166. ^ a b Chu, Daoliang; Grasby, Stephen E.; Song, Haijun; Dal Corso, Jacopo; Wang, Yao; Mather, Tamsin A.; Wu, Yuyang; Song, Huyue; Shu, Wenchao; Tong, Jinnan; Wignall, Paul Barry (3 January 2020). "Ecological disturbance in tropical peatlands prior to marine Permian-Triassic mass extinction". Geology. 48 (3): 288–292. Bibcode:2020Geo....48..288C. doi:10.1130/G46631.1. S2CID 214468383.
  167. ^ Blomenkemper, Patrick; Kerp, Hans; Abu Hamad, Abdalla; DiMichele, William A.; Bomfleur, Benjamin (21 December 2018). "A hidden cradle of plant evolution in Permian tropical lowlands". Science. 362 (6421): 1414–1416. Bibcode:2018Sci...362.1414B. doi:10.1126/science.aau4061. PMID 30573628. S2CID 56582195.
  168. ^ Viglietti, Pia A.; Benson, Roger B. J.; Smith, Roger M. H.; Botha, Jennifer; Kammerer, Christian F.; Skosan, Zaituna; Butler, Elize; Crean, Annelise; Eloff, Bobby; Kaal, Sheena; Mohoi, Joël; Molehe, William; Mtalana, Nolusindiso; Mtungata, Sibusiso; Ntheri, Nthaopa; Ntsala, Thabang; Nyaphuli, John; October, Paul; Skinner, Georgina; Strong, Mike; Stummer, Hedi; Wolvaart, Frederik P.; Angielczyk, Kenneth D. (27 April 2021). "Evidence from South Africa for a protracted end-Permian extinction on land". Proceedings of the National Academy of Sciences of the United States of America. 118 (17). doi:10.1073/pnas.2017045118. ISSN 0027-8424. PMC 8092562. PMID 33875588.
  169. ^ Smith, Roger M. H.; Ward, Peter D. (1 December 2001). "Pattern of vertebrate extinctions across an event bed at the Permian-Triassic boundary in the Karoo Basin of South Africa". Geology. 29 (12): 1147–1150. Bibcode:2001Geo....29.1147S. doi:10.1130/0091-7613(2001)029<1147:POVEAA>2.0.CO;2. Retrieved 23 May 2023.
  170. ^ Niedźwiedzki, Grzegorz; Bajdek, Piotr; Qvarnström, Martin; Sulej, Tomasz; Sennikov, Andrey G.; Golubev, Valeriy K. (15 May 2016). "Reduction of vertebrate coprolite diversity associated with the end-Permian extinction event in Vyazniki region, European Russia". Palaeogeography, Palaeoclimatology, Palaeoecology. 450: 77–90. doi:10.1016/j.palaeo.2016.02.057. Retrieved 18 June 2024 – via Elsevier Science Direct.
  171. ^ a b Smith, Roger M. H.; Botha-Brink, Jennifer (15 February 2014). "Anatomy of a mass extinction: Sedimentological and taphonomic evidence for drought-induced die-offs at the Permo-Triassic boundary in the main Karoo Basin, South Africa". Palaeogeography, Palaeoclimatology, Palaeoecology. 396: 99–118. Bibcode:2014PPP...396...99S. doi:10.1016/j.palaeo.2014.01.002. Retrieved 23 May 2023.
  172. ^ Gastaldo, Robert A.; Neveling, Johann (1 April 2016). "Comment on: "Anatomy of a mass extinction: Sedimentological and taphonomic evidence for drought-induced die-offs at the Permo–Triassic boundary in the main Karoo Basin, South Africa" by R.M.H. Smith and J. Botha-Brink, Palaeogeography, Palaeoclimatology, Palaeoecology 396:99-118". Palaeogeography, Palaeoclimatology, Palaeoecology. 447: 88–91. Bibcode:2016PPP...447...88G. doi:10.1016/j.palaeo.2014.06.027.
  173. ^ Maxwell, W.D. (1992). "Permian and Early Triassic extinction of non-marine tetrapods". Palaeontology. 35: 571–583.
  174. ^ Ketchum, Hilary F; Barrett, Paul M (21 January 2004). "New reptile material from the Lower Triassic of Madagascar: implications for the PermianTriassic extinction event". Canadian Journal of Earth Sciences. 41 (1): 1–8. doi:10.1139/e03-084. ISSN 0008-4077. Retrieved 13 October 2024 – via GeoScienceWorld.
  175. ^ Benoit, Julien; Kammerer, Christian F.; Dollman, Kathleen; Groenewald, David P.; Smith, Roger M.H. (15 March 2024). "Did gorgonopsians survive the end-Permian "Great Dying"? A re-appraisal of three gorgonopsian specimens (Therapsida, Theriodontia) reported from the Triassic Lystrosaurus declivis Assemblage Zone, Karoo Basin, South Africa". Palaeogeography, Palaeoclimatology, Palaeoecology. 638: 112044. doi:10.1016/j.palaeo.2024.112044. Retrieved 21 May 2024 – via Elsevier Science Direct.
  176. ^ Pitrat, Charles W. (December 1973). "Vertebrates and the Permo-Triassic extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 14 (4): 249–264. doi:10.1016/0031-0182(73)90011-4. Retrieved 13 October 2024 – via Elsevier Science Direct.
  177. ^ Bakaev, Aleksandr S. (17 July 2024). "Actinopterygians from the continental Permian–Triassic boundary section at Babiy Kamen (Kuznetsk Basin, Siberia, Russia)". Palaeoworld. doi:10.1016/j.palwor.2024.05.007. Retrieved 13 October 2024 – via Elsevier Science Direct.
  178. ^ "Bristol University – News – 2008: Mass extinction".
  179. ^ a b c Knoll AH, Bambach RK, Payne JL, Pruss S, Fischer WW (2007). "Paleophysiology and end-Permian mass extinction" (PDF). Earth and Planetary Science Letters. 256 (3–4): 295–313. Bibcode:2007E&PSL.256..295K. doi:10.1016/j.epsl.2007.02.018. Retrieved 13 December 2021.
  180. ^ Smith, Roger M. H.; Botha, Jennifer (September–October 2005). "The recovery of terrestrial vertebrate diversity in the South African Karoo Basin after the end-Permian extinction". Comptes Rendus Palevol. 4 (6–7): 623–636. Bibcode:2005CRPal...4..623S. doi:10.1016/j.crpv.2005.07.005. Retrieved 26 May 2023.
  181. ^ Botha, Jennifer; Smith, Roger M. H. (August 2006). "Rapid vertebrate recuperation in the Karoo Basin of South Africa following the End-Permian extinction". Journal of African Earth Sciences. 45 (4–5): 502–514. Bibcode:2006JAfES..45..502B. doi:10.1016/j.jafrearsci.2006.04.006. Retrieved 26 May 2023.
  182. ^ Sepkoski, J. John (8 February 2016). "A kinetic model of Phanerozoic taxonomic diversity. III. Post-Paleozoic families and mass extinctions". Paleobiology. 10 (2): 246–267. Bibcode:1984Pbio...10..246S. doi:10.1017/S0094837300008186. S2CID 85595559.
  183. ^ Clapham, Matthew E.; Bottjer, David J. (19 June 2007). "Permian marine paleoecology and its implications for large-scale decoupling of brachiopod and bivalve abundance and diversity during the Lopingian (Late Permian)". Palaeogeography, Palaeoclimatology, Palaeoecology. 249 (3–4): 283–301. Bibcode:2007PPP...249..283C. doi:10.1016/j.palaeo.2007.02.003. Retrieved 2 April 2023.
  184. ^ a b Clapham, Matthew E.; Bottjer, David J. (7 August 2007). "Prolonged Permian–Triassic ecological crisis recorded by molluscan dominance in Late Permian offshore assemblages". Proceedings of the National Academy of Sciences of the United States of America. 104 (32): 12971–12975. doi:10.1073/pnas.0705280104. PMC 1941817. PMID 17664426.
  185. ^ Romano, Carlo; Koot, Martha B.; Kogan, Ilja; Brayard, Arnaud; Minikh, Alla V.; Brinkmann, Winand; et al. (February 2016). "Permian–Triassic Osteichthyes (bony fishes): diversity dynamics and body size evolution". Biological Reviews. 91 (1): 106–147. doi:10.1111/brv.12161. PMID 25431138. S2CID 5332637.
  186. ^ Scheyer, Torsten M.; Romano, Carlo; Jenks, Jim; Bucher, Hugo (19 March 2014). "Early Triassic Marine Biotic Recovery: The Predators' Perspective". PLOS ONE. 9 (3): e88987. Bibcode:2014PLoSO...988987S. doi:10.1371/journal.pone.0088987. PMC 3960099. PMID 24647136.
  187. ^ Komatsu, Toshifumi; Huyen, Dang Tran; Jin-Hua, Chen (1 June 2008). "Lower Triassic bivalve assemblages after the end-Permian mass extinction in South China and North Vietnam". Paleontological Research. 12 (2): 119–128. doi:10.2517/1342-8144(2008)12[119:LTBAAT]2.0.CO;2. S2CID 131144468. Retrieved 6 November 2022.
  188. ^ Gould, S.J.; Calloway, C.B. (1980). "Clams and brachiopodsships that pass in the night". Paleobiology. 6 (4): 383–396. Bibcode:1980Pbio....6..383G. doi:10.1017/S0094837300003572. S2CID 132467749.
  189. ^ Jablonski, D. (8 May 2001). "Lessons from the past: Evolutionary impacts of mass extinctions". Proceedings of the National Academy of Sciences of the United States of America. 98 (10): 5393–5398. Bibcode:2001PNAS...98.5393J. doi:10.1073/pnas.101092598. PMC 33224. PMID 11344284.
  190. ^ Kaim, Andrzej; Nützel, Alexander (July 2011). "Dead bellerophontids walking – The short Mesozoic history of the Bellerophontoidea (Gastropoda)". Palaeogeography, Palaeoclimatology, Palaeoecology. 308 (1–2): 190–199. Bibcode:2011PPP...308..190K. doi:10.1016/j.palaeo.2010.04.008.
  191. ^ Hautmann, Michael (29 September 2009). "The first scallop" (PDF). Paläontologische Zeitschrift. 84 (2): 317–322. doi:10.1007/s12542-009-0041-5. S2CID 84457522.
  192. ^ Hautmann, Michael; Ware, David; Bucher, Hugo (August 2017). "Geologically oldest oysters were epizoans on Early Triassic ammonoids". Journal of Molluscan Studies. 83 (3): 253–260. doi:10.1093/mollus/eyx018.
  193. ^ Zhang, Shu-han; Shen, Shu-zhong; Erwin, Douglas H. (September 2022). "Two cosmopolitanism events driven by different extreme paleoclimate regimes". Global and Planetary Change. 216: 103899. Bibcode:2022GPC...21603899Z. doi:10.1016/j.gloplacha.2022.103899. S2CID 251174662. Retrieved 26 May 2023.
  194. ^ a b Petsios, Elizabeth; Bottjer, David P. (28 April 2016). "Quantitative analysis of the ecological dominance of benthic disaster taxa in the aftermath of the end-Permian mass extinction". Paleobiology. 42 (3): 380–393. Bibcode:2016Pbio...42..380P. doi:10.1017/pab.2015.47. S2CID 88450377. Retrieved 23 March 2023.
  195. ^ Yuan, Dong-Xun; Zhang, Yi-Chun; Shen, Shu-Zhong (1 February 2018). "Conodont succession and reassessment of major events around the Permian-Triassic boundary at the Selong Xishan section, southern Tibet, China". Global and Planetary Change. 161: 194–210. Bibcode:2018GPC...161..194Y. doi:10.1016/j.gloplacha.2017.12.024. ISSN 0921-8181. Retrieved 24 November 2023.
  196. ^ Groves, John R.; Rettori, Roberto; Payne, Jonathan L.; Boyce, Matthew D.; Altiner, Demír (14 July 2015). "End-Permian mass extinction of lagenide foraminifers in the Southern Alps (Northern Italy)". Journal of Paleontology. 81 (3): 415–434. doi:10.1666/05123.1. S2CID 32920409. Retrieved 23 March 2023.
  197. ^ Kolar-Jurkovšek, Tea; Hrvatović, Hazim; Aljinović, Dunja; Nestell, Galina P.; Jurkovšek, Bogdan; Skopljak, Ferid (1 May 2021). "Permian-Triassic biofacies of the Teočak section, Bosnia and Herzegovina". Global and Planetary Change. 200: 103458. Bibcode:2021GPC...20003458K. doi:10.1016/j.gloplacha.2021.103458. ISSN 0921-8181. S2CID 233858016. Retrieved 24 November 2023.
  198. ^ Schubert, Jennifer K.; Bottjer, David J. (June 1995). "Aftermath of the Permian-Triassic mass extinction event: Paleoecology of Lower Triassic carbonates in the western USA". Palaeogeography, Palaeoclimatology, Palaeoecology. 116 (1–2): 1–39. Bibcode:1995PPP...116....1S. doi:10.1016/0031-0182(94)00093-N. Retrieved 2 April 2023.
  199. ^ Song, Haijun; Huang, Shan; Jia, Enhao; Dai, Xu; Wignall, Paul Barry; Dunhill, Alexander M. (28 July 2020). "Flat latitudinal diversity gradient caused by the Permian–Triassic mass extinction". Proceedings of the National Academy of Sciences of the United States of America. 117 (30): 17578–17583. doi:10.1073/pnas.1918953117. ISSN 0027-8424. PMC 7395496. PMID 32631978.
  200. ^ "It took Earth ten million years to recover from greatest mass extinction". ScienceDaily. 27 May 2012. Retrieved 28 May 2012.
  201. ^ a b Brayard, Arnaud; Krumenacker, L. J.; Botting, Joseph P.; Jenks, James F.; Bylund, Kevin G.; Fara, Emmanuel; Vennin, Emmanuelle; Olivier, Nicolas; Goudemand, Nicolas; Saucède, Thomas; Charbonnier, Sylvain; Romano, Carlo; Doguzhaeva, Larisa; Thuy, Ben; Hautmann, Michael; Stephen, Daniel A.; Thomazo, Christophe; Escarguel, Gilles (15 February 2017). "Unexpected Early Triassic marine ecosystem and the rise of the Modern evolutionary fauna". Science Advances. 13 (2): e1602159. Bibcode:2017SciA....3E2159B. doi:10.1126/sciadv.1602159. PMC 5310825. PMID 28246643.
  202. ^ Smith, Christopher P. A.; Laville, Thomas; Fara, Emmanuel; Escarguel, Gilles; Olivier, Nicolas; Vennin, Emmanuelle; et al. (2021-10-04). "Exceptional fossil assemblages confirm the existence of complex Early Triassic ecosystems during the early Spathian". Scientific Reports. 11 (1): 19657. Bibcode:2021NatSR..1119657S. doi:10.1038/s41598-021-99056-8. ISSN 2045-2322. PMC 8490361. PMID 34608207.
  203. ^ Hofmann, Richard; Goudemand, Nicolas; Wasmer, Martin; Bucher, Hugo; Hautmann, Michael (1 October 2011). "New trace fossil evidence for an early recovery signal in the aftermath of the end-Permian mass extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 310 (3–4): 216–226. Bibcode:2011PPP...310..216H. doi:10.1016/j.palaeo.2011.07.014. Retrieved 20 December 2022.
  204. ^ Lewis, Dyani (9 February 2023). "Sea life bounced back fast after the 'mother of mass extinctions'". Nature. doi:10.1038/d41586-023-00383-9. PMID 36759740. S2CID 256738043. Retrieved 23 February 2023.
  205. ^ 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. Bibcode:2023Sci...379..567D. doi:10.1126/science.adf1622. PMID 36758082. S2CID 256697946.
  206. ^ Hautmann, Michael; Bucher, Hugo; Brühwiler, Thomas; Goudemand, Nicolas; Kaim, Andrzej; Nützel, Alexander (January–February 2011). "An unusually diverse mollusc fauna from the earliest Triassic of South China and its implications for benthic recovery after the end-Permian biotic crisis". Geobios. 44 (1): 71–85. Bibcode:2011Geobi..44...71H. doi:10.1016/j.geobios.2010.07.004. Retrieved 2 April 2023.
  207. ^ Zhou, C.Y.; Zhang, Q.Y.; Wen, W.; Huang, J.Y.; Hu, S.X.; Liu, W.; Min, X.; Ma, Z.X.; Wen, Q.Q. (2024). "A new Early Triassic fossil Lagerstätte from Wangmo, Guizhou Province". Sedimentary Geology and Tethyan Geology. 44 (1): 1–8. doi:10.19826/j.cnki.1009-3850.2022.06011.
  208. ^ Wheeley, James; Twitchett, Richard J. (2 January 2007). "Palaeoecological significance of a new Griesbachian (Early Triassic) gastropod assemblage from Oman". Lethaia. 38 (1): 37–45. doi:10.1080/0024116051003150. Retrieved 8 April 2023.
  209. ^ Pruss, Sara B.; Bottjer, David J. (1 December 2004). "Early Triassic Trace Fossils of the Western United States and their Implications for Prolonged Environmental Stress from the End-Permian Mass Extinction". PALAIOS. 19 (6): 551–564. doi:10.1669/0883-1351(2004)019<0551:ETTFOT>2.0.CO;2. ISSN 0883-1351. Retrieved 18 June 2024 – via GeoScienceWorld.
  210. ^ Flannery-Sutherland, Joseph T.; Silvestro, Daniele; Benton, Michael J. (18 May 2022). "Global diversity dynamics in the fossil record are regionally heterogeneous". Nature Communications. 13 (1): 2751. Bibcode:2022NatCo..13.2751F. doi:10.1038/s41467-022-30507-0. ISSN 2041-1723. PMC 9117201. PMID 35585069.
  211. ^ Woods, Adam D.; Bottjer, David J.; Mutti, Maria; Morrison, Jean (1 July 1999). "Lower Triassic large sea-floor carbonate cements: Their origin and a mechanism for the prolonged biotic recovery from the end-Permian mass extinction". Geology. 27 (7): 645–648. Bibcode:1999Geo....27..645W. doi:10.1130/0091-7613(1999)027<0645:LTLSFC>2.3.CO;2. ISSN 0091-7613. Retrieved 24 November 2023.
  212. ^ a b Woods, Adam D.; Alms, Paul D.; Monarrez, Pedro M.; Mata, Scott (1 January 2019). "The interaction of recovery and environmental conditions: An analysis of the outer shelf edge of western North America during the early Triassic". Palaeogeography, Palaeoclimatology, Palaeoecology. 513: 52–64. Bibcode:2019PPP...513...52W. doi:10.1016/j.palaeo.2018.05.014. Retrieved 20 December 2022.
  213. ^ Dineen, Ashley M.; Fraiser, Margaret L.; Sheehan, Peter M. (September 2014). "Quantifying functional diversity in pre- and post-extinction paleocommunities: A test of ecological restructuring after the end-Permian mass extinction". Earth-Science Reviews. 136: 339–349. Bibcode:2014ESRv..136..339D. doi:10.1016/j.earscirev.2014.06.002. Retrieved 2 April 2023.
  214. ^ Song, Haijun; Wignall, Paul Barry; Dunhill, Alexander M. (10 October 2018). "Decoupled taxonomic and ecological recoveries from the Permo-Triassic extinction". Science Advances. 4 (10): eaat5091. Bibcode:2018SciA....4.5091S. doi:10.1126/sciadv.aat5091. PMC 6179380. PMID 30324133.
  215. ^ Hofmann, Richard; Hautmann, Michael; Brayard, Arnaud; Nützel, Alexander; Bylund, Kevin G.; Jenks, James F.; et al. (May 2014). "Recovery of benthic marine communities from the end-Permian mass extinction at the low latitudes of eastern Panthalassa" (PDF). Palaeontology. 57 (3): 547–589. Bibcode:2014Palgy..57..547H. doi:10.1111/pala.12076. S2CID 6247479.
  216. ^ Feng, Xueqian; Chen, Zhong-Qiang; Benton, Michael James; Su, Chunmei; Bottjer, David J.; Cribb, Alison T.; Li, Ziheng; Zhao, Liashi; Zhu, Guangyou; Huang, Yuangeng; Guo, Zhen (29 June 2022). "Resilience of infaunal ecosystems during the Early Triassic greenhouse Earth". Science Advances. 8 (26): eabo0597. Bibcode:2022SciA....8O.597F. doi:10.1126/sciadv.abo0597. PMC 9242451. PMID 35767613. S2CID 250114146.
  217. ^ Brayard, A.; Escarguel, G.; Bucher, H.; Monnet, C.; Bruhwiler, T.; Goudemand, N.; et al. (27 August 2009). "Good genes and good luck: Ammonoid diversity and the end-Permian mass extinction". Science. 325 (5944): 1118–1121. Bibcode:2009Sci...325.1118B. doi:10.1126/science.1174638. PMID 19713525. S2CID 1287762.
  218. ^ Chen, Zhong-Qiang; Benton, Michael James (27 May 2012). "The timing and pattern of biotic recovery following the end-Permian mass extinction". Nature Geoscience. 5 (6): 375–383. Bibcode:2012NatGe...5..375C. doi:10.1038/ngeo1475. Retrieved 28 March 2023.
  219. ^ Hautmann, Michael; Bagherpour, Borhan; Brosse, Morgane; Frisk, Åsa; Hofmann, Richard; Baud, Aymon; et al. (September 2015). "Competition in slow motion: The unusual case of benthic marine communities in the wake of the end-Permian mass extinction". Palaeontology. 58 (5): 871–901. Bibcode:2015Palgy..58..871H. doi:10.1111/pala.12186. S2CID 140688908.
  220. ^ a b Grosse, Morgane; Bucher, Hugo; Baud, Aymon; Frisk, Åsa M.; Goudemand, Nicolas; Hagdorn, Hans; Nützel, Alexander; Ware, David; Hautmann, Michael (31 July 2018). "New data from Oman indicate benthic high biomass productivity coupled with low taxonomic diversity in the aftermath of the Permian–Triassic Boundary mass extinction". Lethaia. 52 (2): 165–187. doi:10.1111/let.12281. S2CID 135442906. Retrieved 8 April 2023.
  221. ^ Friesenbichler, Evelyn; Hautmann, Michael; Bucher, Hugo (20 July 2021). "The main stage of recovery after the end-Permian mass extinction: taxonomic rediversification and ecologic reorganization of marine level-bottom communities during the Middle Triassic". PeerJ. 9: e11654. doi:10.7717/peerj.11654. PMC 8300500. PMID 34322318.
  222. ^ Erwin, Douglas H. (January–September 2007). "Increasing returns, ecological feedback and the Early Triassic recovery". Palaeoworld. 16 (1–3): 9–15. doi:10.1016/j.palwor.2007.05.013. Retrieved 3 July 2023.
  223. ^ Wei, Hengye; Shen, Jun; Schoepfer, Shane D.; Krystyn, Leo; Richoz, Sylvain; Algeo, Thomas J. (October 2015). "Environmental controls on marine ecosystem recovery following mass extinctions, with an example from the Early Triassic". Earth-Science Reviews. 149: 108–135. doi:10.1016/j.earscirev.2014.10.007. Retrieved 18 June 2024 – via Elsevier Science Direct.
  224. ^ Stanley, Steven M. (8 September 2009). "Evidence from ammonoids and conodonts for multiple Early Triassic mass extinctions". Proceedings of the National Academy of Sciences of the United States of America. 106 (36): 15264–15267. Bibcode:2009PNAS..10615264S. doi:10.1073/pnas.0907992106. PMC 2741239. PMID 19721005.
  225. ^ Sun, Y. D.; Wignall, Paul Barry; Joachimski, Michael M.; Bond, David P. G.; Grasby, Stephen E.; Sun, S.; Yan, C. B.; Wang, L. N.; Chen, Y. L.; Lai, X. L. (1 June 2015). "High amplitude redox changes in the late Early Triassic of South China and the Smithian–Spathian extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 427: 62–78. Bibcode:2015PPP...427...62S. doi:10.1016/j.palaeo.2015.03.038. ISSN 0031-0182. Retrieved 24 November 2023.
  226. ^ Lloret, Joan; De la Horra, Raúl; Gretter, Nicola; Borruel-Abadía, Violeta; Barrenechea, José F.; Ronchi, Ausonio; Diez, José B.; Arche, Alfredo; López-Gómez, José (1 September 2020). "Gradual changes in the Olenekian-Anisian continental record and biotic implications in the Central-Eastern Pyrenean basin, NE Spain". Global and Planetary Change. 192: 103252. Bibcode:2020GPC...19203252L. doi:10.1016/j.gloplacha.2020.103252. ISSN 0921-8181. S2CID 225301237. Retrieved 24 November 2023.
  227. ^ Petsios, Elizabeth; Thompson, Jeffrey R.; Pietsch, Carlie; Bottjer, David J. (1 January 2019). "Biotic impacts of temperature before, during, and after the end-Permian extinction: A multi-metric and multi-scale approach to modeling extinction and recovery dynamics". Palaeogeography, Palaeoclimatology, Palaeoecology. 513: 86–99. Bibcode:2019PPP...513...86P. doi:10.1016/j.palaeo.2017.08.038. S2CID 134149088. Retrieved 2 December 2022.
  228. ^ a b Cao, Cheng; Bataille, Clément P.; Song, Haijun; Saltzman, Matthew R.; Cramer, Kate Tierney; Wu, Huaichun; Korte, Christoph; Zhang, Zhaofeng; Liu, Xiao-Ming (3 October 2022). "Persistent late Permian to Early Triassic warmth linked to enhanced reverse weathering". Nature Geoscience. 15 (1): 832–838. Bibcode:2022NatGe..15..832C. doi:10.1038/s41561-022-01009-x. S2CID 252708876. Retrieved 20 April 2023.
  229. ^ Wood, Rachel; Erwin, Douglas H. (16 October 2017). "Innovation not recovery: dynamic redox promotes metazoan radiations". Biology Reviews. 93 (2): 863–873. doi:10.1111/brv.12375. hdl:20.500.11820/a962bbdd-5d2a-45f9-8cbf-2fdd1d539ba1. PMID 29034568. S2CID 207103048. Retrieved 20 April 2023.
  230. ^ a b Wignall, P. B.; Twitchett, Richard J. (24 May 1996). "Oceanic Anoxia and the End Permian Mass Extinction". Science. 272 (5265): 1155–1158. Bibcode:1996Sci...272.1155W. doi:10.1126/science.272.5265.1155. PMID 8662450. S2CID 35032406.
  231. ^ Hofmann, Richard; Hautmann, Michael; Bucher, Hugo (October 2015). "Recovery dynamics of benthic marine communities from the Lower Triassic Werfen Formation, northern Italy". Lethaia. 48 (4): 474–496. Bibcode:2015Letha..48..474H. doi:10.1111/let.12121.
  232. ^ Foster, William J.; Danise, Silvia; Price, Gregory D.; Twitchett, Richard J. (15 March 2017). "Subsequent biotic crises delayed marine recovery following the late Permian mass extinction event in northern Italy". PLOS ONE. 12 (3): e0172321. Bibcode:2017PLoSO..1272321F. doi:10.1371/journal.pone.0172321. PMC 5351997. PMID 28296886.
  233. ^ Dineen, Ashley A.; Fraiser, Margaret L.; Tong, Jinnan (October 2015). "Low functional evenness in a post-extinction Anisian (Middle Triassic) paleocommunity: A case study of the Leidapo Member (Qingyan Formation), south China". Global and Planetary Change. 133: 79–86. Bibcode:2015GPC...133...79D. doi:10.1016/j.gloplacha.2015.08.001. Retrieved 2 April 2023.
  234. ^ Friesenbichler, Evelyn; Hautmann, Michael; Nützel, Alexander; Urlichs, Max; Bucher, Hugo (24 July 2018). "Palaeoecology of Late Ladinian (Middle Triassic) benthic faunas from the Schlern/Sciliar and Seiser Alm/Alpe di Siusi area (South Tyrol, Italy)" (PDF). PalZ. 93 (1): 1–29. doi:10.1007/s12542-018-0423-7. S2CID 134192673.
  235. ^ Friesenbichler, Evelyn; Hautmann, Michael; Grădinaru, Eugen; Bucher, Hugo; Brayard, Arnaud (12 October 2019). "A highly diverse bivalve fauna from a Bithynian (Anisian, Middle Triassic) – microbial buildup in North Dobrogea (Romania)" (PDF). Papers in Palaeontology. doi:10.1002/spp2.1286. S2CID 208555999.
  236. ^ Sepkoski, J. John (1997). "Biodiversity: Past, Present, and Future". Journal of Paleontology. 71 (4): 533–539. Bibcode:1997JPal...71..533S. doi:10.1017/S0022336000040026. PMID 11540302. S2CID 27430390.
  237. ^ a b Wagner PJ, Kosnik MA, Lidgard S (2006). "Abundance Distributions Imply Elevated Complexity of Post-Paleozoic Marine Ecosystems". Science. 314 (5803): 1289–1292. Bibcode:2006Sci...314.1289W. doi:10.1126/science.1133795. PMID 17124319. S2CID 26957610.
  238. ^ Salamon, Mariusz A.; Niedźwiedzki, Robert; Gorzelak, Przemysław; Lach, Rafał; Surmik, Dawid (1 March 2012). "Bromalites from the Middle Triassic of Poland and the rise of the Mesozoic Marine Revolution". Palaeogeography, Palaeoclimatology, Palaeoecology. 321–322: 142–150. Bibcode:2012PPP...321..142S. doi:10.1016/j.palaeo.2012.01.029. Retrieved 24 November 2023.
  239. ^ Nakajima, Yasuhisa; Izumi, Kentaro (15 November 2014). "Coprolites from the upper Osawa Formation (upper Spathian), northeastern Japan: Evidence for predation in a marine ecosystem 5Myr after the end-Permian mass extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 414: 225–232. doi:10.1016/j.palaeo.2014.08.014. Retrieved 18 June 2024 – via Elsevier Science Direct.
  240. ^ Wen, Wen; Zhang, Qiyue; Kriwet, Jürgen; Hu, Shixue; Zhou, Changyong; Huang, Jinyuan; Cui, Xindong; Min, Xiao; Benton, Michael James (1 May 2023). "First occurrence of hybodontid teeth in the Luoping Biota (Middle Triassic, Anisian) and recovery of the marine ecosystem after the end-Permian mass extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 617: 111471. doi:10.1016/j.palaeo.2023.111471.
  241. ^ Nakajima, Yasuhisa; Shigeta, Yasunari; Houssaye, Alexandra; Zakharov, Yuri D.; Popov, Alexander M.; Sander, P. Martin (1 April 2022). "Early Triassic ichthyopterygian fossils from the Russian Far East". Scientific Reports. 12 (1): 5546. doi:10.1038/s41598-022-09481-6. ISSN 2045-2322. PMC 8976075. PMID 35365703.
  242. ^ Song, Ting; Tong, Jinnan; Tian, Li; Chu, Daoliang; Huang, Yunfei (1 April 2019). "Taxonomic and ecological variations of Permian-Triassic transitional bivalve communities from the littoral clastic facies in southwestern China". Palaeogeography, Palaeoclimatology, Palaeoecology. 519: 108–123. Bibcode:2019PPP...519..108S. doi:10.1016/j.palaeo.2018.02.027. S2CID 134649495. Retrieved 2 April 2023.
  243. ^ Chen, Zhong-Qiang; Tong, Jinnan; Liao, Zhuo-Ting; Chen, Jing (August 2010). "Structural changes of marine communities over the Permian–Triassic transition: Ecologically assessing the end-Permian mass extinction and its aftermath". Global and Planetary Change. 73 (1–2): 123–140. Bibcode:2010GPC....73..123C. doi:10.1016/j.gloplacha.2010.03.011. Retrieved 6 November 2022.
  244. ^ Guo, Zhen; Flannery-Sutherland, Joseph T.; Benton, Michael J.; Chen, Zhong-Qiang (9 September 2023). "Bayesian analyses indicate bivalves did not drive the downfall of brachiopods following the Permian-Triassic mass extinction". Nature Communications. 14 (1): 5566. Bibcode:2023NatCo..14.5566G. doi:10.1038/s41467-023-41358-8. ISSN 2041-1723. PMC 10492784. PMID 37689772. Retrieved 24 November 2023.
  245. ^ Afanasjeva, G. A.; Viskova, L. A. (20 December 2021). "Morphophysiological Peculiarities of Articulated Brachiopods and Marine Bryozoans as a Reason for Their Different Evolutionary Consequences of the Permian–Triassic Crisis". Paleontological Journal. 55 (7): 742–751. doi:10.1134/S0031030121070029. ISSN 0031-0301. Retrieved 18 June 2024 – via Springer Link.
  246. ^ a b Greene, Sarah E.; Bottjer, David J.; Hagdorn, Hans; Zonneveld, John-Paul (15 July 2011). "The Mesozoic return of Paleozoic faunal constituents: A decoupling of taxonomic and ecological dominance during the recovery from the end-Permian mass extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 308 (1–2): 224–232. Bibcode:2011PPP...308..224G. doi:10.1016/j.palaeo.2010.08.019. Retrieved 2 April 2023.
  247. ^ Clapham ME, Bottjer DJ, Shen S (2006). "Decoupled diversity and ecology during the end-Guadalupian extinction (late Permian)". Geological Society of America Abstracts with Programs. 38 (7): 117. Archived from the original on 2015-12-08. Retrieved 28 March 2008.
  248. ^ McRoberts, Christopher A. (1 April 2001). "Triassic bivalves and the initial marine Mesozoic revolution: A role for predators?". Geology. 29 (4): 359. Bibcode:2001Geo....29..359M. doi:10.1130/0091-7613(2001)029<0359:TBATIM>2.0.CO;2. ISSN 0091-7613. Retrieved 20 September 2023.
  249. ^ Posenato, Renato; Holmer, Lars E.; Prinoth, Herwig (1 April 2014). "Adaptive strategies and environmental significance of lingulid brachiopods across the late Permian extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 399: 373–384. Bibcode:2014PPP...399..373P. doi:10.1016/j.palaeo.2014.01.028. Retrieved 14 January 2023.
  250. ^ Ke, Yan; Shen, Shu-zhong; Shi, Guang R.; Fan, Jun-xuan; Zhang, Hua; Qiao, Li; Zeng, Yong (15 April 2016). "Global brachiopod palaeobiogeographical evolution from Changhsingian (Late Permian) to Rhaetian (Late Triassic)". Palaeogeography, Palaeoclimatology, Palaeoecology. 448: 4–25. Bibcode:2016PPP...448....4K. doi:10.1016/j.palaeo.2015.09.049. Retrieved 3 July 2023.
  251. ^ Chen, Zhong-Qiang; Shi, Guang R.; Kaiho, Kunio (24 November 2003). "A New Genus of Rhynchonellid Brachiopod from the Lower Triassic of South China and Implications for Timing the Recovery of Brachiopoda After the End-Permian Mass Extinction". Palaeontology. 45 (1): 149–164. doi:10.1111/1475-4983.00231. S2CID 128441580.
  252. ^ Shen, Jing; Tong, Jinnan; Song, Haijun; Luo, Mao; Huang, Yunfei; Xiang, Ye (1 September 2015). "Recovery pattern of brachiopods after the Permian–Triassic crisis in South China". Palaeogeography, Palaeoclimatology, Palaeoecology. 433: 91–105. Bibcode:2015PPP...433...91C. doi:10.1016/j.palaeo.2015.05.020. Retrieved 26 June 2023.
  253. ^ Rong, Jia-yu; Shen, Shu-zhong (1 December 2002). "Comparative analysis of the end-Permian and end-Ordovician brachiopod mass extinctions and survivals in South China". Palaeogeography, Palaeoclimatology, Palaeoecology. 188 (1–2): 25–38. Bibcode:2002PPP...188...25R. doi:10.1016/S0031-0182(02)00507-2. Retrieved 18 April 2023.
  254. ^ Forel, Marie-Béatrice; Crasquin, Sylvie (6 September 2011). "Lower Triassic ostracods (Crustacea) from the Meishan section, Permian–Triassic boundary GSSP (Zhejiang Province, South China)". Journal of Systematic Palaeontology. 9 (3): 455–466. Bibcode:2011JSPal...9..455F. doi:10.1080/14772019.2010.526638. S2CID 128739624. Retrieved 23 May 2023.
  255. ^ Crasquin, Sylvie; Forel, Marie-Béatrice (October 2014). "Ostracods (Crustacea) through Permian–Triassic events". Earth-Science Reviews. 137: 52–64. Bibcode:2014ESRv..137...52C. doi:10.1016/j.earscirev.2013.01.006. Retrieved 23 May 2023.
  256. ^ Forel, Marie-Béatrice; Crasquin, Sylvie; Kershaw, S.; Feng, Q. L.; Collin, P.-Y. (13 July 2009). "Ostracods (Crustacea) and water oxygenation in the earliest Triassic of South China: implications for oceanic events at the end-Permian mass extinction". Australian Journal of Earth Sciences. 56 (6): 815–823. Bibcode:2009AuJES..56..815F. doi:10.1080/08120090903002631. S2CID 128700856. Retrieved 23 May 2023.
  257. ^ Powers, Catherine M.; Pachut, Joseph F. (20 May 2016). "Diversity and distribution of Triassic bryozoans in the aftermath of the end-Permian mass extinction". Journal of Paleontology. 82 (2): 362–371. doi:10.1666/06-131.1. ISSN 0022-3360. S2CID 130929034. Retrieved 30 December 2023.
  258. ^ Afanasjeva, G. A.; Viskova, L. A. (20 December 2021). "Morphophysiological Peculiarities of Articulated Brachiopods and Marine Bryozoans as a Reason for Their Different Evolutionary Consequences of the Permian–Triassic Crisis". Paleontological Journal. 55 (7): 742–751. doi:10.1134/S0031030121070029. ISSN 0031-0301. Retrieved 28 October 2024 – via Springer Link.
  259. ^ Foote, M. (1999). "Morphological diversity in the evolutionary radiation of Paleozoic and post-Paleozoic crinoids". Paleobiology. 25 (sp1): 1–116. doi:10.1666/0094-8373(1999)25[1:MDITER]2.0.CO;2. ISSN 0094-8373. JSTOR 2666042. S2CID 85586709.
  260. ^ Twitchett, Richard J.; Oji, Tatsuo (September–October 2005). "Early Triassic recovery of echinoderms". Comptes Rendus Palevol. 4 (6–7): 531–542. Bibcode:2005CRPal...4..531T. doi:10.1016/j.crpv.2005.02.006. Retrieved 2 April 2023.
  261. ^ Baumiller, T. K. (2008). "Crinoid Ecological Morphology". Annual Review of Earth and Planetary Sciences. 36 (1): 221–249. Bibcode:2008AREPS..36..221B. doi:10.1146/annurev.earth.36.031207.124116.
  262. ^ Thompson, Jeffrey R.; Hu, Shi-xue; Zhang, Qi-Yue; Petsios, Elizabeth; Cotton, Laura J.; Huang, Jin-Yuan; Zhou, Chang-yong; Wen, Wen; Bottjer, David J. (21 December 2017). "A new stem group echinoid from the Triassic of China leads to a revised macroevolutionary history of echinoids during the end-Permian mass extinction". Royal Society Open Science. 5 (1): 171548. doi:10.1098/rsos.171548. ISSN 2054-5703. PMC 5792935. PMID 29410858.
  263. ^ Thompson, Jeffrey R.; Posenato, Renato; Bottjer, David J.; Petsios, Elizabeth (30 August 2019). "Echinoids from the Tesero Member (Werfen Formation) of the Dolomites (Italy): implications for extinction and survival of echinoids in the aftermath of the end-Permian mass extinction". PeerJ. 7: e7361. doi:10.7717/peerj.7361. ISSN 2167-8359. PMC 6718154. PMID 31531267.
  264. ^ Metcalfe, Ian; Isozaki, Yukio (2 November 2009). "Current perspectives on the Permian–Triassic boundary and end-Permian mass extinction: Preface". Journal of Asian Earth Sciences. End-Permian Mass Extinction: Events & Processes, Age & Timescale, Causative mechanism(s), & Recovery. 36 (6): 407–412. Bibcode:2009JAESc..36..407M. doi:10.1016/j.jseaes.2009.07.009. ISSN 1367-9120. Retrieved 24 November 2023.
  265. ^ Martínez-Pérez, Carlos; Plasencia, Pablo; Cascales-Miñana, Borja; Mazza, Michele; Botella, Héctor (3 September 2014). "New insights into the diversity dynamics of Triassic conodonts". Historical Biology. 26 (5): 591–602. Bibcode:2014HBio...26..591M. doi:10.1080/08912963.2013.808632. hdl:2434/223492. ISSN 0891-2963. S2CID 53364944. Retrieved 24 November 2023.
  266. ^ Chen, Yanlong; Jiang, Haishui; Lai, Xulong; Yan, Chunbo; Richoz, Sylvain; Liu, Xiaodan; Wang, Lina (1 June 2015). "Early Triassic conodonts of Jiarong, Nanpanjiang Basin, southern Guizhou Province, South China". Journal of Asian Earth Sciences. 105: 104–121. Bibcode:2015JAESc.105..104C. doi:10.1016/j.jseaes.2015.03.014. ISSN 1367-9120. Retrieved 24 November 2023.
  267. ^ Kiliç, Ali Murat; Plasencia, Pablo; Ishida, Keisuke; Guex, Jean; Hirsch, Francis (1 January 2016). "Proteromorphosis of Neospathodus (Conodonta) during the Permian–Triassic crisis and recovery". Revue de Micropaléontologie. 59 (1): 33–39. Bibcode:2016RvMic..59...33K. doi:10.1016/j.revmic.2016.01.003. ISSN 0035-1598. Retrieved 24 November 2023.
  268. ^ Altıner, Demir; Payne, Jonathan L.; Lehrmann, Daniel J.; Özkan-Altıner, Sevinç; Kelley, Brian M.; Summers, Mindi M.; Yu, Meiyi (December 2021). "Triassic Foraminifera from the Great Bank of Guizhou, Nanpanjiang Basin, south China: taxonomic account, biostratigraphy, and implications for recovery from end-Permian mass extinction". Journal of Paleontology. 95 (S84): 1–53. doi:10.1017/jpa.2021.10. ISSN 0022-3360. Retrieved 12 May 2024 – via Cambridge Core.
  269. ^ Márquez, Leopoldo (12 December 2005). "Foraminiferal fauna recovered after the Late Permian extinctions in Iberia and the westernmost Tethys area". Palaeogeography, Palaeoclimatology, Palaeoecology. 229 (1–2): 137–157. doi:10.1016/j.palaeo.2005.06.035. Retrieved 28 October 2024 – via Elsevier Science Direct.
  270. ^ Brayard, Arnaud; Vennin, Emmanuelle; Olivier, Nicolas; Bylund, Kevin G.; Jenks, Jim; Stephen, Daniel A.; Bucher, Hugo; Hofmann, Richard; Goudemand, Nicolas; Escarguel, Gilles (18 September 2011). "Transient metazoan reefs in the aftermath of the end-Permian mass extinction". Nature Geoscience. 4 (1): 693–697. Bibcode:2011NatGe...4..693B. doi:10.1038/ngeo1264. Retrieved 8 May 2023.
  271. ^ باقرپور, برهان; سلیمانی, معصومه; فقیه, علی (February 2024). "Morphological diversity of microbialites and the significance of sponge remains in the Permian Triassic transition interval from Hambast Range, Central Iran". Advanced Applied Geology. 14 (2): 350–369. doi:10.22055/aag.2024.43965.2378. Retrieved 11 September 2024.
  272. ^ Kershaw, S.; Crasquin, S.; Li, Y.; Collin, P.-Y.; Forel, M.-B.; Mu, X.; Baud, A.; Wang, Y.; Xie, S.; Maurer, F.; Guo, L. (13 November 2011). "Microbialites and global environmental change across the Permian–Triassic boundary: a synthesis". Geobiology. 10 (1): 25–47. doi:10.1111/j.1472-4669.2011.00302.x. ISSN 1472-4677. PMID 22077322. Retrieved 1 August 2024 – via Wiley Online Library.
  273. ^ a b Wu, Siqi; Chen, Zhong-Qiang; Fang, Yuheng; Pei, Yu; Yang, Hao; Ogg, James (15 November 2017). "A Permian-Triassic boundary microbialite deposit from the eastern Yangtze Platform (Jiangxi Province, South China): Geobiologic features, ecosystem composition and redox conditions". Palaeogeography, Palaeoclimatology, Palaeoecology. 486: 58–73. Bibcode:2017PPP...486...58W. doi:10.1016/j.palaeo.2017.05.015. Retrieved 8 May 2023.
  274. ^ Zhang, Xi-Yang; Zheng, Quan-Feng; Li, Yue; Yang, Hong-Qiang; Zhang, Hua; Wang, Wen-Qian; Shen, Shu-Zhong (15 August 2020). "Polybessurus-like fossils as key contributors to Permian–Triassic boundary microbialites in South China". Palaeogeography, Palaeoclimatology, Palaeoecology. 552: 109770. Bibcode:2020PPP...55209770Z. doi:10.1016/j.palaeo.2020.109770. ISSN 0031-0182. S2CID 219090219. Retrieved 24 November 2023.
  275. ^ Friesenbichler, Evelyn; Rychoz, Sylvain; Baud, Aymon; Krystyn, Leopold; Sahakyan, Lilit; Vardanyan, Sargis; Peckmann, Jörn; Reitner, Joachim; Heindel, Katrin (15 January 2018). "Sponge-microbial build-ups from the lowermost Triassic Chanakhchi section in southern Armenia: Microfacies and stable carbon isotopes". Palaeogeography, Palaeoclimatology, Palaeoecology. 490: 653–672. Bibcode:2018PPP...490..653F. doi:10.1016/j.palaeo.2017.11.056. Retrieved 8 May 2023.
  276. ^ a b c Martindale, Rowan C.; Foster, William J.; Velledits, Felicitász (1 January 2019). "The survival, recovery, and diversification of metazoan reef ecosystems following the end-Permian mass extinction event". Palaeogeography, Palaeoclimatology, Palaeoecology. 513: 100–115. Bibcode:2019PPP...513..100M. doi:10.1016/j.palaeo.2017.08.014. S2CID 135338869. Retrieved 2 December 2022.
  277. ^ Wu, Siqi; Chen, Zhong-Qiang; Su, Chunmei; Fang, Yuheng; Yang, Hao (April 2022). "Keratose sponge fabrics from the lowermost Triassic microbialites in South China: Geobiologic features and Phanerozoic evolution". Global and Planetary Change. 211: 103787. Bibcode:2022GPC...21103787W. doi:10.1016/j.gloplacha.2022.103787. S2CID 247491746. Retrieved 8 May 2023.
  278. ^ Baud, Aymon; Richoz, Sylvain; Brandner, Rainier; Krystyn, Leopold; Hindel, Katrin; Mohdat, Tayebeh; Mohtat-Aghai, Parvin; Horacek, Micha (28 May 2021). "Sponge Takeover from End-Permian Mass Extinction to Early Induan Time: Records in Central Iran Microbial Buildups". Frontiers in Earth Science. 9: 355. Bibcode:2021FrEaS...9..355B. doi:10.3389/feart.2021.586210.
  279. ^ Wu, Siqi; Reitner, Joachim; Harper, David A. T.; Yu, Jianxin; Chen, Zhong-Qiang (27 December 2023). "New keratose sponges after the end-Permian extinction provide insights into biotic recoveries". Geobiology. 22 (1): e12582. doi:10.1111/gbi.12582. ISSN 1472-4677. PMID 38385600. Retrieved 11 September 2024 – via Wiley Online Library.
  280. ^ Senowbari-Daryan, Baba; Zühlke, Rainer; Bechstädt, Thilo; Flügel, Erik (December 1993). "Anisian (middle triassic) buildups of the Northern Dolomites (Italy): The recovery of reef communities after the permian/triassic crisis". Facies. 28 (1): 181–256. Bibcode:1993Faci...28..181S. doi:10.1007/BF02539736. ISSN 0172-9179. S2CID 129651368. Retrieved 20 September 2023.
  281. ^ Chen, Zhong-Qiang; Tu, Chenyi; Pei, Yu; Ogg, James; Fang, Yuheng; Wu, Siqi; Feng, Xueqian; Huang, Yuangeng; Guo, Zhen; Yang, Hao (February 2019). "Biosedimentological features of major microbe-metazoan transitions (MMTs) from Precambrian to Cenozoic". Earth-Science Reviews. 189: 21–50. Bibcode:2019ESRv..189...21C. doi:10.1016/j.earscirev.2019.01.015. S2CID 134828705. Retrieved 20 February 2023.
  282. ^ Kelley, Brian M.; Yu, Meiyi; Lehrmann, Daniel J.; Altıner, Demir; Payne, Jonathan L. (14 August 2023). "Prolonged and gradual recovery of metazoan-algal reefs following the end-Permian mass extinction". Geology. 51 (11): 1011–1016. doi:10.1130/G51058.1. ISSN 0091-7613. Retrieved 12 May 2024 – via GeoScienceWorld.
  283. ^ Xu, Yanling; Chen, Zhong-Qiang; Feng, Xueqian; Wu, Siqi; Shi, Guang R.; Tu, Chenyi (15 May 2017). "Proliferation of MISS-related microbial mats following the end-Permian mass extinction in the northern Paleo-Tethys: Evidence from southern Qilianshan region, western China". Palaeogeography, Palaeoclimatology, Palaeoecology. 474: 198–213. Bibcode:2017PPP...474..198X. doi:10.1016/j.palaeo.2016.04.045. Retrieved 8 May 2023.
  284. ^ Foster, William J.; Lehrmann, Daniel J.; Yu, Meiyi; Martindale, Rowan Clare (23 May 2019). "Facies selectivity of benthic invertebrates in a Permian/Triassic boundary microbialite succession: Implications for the "microbialite refuge" hypothesis". Geobiology. 17 (5): 523–535. doi:10.1111/gbi.12343. ISSN 1472-4677. PMID 31120196. Retrieved 18 June 2024 – via Wiley Online Library.
  285. ^ Zhao, Xiaoming; Tong, Jinnan; Yao, Huazhou; Niu, Zhijun; Luo, Mao; Huang, Yunfei; Song, Haijun (1 July 2015). "Early Triassic trace fossils from the Three Gorges area of South China: Implications for the recovery of benthic ecosystems following the Permian–Triassic extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 429: 100–116. Bibcode:2015PPP...429..100Z. doi:10.1016/j.palaeo.2015.04.008. Retrieved 20 January 2023.
  286. ^ Feng, Xueqian; Chen, Zhong-Qiang; Woods, Adam; Pei, Yu; Wu, Siqi; Fang, Yuheng; Luo, Mao; Xu, Yaling (October 2017). "Anisian (Middle Triassic) marine ichnocoenoses from the eastern and western margins of the Kamdian Continent, Yunnan Province, SW China: Implications for the Triassic biotic recovery". Global and Planetary Change. 157: 194–213. Bibcode:2017GPC...157..194F. doi:10.1016/j.gloplacha.2017.09.004. Retrieved 20 January 2023.
  287. ^ Zonneveld, John-Paul; Gingras, Murray; Beatty, Tyler W. (1 January 2010). "Diverse Ichnofossil Assemblages Following the P-T Mass Extinction, Lower Triassic, Alberta and British Columbia, Canada: Evidence for Shallow Marine Refugia on the Northwestern Coast of Pangaea". PALAIOS. 25 (6): 368–392. Bibcode:2010Palai..25..368Z. doi:10.2110/palo.2009.p09-135r. S2CID 128837115. Retrieved 20 January 2023.
  288. ^ Knaust, Dirk (11 May 2010). "The end-Permian mass extinction and its aftermath on an equatorial carbonate platform: insights from ichnology". Terra Nova. 22 (3): 195–202. Bibcode:2010TeNov..22..195K. doi:10.1111/j.1365-3121.2010.00934.x. S2CID 128722898. Retrieved 20 January 2023.
  289. ^ Cribb, Alison T.; Bottjer, David J. (14 January 2020). "Complex marine bioturbation ecosystem engineering behaviors persisted in the wake of the endPermian mass extinction". Scientific Reports. 10 (1): 203. Bibcode:2020NatSR..10..203C. doi:10.1038/s41598-019-56740-0. PMC 6959249. PMID 31937801. S2CID 210169203. Retrieved 8 April 2023.
  290. ^ Feng, Xueqian; Chen, Zhong-Qiang; Zhao, Laishi; Lan, Zhongwu (February 2021). "Middle Permian trace fossil assemblages from the Carnarvon Basin of Western Australia: Implications for the evolution of ichnofaunas in wave-dominated siliciclastic shoreface settings across the Permian-Triassic boundary". Global and Planetary Change. 197: 103392. Bibcode:2021GPC...19703392F. doi:10.1016/j.gloplacha.2020.103392. S2CID 229416157. Retrieved 20 January 2023.
  291. ^ Aftabuzzaman, Md.; Kaiho, Kunio; Biswas, Raman Kumar; Liu, Yuqing; Saito, Ryosuke; Tian, Li; Bhat, Ghulam M.; Chen, Zhong-Qiang (October 2021). "End-Permian terrestrial disturbance followed by the complete plant devastation, and the vegetation proto-recovery in the earliest-Triassic recorded in coastal sea sediments". Global and Planetary Change. 205: 103621. Bibcode:2021GPC...20503621A. doi:10.1016/j.gloplacha.2021.103621.
  292. ^ Yang, Wan; Wan, Mingli; Crowley, James L.; Wang, Jun; Luo, Xiaorong; Tabor, Neil; Angielczyk, Kenneth D.; Gastaldo, Robert; Geissman, John; Liu, Feng; Roopnarine, Peter; Sidor, Christian A. (November 2021). "Paleoenvironmental and paleoclimatic evolution and cyclo- and chrono-stratigraphy of upper Permian–Lower Triassic fluvial-lacustrine deposits in Bogda Mountains, NW China — Implications for diachronous plant evolution across the Permian–Triassic boundary". Earth-Science Reviews. 222: 103741. doi:10.1016/j.earscirev.2021.103741. Retrieved 13 October 2024 – via Elsevier Science Direct.
  293. ^ Hochuli, Peter A.; Sanson-Barrera, Anna; Schneebeli-Hermann, Elke; Bucher, Hugo (24 June 2016). "Severest crisis overlooked—Worst disruption of terrestrial environments postdates the Permian–Triassic mass extinction". Scientific Reports. 6 (1): 28372. Bibcode:2016NatSR...628372H. doi:10.1038/srep28372. PMC 4920029. PMID 27340926.
  294. ^ Looy, C. V.; Brugman, W. A.; Dilcher, D. L.; Visscher, H. (1999). "The delayed resurgence of equatorial forests after the Permian–Triassic ecologic crisis". Proceedings of the National Academy of Sciences of the United States of America. 96 (24): 13857–13862. Bibcode:1999PNAS...9613857L. doi:10.1073/pnas.96.24.13857. PMC 24155. PMID 10570163.
  295. ^ a b Hochuli, Peter A.; Hermann, Elke; Vigran, Jorunn Os; Bucher, Hugo; Weissert, Helmut (December 2010). "Rapid demise and recovery of plant ecosystems across the end-Permian extinction event". Global and Planetary Change. 74 (3–4): 144–155. Bibcode:2010GPC....74..144H. doi:10.1016/j.gloplacha.2010.10.004. Retrieved 24 December 2022.
  296. ^ Grauvogel-Stamm, Léa; Ash, Sidney R. (September–October 2005). "Recovery of the Triassic land flora from the end-Permian life crisis". Comptes Rendus Palevol. 4 (6–7): 593–608. Bibcode:2005CRPal...4..593G. doi:10.1016/j.crpv.2005.07.002. Retrieved 24 December 2022.
  297. ^ a b Xu, Zhen; Hilton, Jason; Yu, Jianxin; Wignall, Paul Barry; Yin, Hongfu; Xue, Qing; Ran, Weiju; Li, Hui; Shen, Jun; Meng, Fansong (September 2022). "End Permian to Middle Triassic plant species richness and abundance patterns in South China: Coevolution of plants and the environment through the Permian–Triassic transition". Earth-Science Reviews. 232: 104136. Bibcode:2022ESRv..23204136X. doi:10.1016/j.earscirev.2022.104136. S2CID 251031028. Archived from the original on 14 January 2023. Retrieved 26 June 2023.{{cite journal}}: CS1 maint: bot: original URL status unknown (link)
  298. ^ a b Chu, Daoliang; Dal Corso, Jacopo; Shu, Wenchao; Song, Haijun; Wignall, Paul Barry; Grasby, Stephen E.; Van de Schootbrugge, Bas; Zong, Keqing; Wu, Yuyang; Tong, Jinnan (5 February 2021). "Metal-induced stress in survivor plants following the end-Permian collapse of land ecosystems". Geology. 49 (6): 657–661. Bibcode:2021Geo....49..657C. doi:10.1130/G48333.1. S2CID 234074046.
  299. ^ Liu, Feng; Peng, Huiping; Bomfleur, Benjamin; Kerp, Hans; Zhu, Huaicheng; Shen, Shuzhong (October 2020). "Palynology and vegetation dynamics across the Permian–Triassic boundary in southern Tibet". Earth-Science Reviews. 209: 103278. Bibcode:2020ESRv..20903278L. doi:10.1016/j.earscirev.2020.103278. S2CID 225585090.
  300. ^ Hermann, Elke; Hochuli, Peter A.; Bucher, Hugo; Brühwiler, Thomas; Hautmann, Michael; Ware, David; Roohi, Ghazala (September 2011). "Terrestrial ecosystems on North Gondwana following the end-Permian mass extinction". Gondwana Research. 20 (2–3): 630–637. Bibcode:2011GondR..20..630H. doi:10.1016/j.gr.2011.01.008. Retrieved 31 May 2023.
  301. ^ Retallack, Gregory J. (1 January 1999). "Postapocalyptic greenhouse paleoclimate revealed by earliest Triassic paleosols in the Sydney Basin, Australia". Geological Society of America Bulletin. 111 (1): 52–70. doi:10.1130/0016-7606(1999)111<0052:PGPRBE>2.3.CO;2. Retrieved 31 May 2023.
  302. ^ Michaelsen, P. (2002). "Mass extinction of peat-forming plants and the effect on fluvial styles across the Permian–Triassic boundary, northern Bowen Basin, Australia". Palaeogeography, Palaeoclimatology, Palaeoecology. 179 (3–4): 173–188. Bibcode:2002PPP...179..173M. doi:10.1016/S0031-0182(01)00413-8.
  303. ^ Liu, Jun; Abdala, Fernando; Angielczyk, Kenneth D.; Sidor, Christian A. (January 2022). "Tetrapod turnover during the Permo-Triassic transition explained by temperature change". Earth-Science Reviews. 224: 103886. Bibcode:2022ESRv..22403886L. doi:10.1016/j.earscirev.2021.103886. S2CID 244900399. Retrieved 31 May 2023.
  304. ^ Liu, Jun; Angielczyk, Kenneth D.; Abdala, Fernando (October 2021). "Permo-Triassic tetrapods and their climate implications". Global and Planetary Change. 205: 103618. Bibcode:2021GPC...20503618L. doi:10.1016/j.gloplacha.2021.103618. Retrieved 10 August 2023.
  305. ^ a b Zhu, Zhicai; Liu, Yongqing; Kuang, Hongwei; Newell, Andrew J.; Peng, Nan; Cui, Mingming; Benton, Michael J. (September 2022). "Improving paleoenvironment in North China aided Triassic biotic recovery on land following the end-Permian mass extinction". Global and Planetary Change. 216: 103914. Bibcode:2022GPC...21603914Z. doi:10.1016/j.gloplacha.2022.103914.
  306. ^ a b Yu, Yingyue; Tian, Li; Chu, Daoliang; Song, Huyue; Guo, Wenwei; Tong, Jinnan (1 January 2022). "Latest Permian–Early Triassic paleoclimatic reconstruction by sedimentary and isotopic analyses of paleosols from the Shichuanhe section in central North China Basin". Palaeogeography, Palaeoclimatology, Palaeoecology. 585: 110726. Bibcode:2022PPP...58510726Y. doi:10.1016/j.palaeo.2021.110726. S2CID 239498183. Retrieved 6 November 2022.
  307. ^ Benton, Michael James; Tverdokhlebov, V. P.; Surkov, M. V. (4 November 2004). "Ecosystem remodelling among vertebrates at the Permian–Triassic boundary in Russia". Nature. 432 (7013): 97–100. Bibcode:2004Natur.432...97B. doi:10.1038/nature02950. PMID 15525988. S2CID 4388173. Retrieved 31 May 2023.
  308. ^ Yao, Mingtao; Sun, Zuoyu; Meng, Qingqiang; Li, Jiachun; Jiang, Dayong (15 August 2022). "Vertebrate coprolites from Middle Triassic Chang 7 Member in Ordos Basin, China: Palaeobiological and palaeoecological implications". Palaeogeography, Palaeoclimatology, Palaeoecology. 600: 111084. Bibcode:2022PPP...60011084M. doi:10.1016/j.palaeo.2022.111084. S2CID 249186414. Retrieved 8 January 2023.
  309. ^ Sidor, Christian A.; Vilhena, Daril A.; Angielczyk, Kenneth D.; Huttenlocker, Adam K.; Nisbett, Sterling J.; Peecook, Brandon R.; Steyer, J. Sébastien; Smith, Roger M. H.; Tsuji, Linda A. (29 April 2013). "Provincialization of terrestrial faunas following the end-Permian mass extinction". Proceedings of the National Academy of Sciences of the United States of America. 110 (20): 8129–8133. Bibcode:2013PNAS..110.8129S. doi:10.1073/pnas.1302323110. PMC 3657826. PMID 23630295.
  310. ^ Modesto, Sean P. (16 December 2020). "The Disaster Taxon Lystrosaurus: A Paleontological Myth". Frontiers in Earth Science. 8: 617. Bibcode:2020FrEaS...8..617M. doi:10.3389/feart.2020.610463.
  311. ^ Angielczyk, Kenneth D.; Liu, Jun; Sidor, Christian A.; Yang, Wan (November 2022). "The stratigraphic and geographic occurrences of Permo-Triassic tetrapods in the Bogda Mountains, NW China — Implications of a new cyclostratigraphic framework and Bayesian age model". Journal of African Earth Sciences. 195: 104655. doi:10.1016/j.jafrearsci.2022.104655.
  312. ^ Viglietti, Pia A.; Smith, Roger M. H.; Compton, John S. (15 December 2013). "Origin and palaeoenvironmental significance of Lystrosaurus bonebeds in the earliest Triassic Karoo Basin, South Africa". Palaeogeography, Palaeoclimatology, Palaeoecology. 392: 9–21. Bibcode:2013PPP...392....9V. doi:10.1016/j.palaeo.2013.08.015. Retrieved 31 May 2023.
  313. ^ Botha-Brink, Jennifer; Angielczyk, Kenneth D. (26 July 2010). "Do extraordinarily high growth rates in Permo-Triassic dicynodonts (Therapsida, Anomodontia) explain their success before and after the end-Permian extinction?". Zoological Journal of the Linnean Society. 160 (2): 341–365. doi:10.1111/j.1096-3642.2009.00601.x.
  314. ^ Botha-Brink, Jennifer; Codron, Daryl; Huttenlocker, Adam K.; Angielczyk, Kenneth D.; Ruta, Marcello (5 April 2016). "Breeding Young as a Survival Strategy during Earth's Greatest Mass Extinction". Scientific Reports. 6 (1): 24053. Bibcode:2016NatSR...624053B. doi:10.1038/srep24053. PMC 4820772. PMID 27044713.
  315. ^ Fröbisch, Jörg; Angielczyk, Kenneth D.; Sidor, Christian A. (3 December 2009). "The Triassic dicynodont Kombuisia (Synapsida, Anomodontia) from Antarctica, a refuge from the terrestrial Permian-Triassic mass extinction". Naturwissenschaften. 97 (2): 187–196. doi:10.1007/s00114-009-0626-6. PMID 19956920. S2CID 20557454. Retrieved 31 May 2023.
  316. ^ Marchetti, Lorenzo; Klein, Hendrik; Buchwitz, Michael; Ronchi, Ausonio; Smith, Roger M. H.; De Klerk, William J.; Sciscio, Lara; Groenewald, Gideon H. (August 2019). "Permian-Triassic vertebrate footprints from South Africa: Ichnotaxonomy, producers and biostratigraphy through two major faunal crises". Gondwana Research. 72: 139–168. Bibcode:2019GondR..72..139M. doi:10.1016/j.gr.2019.03.009. S2CID 133781923.
  317. ^ a b Huttenlocker, Adam K. (3 February 2014). "Body Size Reductions in Nonmammalian Eutheriodont Therapsids (Synapsida) during the End-Permian Mass Extinction". PLOS ONE. 9 (2): e87553. Bibcode:2014PLoSO...987553H. doi:10.1371/journal.pone.0087553. PMC 3911975. PMID 24498335.
  318. ^ Rey, Kévin; Amiot, Romain; Fourel, François; Abdala, Fernando; Fluteau, Frédéric; Jalil, Nour-Eddine; Liu, Jun; Rubidge, Bruce S.; Smith, Roger M. H.; Steyer, J. Sébastien; Viglietti, Pia A.; Wang, Xu; Lécuyer, Christophe (18 July 2017). "Oxygen isotopes suggest elevated thermometabolism within multiple Permo-Triassic therapsid clades". eLife. 6: e28589. doi:10.7554/eLife.28589. PMC 5515572. PMID 28716184.
  319. ^ Fontanarrosa, Gabriela; Abdala, Fernando; Kümmell, Susanna; Gess, Robert (26 March 2019). "The manus of Tetracynodon (Therapsida: Therocephalia) provides evidence for survival strategies following the Permo-Triassic extinction". Journal of Vertebrate Paleontology. 38 (4): (1)-(13). doi:10.1080/02724634.2018.1491404. hdl:11336/91246. S2CID 109228166. Retrieved 31 May 2023.
  320. ^ Botha, J. & Smith, R.M.H. (2007). "Lystrosaurus species composition across the Permo–Triassic boundary in the Karoo Basin of South Africa" (PDF). Lethaia. 40 (2): 125–137. Bibcode:2007Letha..40..125B. doi:10.1111/j.1502-3931.2007.00011.x. Archived from the original (PDF) on 2008-09-10. Retrieved 2008-07-02.
  321. ^ Huttenlocker, Adam K.; Sidor, Christian A.; Smith, Roger M. H. (21 March 2011). "A new specimen of Promoschorhynchus (Therapsida: Therocephalia: Akidnognathidae) from the Lower Triassic of South Africa and its implications for theriodont survivorship across the Permo-Triassic boundary". Journal of Vertebrate Paleontology. 31 (2): 405–421. Bibcode:2011JVPal..31..405H. doi:10.1080/02724634.2011.546720. S2CID 129242450. Retrieved 31 May 2023.
  322. ^ Huttenlocker, Adam K.; Botha-Brink, Jennifer (8 April 2014). "Bone microstructure and the evolution of growth patterns in Permo-Triassic therocephalians (Amniota, Therapsida) of South Africa". PeerJ. 2: e325. doi:10.7717/peerj.325. PMC 3994631. PMID 24765566.
  323. ^ Hoffmann, Devin K.; Hancox, John P.; Nesbitt, Sterling J. (1 May 2023). "A diverse diapsid tooth assemblage from the Early Triassic (Driefontein locality, South Africa) records the recovery of diapsids following the end-Permian mass extinction". PLOS ONE. 18 (5): e0285111. Bibcode:2023PLoSO..1885111H. doi:10.1371/journal.pone.0285111. PMC 10150976. PMID 37126508.
  324. ^ Benton, M.J. (2004). Vertebrate Paleontology. Blackwell Publishers. xii–452. ISBN 978-0-632-05614-9.
  325. ^ Ruben, J.A. & Jones, T.D. (2000). "Selective Factors Associated with the Origin of Fur and Feathers". American Zoologist. 40 (4): 585–596. doi:10.1093/icb/40.4.585.
  326. ^ Benton, Michael James (December 2021). "The origin of endothermy in synapsids and archosaurs and arms races in the Triassic". Gondwana Research. 100: 261–289. Bibcode:2021GondR.100..261B. doi:10.1016/j.gr.2020.08.003. S2CID 222247711.
  327. ^ McHugh, Julia Beth (May 2012). "ASSESSING TEMNOSPONDYL EVOLUTION AND ITS IMPLICATIONS FOR THE TERRESTRIAL PERMO-TRIASSIC MASS EXTINCTION". Temnospondyl ontogeny and phylogeny, a window into terrestrial ecosystems during the Permian-Triassic mass extinction (PhD). University of Iowa. ProQuest 1030963218. Retrieved 20 September 2023.
  328. ^ a b Ruta, Marcello; Benton, Michael J. (19 November 2011). "Calibrated Diversity, Tree Topology and the Mother of Mass Extinctions: The Lesson of Temnospondyls". Palaeontology. 51 (6): 1261–1288. doi:10.1111/j.1475-4983.2008.00808.x. S2CID 85411546.
  329. ^ Scholze, Frank; Golubev, Valeriy K.; Niedźwiedzki, Grzegorz; Sennikov, Andrey G.; Schneider, Jörg W.; Silantiev, Vladimir V. (1 July 2015). "Early Triassic Conchostracans (Crustacea: Branchiopoda) from the terrestrial Permian–Triassic boundary sections in the Moscow syncline". Palaeogeography, Palaeoclimatology, Palaeoecology. 429: 22–40. Bibcode:2015PPP...429...22S. doi:10.1016/j.palaeo.2015.04.002. ISSN 0031-0182. Retrieved 24 November 2023.
  330. ^ Tarailo, David A. (5 November 2018). "Taxonomic and ecomorphological diversity of temnospondyl amphibians across the Permian–Triassic boundary in the Karoo Basin (South Africa)". Journal of Morphology. 279 (12): 1840–1848. doi:10.1002/jmor.20906. PMID 30397933. S2CID 53234826. Retrieved 31 May 2023.
  331. ^ Yates, A. M.; Warren, A. A. (2000). "The phylogeny of the 'higher' temnospondyls (Vertebrata: Choanata) and its implications for the monophyly and origins of the Stereospondyli". Zoological Journal of the Linnean Society. 128 (1): 77–121. doi:10.1111/j.1096-3642.2000.tb00650.x.
  332. ^ Zheng, Wei; Wan, En-zhao; Xu, Xin; Li, Da; Dai, Ming-yue; Qi, Yong-An; Xing, Zhi-Feng; Liu, Yun-Long (March 2023). "The variations of terrestrial trace fossils and sedimentary substrates after the end-Permian extinction in the Dengfeng area, North China". Geological Journal. 58 (3): 1223–1238. Bibcode:2023GeolJ..58.1223Z. doi:10.1002/gj.4657. S2CID 254207982. Retrieved 8 April 2023.
  333. ^ a b Chu, Daoliang; Tong, Jinnan; Bottjer, David J.; Song, Haijun; Song, Huyue; Benton, Michael James; Tian, Li; Guo, Wenwei (15 May 2017). "Microbial mats in the terrestrial Lower Triassic of North China and implications for the Permian–Triassic mass extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 474: 214–231. Bibcode:2017PPP...474..214C. doi:10.1016/j.palaeo.2016.06.013. hdl:1983/95966174-157e-4814-b73f-6901ff9b9bf8. Retrieved 23 December 2022.
  334. ^ Wignall, Paul Barry; Bond, David P. G.; Grasby, Stephen E.; Pruss, Sarah B.; Peakall, Jeffrey (30 August 2019). "Controls on the formation of microbially induced sedimentary structures and biotic recovery in the Lower Triassic of Arctic Canada". Geological Society of America Bulletin. 132 (5–6): 918–930. doi:10.1130/B35229.1. S2CID 202194000.
  335. ^ Andy Saunders; Marc Reichow (2009). "The Siberian Traps – area and volume". Retrieved 2009-10-18.
  336. ^ "Largest Countries in the World (by area)". 2023. Retrieved 2023-05-25.
  337. ^ Saunders, Andy & Reichow, Marc (January 2009). "The Siberian Traps and the End-Permian mass extinction: a critical review" (PDF). Chinese Science Bulletin. 54 (1): 20–37. Bibcode:2009ChSBu..54...20S. doi:10.1007/s11434-008-0543-7. hdl:2381/27540. S2CID 1736350.
  338. ^ Reichow, Marc K.; Pringle, M.S.; Al'Mukhamedov, A.I.; Allen, M.B.; Andreichev, V.L.; Buslov, M.M.; et al. (2009). "The timing and extent of the eruption of the Siberian Traps large igneous province: Implications for the end-Permian environmental crisis" (PDF). Earth and Planetary Science Letters. 277 (1–2): 9–20. Bibcode:2009E&PSL.277....9R. doi:10.1016/j.epsl.2008.09.030. hdl:2381/4204.
  339. ^ Augland, L. E.; Ryabov, V. V.; Vernikovsky, V. A.; Planke, S.; Polozov, A. G.; Callegaro, S.; Jerram, D. A.; Svensen, H. H. (10 December 2019). "The main pulse of the Siberian Traps expanded in size and composition". Scientific Reports. 9 (1): 18723. Bibcode:2019NatSR...918723A. doi:10.1038/s41598-019-54023-2. PMC 6904769. PMID 31822688.
  340. ^ Kamo, SL (2003). "Rapid eruption of Siberian flood-volcanic rocks and evidence for coincidence with the Permian–Triassic boundary and mass extinction at 251 Ma". Earth and Planetary Science Letters. 214 (1–2): 75–91. Bibcode:2003E&PSL.214...75K. doi:10.1016/S0012-821X(03)00347-9.
  341. ^ Black, Benjamin A.; Weiss, Benjamin P.; Elkins-Tanton, Linda T.; Veselovskiy, Roman V.; Latyshev, Anton (30 April 2015). "Siberian Traps volcaniclastic rocks and the role of magma-water interactions". Geological Society of America Bulletin. 127 (9–10): B31108.1. Bibcode:2015GSAB..127.1437B. doi:10.1130/B31108.1. ISSN 0016-7606.
  342. ^ Fischman, Josh. Giant eruptions and giant extinctions. Scientific American (video). Retrieved 2016-03-11.
  343. ^ Pavlov, Vladimir E.; Fluteau, Frederic; Latyshev, Anton V.; Fetisova, Anna M.; Elkins-Tanton, Linda T.; Black, Ben A.; Burgess, Seth D.; Veselovskiy, Roman V. (17 January 2019). "Geomagnetic Secular Variations at the Permian-Triassic Boundary and Pulsed Magmatism During Eruption of the Siberian Traps". Geochemistry, Geophysics, Geosystems. 20 (2): 773–791. Bibcode:2019GGG....20..773P. doi:10.1029/2018GC007950. S2CID 134521010. Retrieved 20 February 2023.
  344. ^ a b c Jiang, Qiang; Jourdan, Fred; Olierook, Hugo K. H.; Merle, Renaud E.; Bourdet, Julien; Fougerouse, Denis; Godel, Belinda; Walker, Alex T. (25 July 2022). "Volume and rate of volcanic CO2 emissions governed the severity of past environmental crises". Proceedings of the National Academy of Sciences of the United States of America. 119 (31): e2202039119. Bibcode:2022PNAS..11902039J. doi:10.1073/pnas.2202039119. PMC 9351498. PMID 35878029.
  345. ^ Wang, Wen-qian; Zheng, Feifei; Zhang, Shuang; Cui, Ying; Zheng, Quan-feng; Zhang, Yi-chun; Chang, Dong-xun; Zhang, Hua; Xu, Yi-gang; Shen, Shu-zhong (15 January 2023). "Ecosystem responses of two Permian biocrises modulated by CO2 emission rates". Earth and Planetary Science Letters. 602: 117940. Bibcode:2023E&PSL.60217940W. doi:10.1016/j.epsl.2022.117940. S2CID 254660567.
  346. ^ Wignall, Paul Barry; Sun, Yadong; Bond, David P. G.; Izon, Gareth; Newton, Robert J.; Védrine, Stéphanie; Widdowson, Mike; Ali, Jason R.; Lai, Xulong; Jiang, Haishui; Cope, Helen; Bottrell, Simon H. (2009). "Volcanism, Mass Extinction, and Carbon Isotope Fluctuations in the Middle Permian of China". Science. 324 (5931): 1179–1182. Bibcode:2009Sci...324.1179W. doi:10.1126/science.1171956. PMID 19478179. S2CID 206519019. Retrieved 20 February 2023.
  347. ^ Jerram, Dougal A.; Widdowson, Mike; Wignall, Paul Barry; Sun, Yadong; Lai, Xulong; Bond, David P. G.; Torsvik, Trond H. (1 January 2016). "Submarine palaeoenvironments during Emeishan flood basalt volcanism, SW China: Implications for plume–lithosphere interaction during the Capitanian, Middle Permian ('end Guadalupian') extinction event". Palaeogeography, Palaeoclimatology, Palaeoecology. 441: 65–73. Bibcode:2016PPP...441...65J. doi:10.1016/j.palaeo.2015.06.009. Retrieved 19 December 2022.
  348. ^ Zhou, Mei-Fu; Malpas, John; Song, Xie-Yan; Robinson, Paul T.; Sun, Min; Kennedy, Allen K.; Lesher, C. Michael; Keays, Ride R. (15 March 2002). "A temporal link between the Emeishan large igneous province (SW China) and the end-Guadalupian mass extinction". Earth and Planetary Science Letters. 196 (3–4): 113–122. Bibcode:2002E&PSL.196..113Z. doi:10.1016/S0012-821X(01)00608-2. Retrieved 20 February 2023.
  349. ^ a b c d e f White, R. V. (2002). "Earth's biggest 'whodunnit': Unravelling the clues in the case of the end-Permian mass extinction" (PDF). Philosophical Transactions of the Royal Society of London. 360 (1801): 2963–2985. Bibcode:2002RSPTA.360.2963W. doi:10.1098/rsta.2002.1097. PMID 12626276. S2CID 18078072. Archived from the original (PDF) on 2020-11-11. Retrieved 2008-01-12.
  350. ^ Joachimski, Michael M.; Alekseev, A. S.; Grigoryan, A.; Gatovsky, Yu. A. (17 June 2019). "Siberian Trap volcanism, global warming and the Permian-Triassic mass extinction: New insights from Armenian Permian-Triassic sections". Geological Society of America Bulletin. 132 (1–2): 427–443. doi:10.1130/B35108.1. S2CID 197561486. Retrieved 26 May 2023.
  351. ^ Kenny, Ray (16 January 2018). "A geochemical view into continental palaeotemperatures of the end-Permian using oxygen and hydrogen isotope composition of secondary silica in chert rubble breccia: Kaibab Formation, Grand Canyon (USA)". Geochemical Transactions. 19 (2): 2. Bibcode:2018GeoTr..19....2K. doi:10.1186/s12932-017-0047-y. PMC 5770344. PMID 29340852.
  352. ^ Schobben, Martin; Joachimski, Michael M.; Korn, Dieter; Leda, Lucyna; Korte, Christoph (September 2014). "Palaeotethys seawater temperature rise and an intensified hydrological cycle following the end-Permian mass extinction". Gondwana Research. 26 (2): 675–683. Bibcode:2014GondR..26..675S. doi:10.1016/j.gr.2013.07.019. Retrieved 31 May 2023.
  353. ^ Kump, Lee R. (3 September 2018). "Prolonged Late Permian–Early Triassic hyperthermal: failure of climate regulation?". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 376 (2130): 1–9. Bibcode:2018RSPTA.37670078K. doi:10.1098/rsta.2017.0078. PMC 6127386. PMID 30177562. S2CID 52152614.
  354. ^ Zhang, Hongrui; Torsvik, Trond H. (15 April 2022). "Circum-Tethyan magmatic provinces, shifting continents and Permian climate change". Earth and Planetary Science Letters. 584: 117453. Bibcode:2022E&PSL.58417453Z. doi:10.1016/j.epsl.2022.117453. S2CID 247298020. Retrieved 31 May 2023.
  355. ^ "Driver of the largest mass extinction in the history of the Earth identified". phys.org. Retrieved 8 November 2020.
  356. ^ "Large volcanic eruption caused the largest mass extinction". phys.org. Retrieved 8 December 2020.
  357. ^ Shen, Jun; Yu, Jianxin; Chen, Jiubin; Algeo, Thomas J.; Xu, Guozhen; Feng, Qinglai; Shi, Xiao; Planavsky, Noah J.; Shu, Wenchao; Xie, Shucheng (25 September 2019). "Mercury evidence of intense volcanic effects on land during the Permian-Triassic transition". Geology. 47 (12): 1117–1121. Bibcode:2019Geo....47.1117S. doi:10.1130/G46679.1. S2CID 204262451. Retrieved 26 May 2023.
  358. ^ Regelous, Marcel; Regelous, Anette; Grasby, Stephen E.; Bond, David P. G.; Haase, Karsten M.; Gleißner, Stefan; Wignall, Paul Barry (31 October 2020). "Tellurium in Late Permian-Early Triassic Sediments as a Proxy for Siberian Flood Basalt Volcanism". Geochemistry, Geophysics, Geosystems. 21 (11). doi:10.1029/2020GC009064. ISSN 1525-2027. Retrieved 13 March 2024.
  359. ^ Liu, Sheng-Ao; Wu, Huaichun; Shen, Shu-zhong; Jiang, Ganqing; Zhang, Shihong; Lv, Yiwen; Zhang, Hua; Li, Shuguang (1 April 2017). "Zinc isotope evidence for intensive magmatism immediately before the end-Permian mass extinction". Geology. 45 (4): 343–346. Bibcode:2017Geo....45..343L. doi:10.1130/G38644.1. Retrieved 28 March 2023.
  360. ^ a b c Broadley, Michael W.; Barry, Peter H.; Ballentine, Chris J.; Taylor, Lawrence A.; Burgess, Ray (27 August 2018). "End-Permian extinction amplified by plume-induced release of recycled lithospheric volatiles". Nature Geoscience. 11 (9): 682–687. Bibcode:2018NatGe..11..682B. doi:10.1038/s41561-018-0215-4. S2CID 133833819. Retrieved 28 March 2023.
  361. ^ Sobolev, Stephan V.; Sobolev, Alexander V.; Kuzmin, Dmitry V.; Krivolutskaya, Nadezhda A.; Petrunin, Alexey G.; Arndt, Nicholas T.; Radko, Viktor A.; Vasiliev, Yuri R. (14 September 2011). "Linking mantle plumes, large igneous provinces and environmental catastrophes". Nature. 477 (7364): 312–316. Bibcode:2011Natur.477..312S. doi:10.1038/nature10385. ISSN 0028-0836. PMID 21921914. S2CID 205226146. Retrieved 20 September 2023.
  362. ^ Brand, Uwe; Posenato, Renato; Came, Rosemarie; Affek, Hagit; Angiolini, Lucia; Azmy, Karem; Farabegoli, Enzo (5 September 2012). "The end-Permian mass extinction: A rapid volcanic CO2 and CH4-climatic catastrophe". Chemical Geology. 322–323: 121–144. Bibcode:2012ChGeo.322..121B. doi:10.1016/j.chemgeo.2012.06.015. Retrieved 5 March 2023.
  363. ^ Baresel, Björn; Bucher, Hugo; Bagherpour, Borhan; Brosse, Morgane; Guodun, Kuang; Schaltegger, Urs (6 March 2017). "Timing of global regression and microbial bloom linked with the Permian-Triassic boundary mass extinction: implications for driving mechanisms". Scientific Reports. 7: 43630. Bibcode:2017NatSR...743630B. doi:10.1038/srep43630. PMC 5338007. PMID 28262815.
  364. ^ Wignall, Paul Barry; Bond, David P. G. (25 October 2023). "The great catastrophe: causes of the Permo-Triassic marine mass extinction". National Science Review. 11 (1): nwad273. doi:10.1093/nsr/nwad273. ISSN 2095-5138. PMC 10753410. PMID 38156041.
  365. ^ Maruoka, T.; Koeberl, C.; Hancox, P. J.; Reimold, W. U. (30 January 2003). "Sulfur geochemistry across a terrestrial Permian–Triassic boundary section in the Karoo Basin, South Africa". Earth and Planetary Science Letters. 206 (1–2): 101–117. Bibcode:2003E&PSL.206..101M. doi:10.1016/S0012-821X(02)01087-7. Retrieved 31 May 2023.
  366. ^ "Volcanism". Hooper Museum. hoopermuseum.earthsci.carleton.ca. Ottawa, Ontario, Canada: Carleton University.
  367. ^ Elkins-Tanton, L. T.; Grasby, Stephen E.; Black, B. A.; Veselovskiy, R. V.; Ardakani, O. H.; Goodarzi, F. (12 June 2020). "Field evidence for coal combustion links the 252 Ma Siberian Traps with global carbon disruption". Geology. 48 (10): 986–991. Bibcode:2020Geo....48..986E. doi:10.1130/G47365.1.
  368. ^ Svensen, Henrik; Planke, Sverre; Polozov, Alexander G.; Schmidbauer, Norbert; Corfu, Fernando; Podladchikov, Yuri Y.; Jamtveit, Bjørn (30 January 2009). "Siberian gas venting and the end-Permian environmental crisis". Earth and Planetary Science Letters. 297 (3–4): 490–500. Bibcode:2009E&PSL.277..490S. doi:10.1016/j.epsl.2008.11.015. Retrieved 13 January 2023.
  369. ^ Konstantinov, Konstantin M.; Bazhenov, Mikhail L.; Fetisova, Anna M.; Khutorskoy, Mikhail D. (May 2014). "Paleomagnetism of trap intrusions, East Siberia: Implications to flood basalt emplacement and the Permo–Triassic crisis of biosphere". Earth and Planetary Science Letters. 394: 242–253. Bibcode:2014E&PSL.394..242K. doi:10.1016/j.epsl.2014.03.029.
  370. ^ Chen, Chengsheng; Qin, Shengfei; Wang, Yunpeng; Holland, Greg; Wynn, Peter; Zhong, Wanxu; Zhou, Zheng (12 November 2022). "High temperature methane emissions from Large Igneous Provinces as contributors to late Permian mass extinctions". Nature Communications. 13 (1): 6893. Bibcode:2022NatCo..13.6893C. doi:10.1038/s41467-022-34645-3. ISSN 2041-1723. PMC 9653473. PMID 36371500.
  371. ^ Dal Corso, Jacopo; Mills, Benjamin J. W.; Chu, Daoling; Newton, Robert J.; Mather, Tamsin A.; Shu, Wenchao; Wu, Yuyang; Tong, Jinnan; Wignall, Paul Barry (11 June 2020). "Permo–Triassic boundary carbon and mercury cycling linked to terrestrial ecosystem collapse". Nature Communications. 11 (1): 2962. Bibcode:2020NatCo..11.2962D. doi:10.1038/s41467-020-16725-4. PMC 7289894. PMID 32528009.
  372. ^ Kaiho, Kunio (30 April 2024). "Role of volcanism and impact heating in mass extinction climate shifts". Scientific Reports. 14 (1): 9946. doi:10.1038/s41598-024-60467-y. ISSN 2045-2322. PMC 11061309. PMID 38688982.
  373. ^ Verango, Dan (24 January 2011). "Ancient mass extinction tied to torched coal". USA Today.
  374. ^ Grasby, Stephen E.; Sanei, Hamed & Beauchamp, Benoit (January 23, 2011). "Catastrophic dispersion of coal fly ash into oceans during the latest Permian extinction". Nature Geoscience. 4 (2): 104–107. Bibcode:2011NatGe...4..104G. doi:10.1038/ngeo1069.
  375. ^ "Researchers find smoking gun of world's biggest extinction: Massive volcanic eruption, burning coal and accelerated greenhouse gas choked out life" (Press release). University of Calgary. January 23, 2011. Retrieved 26 January 2011.
  376. ^ Hudspith, Victoria A.; Brimmer, Susan M.; Belcher, Claire M. (1 October 2014). "Latest Permian chars may derive from wildfires, not coal combustion". Geology. 42 (10): 879–882. Bibcode:2014Geo....42..879H. doi:10.1130/G35920.1. hdl:10871/20251. Retrieved 31 May 2023.
  377. ^ Yang, Q.Y. (2013). "The chemical compositions and abundances of volatiles in the Siberian large igneous province: Constraints on magmatic CO2 and SO2 emissions into the atmosphere". Chemical Geology. 339: 84–91. Bibcode:2013ChGeo.339...84T. doi:10.1016/j.chemgeo.2012.08.031.
  378. ^ Jones, Morgan T.; Jerram, Dougal A.; Svensen, Henrik H.; Grove, Clayton (1 January 2016). "The effects of large igneous provinces on the global carbon and sulphur cycles". Palaeogeography, Palaeoclimatology, Palaeoecology. Impact, Volcanism, Global changes and Mass Extinctions. 441: 4–21. Bibcode:2016PPP...441....4J. doi:10.1016/j.palaeo.2015.06.042. ISSN 0031-0182. Retrieved 12 January 2024 – via Elsevier Science Direct.
  379. ^ Winguth, Arne M.E.; Shields, Christine A.; Winguth, Cornelia (15 December 2015). "Transition into a Hothouse World at the Permian–Triassic boundary—A model study". Palaeogeography, Palaeoclimatology, Palaeoecology. 440: 316–327. doi:10.1016/j.palaeo.2015.09.008. Retrieved 13 October 2024 – via Elsevier Science Direct.
  380. ^ Isson, Terry T.; Zhang, Shuang; Lau, Kimberly V.; Rauzi, Sofia; Tosca, Nicholas J.; Penman, Donald E.; Planavsky, Noah J. (18 June 2022). "Marine siliceous ecosystem decline led to sustained anomalous Early Triassic warmth". Nature Communications. 13 (1): 3509. Bibcode:2022NatCo..13.3509I. doi:10.1038/s41467-022-31128-3. PMC 9206662. PMID 35717338.
  381. ^ Wang, Xiangdong; Cawood, Peter A.; Zhao, He; Zhao, Laishi; Grasby, Stephen E.; Chen, Zhong-Qiang; Wignall, Paul Barry; Lv, Zhengyi; Han, Chen (15 August 2018). "Mercury anomalies across the end Permian mass extinction in South China from shallow and deep water depositional environments". Earth and Planetary Science Letters. 496: 159–167. doi:10.1016/j.epsl.2018.05.044. Retrieved 13 October 2024 – via Elsevier Science Direct.
  382. ^ Shen, Jun; Shen, Jiubin; Yu, Jianxin; Algeo, Thomas J.; Smith, Roger M. H.; Botha, Jennifer; Frank, Tracy D.; Fielding, Christopher R.; Ward, Peter D.; Mather, Tamsin A. (3 January 2023). "Mercury evidence from southern Pangea terrestrial sections for end-Permian global volcanic effects". Nature Communications. 14 (1): 6. Bibcode:2023NatCo..14....6S. doi:10.1038/s41467-022-35272-8. PMC 9810726. PMID 36596767.
  383. ^ Wang, Xiangdong; Cawood, Peter A.; Zhao, He; Zhao, Laishi; Grasby, Stephen E.; Chen, Zhong-Qiang; Zhang, Lei (1 May 2019). "Global mercury cycle during the end-Permian mass extinction and subsequent Early Triassic recovery". Earth and Planetary Science Letters. 513: 144–155. doi:10.1016/j.epsl.2019.02.026. Retrieved 18 June 2024 – via Elsevier Science Direct.
  384. ^ a b Grasby, Stephen E.; Beauchamp, Benoit; Bond, David P. G.; Wignall, Paul Barry; Talavera, Cristina; Galloway, Jennifer M.; Piepjohn, Karsten; Reinhardt, Lutz; Blomeier, Dirk (1 September 2015). "Progressive environmental deterioration in northwestern Pangea leading to the latest Permian extinction". Geological Society of America Bulletin. 127 (9–10): 1331–1347. Bibcode:2015GSAB..127.1331G. doi:10.1130/B31197.1. Retrieved 14 January 2023.
  385. ^ Grasby, Stephen E.; Beauchamp, Benoit; Bond, David P. G.; Wignall, Paul Barry; Sanei, Hamed (2016). "Mercury anomalies associated with three extinction events (Capitanian Crisis, Latest Permian Extinction and the Smithian/Spathian Extinction) in NW Pangea". Geological Magazine. 153 (2): 285–297. Bibcode:2016GeoM..153..285G. doi:10.1017/S0016756815000436. S2CID 85549730.
  386. ^ Sanei, Hamed; Grasby, Stephen E.; Beauchamp, Benoit (1 January 2012). "Latest Permian mercury anomalies". Geology. 40 (1): 63–66. Bibcode:2012Geo....40...63S. doi:10.1130/G32596.1. Retrieved 14 January 2023.
  387. ^ Grasby, Stephen E.; Liu, Xiaojun; Yin, Runsheng; Ernst, Richard E.; Chen, Zhuoheng (19 May 2020). "Toxic mercury pulses into late Permian terrestrial and marine environments". Geology. 48 (8): 830–833. Bibcode:2020Geo....48..830G. doi:10.1130/G47295.1. S2CID 219495628.
  388. ^ Paterson, Niall W.; Rossi, Valentina M.; Schneebeli-Hermann, Elke (October 2024). "Volcanogenic mercury and plant mutagenesis during the end-Permian mass extinction: Palaeoecological perturbation in northern Pangaea". Gondwana Research. 134: 123–143. doi:10.1016/j.gr.2024.06.018. Retrieved 13 October 2024 – via Elsevier Science Direct.
  389. ^ Grasby, Stephen E.; Sanei, Hamed; Beauchamp, Benoit; Chen, Zhuoheng (2 August 2013). "Mercury deposition through the Permo–Triassic Biotic Crisis". Chemical Geology. 351: 209–216. Bibcode:2013ChGeo.351..209G. doi:10.1016/j.chemgeo.2013.05.022. Retrieved 31 May 2023.
  390. ^ Le Vaillant, Margaux; Barnes, Stephen J.; Mungall, James E.; Mungall, Emma L. (21 February 2017). "Role of degassing of the Noril'sk nickel deposits in the Permian–Triassic mass extinction event". Proceedings of the National Academy of Sciences of the United States of America. 114 (10): 2485–2490. Bibcode:2017PNAS..114.2485L. doi:10.1073/pnas.1611086114. PMC 5347598. PMID 28223492.
  391. ^ Rampino, Michael R.; Rodriguez, Sedelia; Baransky, Eva; Cai, Yue (29 September 2017). "Global nickel anomaly links Siberian Traps eruptions and the latest Permian mass extinction". Scientific Reports. 7 (1): 12416. Bibcode:2017NatSR...712416R. doi:10.1038/s41598-017-12759-9. PMC 5622041. PMID 28963524.
  392. ^ Li, Menghan; Grasby, Stephen E.; Wang, Shui-Jiong; Zhang, Xiaolin; Wasylenki, Laura E.; Xu, Yilun; Sun, Mingzhao; Beauchamp, Benoit; Hu, Dongping; Shen, Yanan (1 April 2021). "Nickel isotopes link Siberian Traps aerosol particles to the end-Permian mass extinction". Nature Communications. 12 (1): 2024. Bibcode:2021NatCo..12.2024L. doi:10.1038/s41467-021-22066-7. PMC 8016954. PMID 33795666.
  393. ^ Clarkson, M. O.; Richoz, Sylvain; Wood, R. A.; Maurer, F.; Krystyn, L.; McGurty, D. J.; Astratti, D. (July 2013). "A new high-resolution δ13C record for the Early Triassic: Insights from the Arabian Platform". Gondwana Research. 24 (1): 233–242. Bibcode:2013GondR..24..233C. doi:10.1016/j.gr.2012.10.002. Retrieved 26 May 2023.
  394. ^ Lehrmann, Daniel J.; Stepchinski, Leanne; Altiner, Demir; Orchard, Michael J.; Montgomery, Paul; Enos, Paul; Ellwood, Brooks B.; Bowring, Samuel A.; Ramezani, Jahandar; Wang, Hongmei; Wei, Jiayong; Yu, Meiyi; Griffiths, James D.; Minzoni, Marcello; Schaal, Ellen K.; Li, Xiaowei; Meyer, Katja M.; Payne, Jonathan L. (August 2015). "An integrated biostratigraphy (conodonts and foraminifers) and chronostratigraphy (paleomagnetic reversals, magnetic susceptibility, elemental chemistry, carbon isotopes and geochronology) for the Permian–Upper Triassic strata of Guandao section, Nanpanjiang Basin, south China". Journal of Asian Earth Sciences. 108: 117–135. doi:10.1016/j.jseaes.2015.04.030. Retrieved 18 June 2024 – via Elsevier Science Direct.
  395. ^ Shen, Jun; Algeo, Thomas J.; Planavsky, Noah J.; Yu, Jianxin; Feng, Qinglai; Song, Haijun; Song, Huyue; Rowe, Harry; Zhou, Lian; Chen, Jiubin (August 2019). "Mercury enrichments provide evidence of Early Triassic volcanism following the end-Permian mass extinction". Earth-Science Reviews. 195: 191–212. doi:10.1016/j.earscirev.2019.05.010. Retrieved 18 June 2024 – via Elsevier Science Direct.
  396. ^ Caravaca, Gwénaël; Thomazo, Christophe; Vennin, Emmanuelle; Olivier, Nicolas; Cocquerez, Théophile; Escarguel, Gilles; Fara, Emmanuel; Jenks, James F.; Bylund, Kevin G.; Stephen, Daniel A.; Brayard, Arnaud (July 2017). "Early Triassic fluctuations of the global carbon cycle: New evidence from paired carbon isotopes in the western USA basin". Global and Planetary Change. 154: 10–22. Bibcode:2017GPC...154...10C. doi:10.1016/j.gloplacha.2017.05.005. S2CID 135330761. Retrieved 13 November 2022.
  397. ^ Nelson, D. A.; Cottle, J. M. (29 March 2019). "Tracking voluminous Permian volcanism of the Choiyoi Province into central Antarctica". Lithosphere. 11 (3): 386–398. Bibcode:2019Lsphe..11..386N. doi:10.1130/L1015.1. S2CID 135130436.
  398. ^ Spalletti, Luis A.; Limarino, Carlos O. (29 September 2017). "The Choiyoi magmatism in south western Gondwana: implications for the end-permian mass extinction - a review". Andean Geology. 44 (3): 328. doi:10.5027/andgeoV44n3-a05. hdl:11336/66408. ISSN 0718-7106. Retrieved 20 September 2023.
  399. ^ a b Shen, Jun; Chen, Jiubin; Algeo, Thomas J.; Feng, Qinglai; Yu, Jianxin; Xu, Yi-Gang; Xu, Guozhen; Lei, Yong; Planavsky, Noah J.; Xie, Shucheng (10 December 2020). "Mercury fluxes record regional volcanism in the South China craton prior to the end-Permian mass extinction". Geology. 49 (4): 452–456. doi:10.1130/G48501.1. S2CID 230524628. Retrieved 28 March 2023.
  400. ^ Zhang, Hua; Zhang, Feifei; Chen, Jiubin; Erwin, Douglas H.; Syverson, Drew D.; Ni, Pei; Rampino, Michael R.; Chi, Zhe; Cai, Yao-Feng; Xiang, Lei; Li, We-Qiang; Liu, Sheng-Ao; Wang, Ru-Cheng; Wang, Xiang-Dong; Feng, Zhuo; Li, Hou-Min; Zhang, Ting; Cai, Mong-Ming; Zheng, Wang; Cui, Ying; Zhu, Xiang-Kun; Hou, Zeng-Qian; Wu, Fu-Yuan; Xu, Yi-Gang; Planavsky, Noah J.; Shen, Shu-zhong (17 November 2021). "Felsic volcanism as a factor driving the end-Permian mass extinction". Science Advances. 7 (47): eabh1390. Bibcode:2021SciA....7.1390Z. doi:10.1126/sciadv.abh1390. PMC 8597993. PMID 34788084.
  401. ^ Dickens, G.R. (2001). "The potential volume of oceanic methane hydrates with variable external conditions". Organic Geochemistry. 32 (10): 1179–1193. Bibcode:2001OrGeo..32.1179D. doi:10.1016/S0146-6380(01)00086-9.
  402. ^ Palfy J, Demeny A, Haas J, Htenyi M, Orchard MJ, Veto I (2001). "Carbon isotope anomaly at the Triassic–Jurassic boundary from a marine section in Hungary". Geology. 29 (11): 1047–1050. Bibcode:2001Geo....29.1047P. doi:10.1130/0091-7613(2001)029<1047:CIAAOG>2.0.CO;2. ISSN 0091-7613.
  403. ^ Reichow MK, Saunders AD, White RV, Pringle MS, Al'Muhkhamedov AI, Medvedev AI, Kirda NP (2002). " 40Ar 39Ar dates from the West Siberian Basin: Siberian flood basalt province doubled" (PDF). Science. 296 (5574): 1846–1849. Bibcode:2002Sci...296.1846R. doi:10.1126/science.1071671. PMID 12052954. S2CID 28964473.
  404. ^ Holser WT, Schoenlaub HP, Attrep Jr M, Boeckelmann K, Klein P, Magaritz M, Orth CJ, Fenninger A, Jenny C, Kralik M, Mauritsch H, Pak E, Schramm JF, Stattegger K, Schmoeller R (1989). "A unique geochemical record at the Permian/Triassic boundary". Nature. 337 (6202): 39–44. Bibcode:1989Natur.337...39H. doi:10.1038/337039a0. S2CID 8035040.
  405. ^ Dobruskina, I.A. (1987). "Phytogeography of Eurasia during the early Triassic". Palaeogeography, Palaeoclimatology, Palaeoecology. 58 (1–2): 75–86. Bibcode:1987PPP....58...75D. doi:10.1016/0031-0182(87)90007-1.
  406. ^ Ryskin, Gregory (September 2003). "Methane-driven oceanic eruptions and mass extinctions". Geology. 31 (9): 741–744. Bibcode:2003Geo....31..741R. doi:10.1130/G19518.1.
  407. ^ Krull, Evelyn S.; Retallack, Gregory J. (1 September 2000). "13C depth profiles from paleosols across the Permian–Triassic boundary: Evidence for methane release". Geological Society of America Bulletin. 112 (9): 1459–1472. Bibcode:2000GSAB..112.1459K. doi:10.1130/0016-7606(2000)112<1459:CDPFPA>2.0.CO;2. ISSN 0016-7606. Retrieved 3 July 2023.
  408. ^ a b Dickens GR, O'Neil JR, Rea DK, Owen RM (1995). "Dissociation of oceanic methane hydrate as a cause of the carbon isotope excursion at the end of the Paleocene". Paleoceanography and Paleoclimatology. 10 (6): 965–971. Bibcode:1995PalOc..10..965D. doi:10.1029/95PA02087.
  409. ^ Schrag DP, Berner RA, Hoffman PF, Halverson GP (2002). "On the initiation of a snowball Earth". Geochemistry, Geophysics, Geosystems. 3 (6): 1–21. Bibcode:2002GGG.....3.1036S. doi:10.1029/2001GC000219. Preliminary abstract at Schrag, D.P. (June 2001). "On the initiation of a snowball Earth". Geological Society of America. Archived from the original on 2018-04-25. Retrieved 2008-04-20.
  410. ^ Benton, Michael James; Twitchett, R.J. (2003). "How to kill (almost) all life: The end-Permian extinction event". Trends in Ecology & Evolution. 18 (7): 358–365. doi:10.1016/S0169-5347(03)00093-4.
  411. ^ Cui, Ying; Li, Mingsong; van Soelen, Elsbeth E.; Peterse, Francien; M. Kürschner, Wolfram (7 September 2021). "Massive and rapid predominantly volcanic CO2 emission during the end-Permian mass extinction". Proceedings of the National Academy of Sciences of the United States of America. 118 (37): e2014701118. Bibcode:2021PNAS..11814701C. doi:10.1073/pnas.2014701118. PMC 8449420. PMID 34493684.
  412. ^ Shen, Shu-Zhong; Cao, Chang-Qun; Henderson, Charles M.; Wang, Xiang-Dong; Shi, Guang R.; Wang, Yue; Wang, Wei (January 2006). "End-Permian mass extinction pattern in the northern peri-Gondwanan region". Palaeoworld. 15 (1): 3–30. doi:10.1016/j.palwor.2006.03.005. Retrieved 26 May 2023.
  413. ^ Majorowicz, J.; Grasby, S. E.; Safanda, J.; Beauchamp, B. (1 May 2014). "Gas hydrate contribution to Late Permian global warming". Earth and Planetary Science Letters. 393: 243–253. Bibcode:2014E&PSL.393..243M. doi:10.1016/j.epsl.2014.03.003. ISSN 0012-821X. Retrieved 12 January 2024 – via Elsevier Science Direct.
  414. ^ Reddin, Carl J.; Nätscher, Paulina; Kocsis, Ádám T.; Pörtner, Hans-Otto; Kiessling, Wolfgang (10 February 2020). "Marine clade sensitivities to climate change conform across timescales". Nature Climate Change. 10 (3): 249–253. Bibcode:2020NatCC..10..249R. doi:10.1038/s41558-020-0690-7. S2CID 211074044. Retrieved 26 March 2023.
  415. ^ Cui, Ying; Kump, Lee R.; Ridgwell, Andy (1 November 2013). "Initial assessment of the carbon emission rate and climatic consequences during the end-Permian mass extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 389: 128–136. Bibcode:2013PPP...389..128C. doi:10.1016/j.palaeo.2013.09.001. Retrieved 26 June 2023.
  416. ^ He, Jiawei; Hu, Xiumian; Li, Juan; Kemp, David B.; Hou, Mingcai; Han, Zhong (15 November 2024). "Millennial-scale sedimentary evolution of carbonate platforms during the Permian–Triassic boundary hyperthermal event". Palaeogeography, Palaeoclimatology, Palaeoecology. 654: 112455. doi:10.1016/j.palaeo.2024.112455. Retrieved 13 October 2024 – via Elsevier Science Direct.
  417. ^ Li, Fei; Yan, Jiaxin; Chen, Zhong-Qiang; Ogg, James G.; Tian, Li; Korngreen, Dorit; Liu, Ke; Ma, Zulu; Woods, Adam D. (October 2015). "Global oolite deposits across the Permian–Triassic boundary: A synthesis and implications for palaeoceanography immediately after the end-Permian biocrisis". Earth-Science Reviews. 149: 163–180. doi:10.1016/j.earscirev.2014.12.006. Retrieved 18 June 2024 – via Elsevier Science Direct.
  418. ^ Wignall, Paul Barry; Chu, Daoliang; Hilton, Jason M.; Dal Corso, Jacopo; Wu, Yuyang; Wang, Yao; Atkinson, Jed; Tong, Jinnan (June 2020). "Death in the shallows: The record of Permo-Triassic mass extinction in paralic settings, southwest China". Global and Planetary Change. 189: 103176. Bibcode:2020GPC...18903176W. doi:10.1016/j.gloplacha.2020.103176. S2CID 216302513.
  419. ^ Foster, William J.; Hirtz, J. A.; Farrell, C.; Reistroffer, M.; Twitchett, Richard J.; Martindale, R. C. (24 January 2022). "Bioindicators of severe ocean acidification are absent from the end-Permian mass extinction". Scientific Reports. 12 (1): 1202. Bibcode:2022NatSR..12.1202F. doi:10.1038/s41598-022-04991-9. PMC 8786885. PMID 35075151.
  420. ^ Wignall, Paul Barry (29 September 2015). "The Killing Seas". The Worst of Times: How Life on Earth Survived Eighty Million Years of Extinctions. Princeton: Princeton University Press. pp. 83–84. ISBN 978-0-691-14209-8.
  421. ^ a b Fraiser, Margaret L.; Bottjer, David P. (20 August 2007). "Elevated atmospheric CO2 and the delayed biotic recovery from the end-Permian mass extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 252 (1–2): 164–175. Bibcode:2007PPP...252..164F. doi:10.1016/j.palaeo.2006.11.041. Retrieved 31 May 2023.
  422. ^ Heydari, Ezat; Arzani, Nasser; Hassanzadeh, Jamshin (7 July 2008). "Mantle plume: The invisible serial killer — Application to the Permian–Triassic boundary mass extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 264 (1–2): 147–162. Bibcode:2008PPP...264..147H. doi:10.1016/j.palaeo.2008.04.013. Retrieved 26 May 2023.
  423. ^ Sephton, Mark A.; Jiao, Dan; Engel, Michael H.; Looy, Cindy V.; Visscher, Henk (1 February 2015). "Terrestrial acidification during the end-Permian biosphere crisis?". Geology. 43 (2): 159–162. Bibcode:2015Geo....43..159S. doi:10.1130/G36227.1. hdl:10044/1/31566. Retrieved 23 December 2022.
  424. ^ Sheldon, Nathan D. (28 February 2006). "Abrupt chemical weathering increase across the Permian–Triassic boundary". Palaeogeography, Palaeoclimatology, Palaeoecology. 231 (3–4): 315–321. Bibcode:2006PPP...231..315S. doi:10.1016/j.palaeo.2005.09.001. Retrieved 24 April 2023.
  425. ^ Buatois, Luis A.; Borruel-Abadía, Violeta; De la Horra, Raúl; Galán-Abellán, Ana Belén; López-Gómez, José; Barrenechea, José F.; Arche, Alfredo (25 March 2021). "Impact of Permian mass extinctions on continental invertebrate infauna". Terra Nova. 33 (5): 455–464. Bibcode:2021TeNov..33..455B. doi:10.1111/ter.12530. S2CID 233616369. Retrieved 23 December 2022.
  426. ^ Xu, Guozhen; Deconinck, Jean-François; Feng, Qinglai; Baudin, François; Pellenard, Pierre; Shen, Jun; Bruneau, Ludovic (15 May 2017). "Clay mineralogical characteristics at the Permian–Triassic Shangsi section and their paleoenvironmental and/or paleoclimatic significance". Palaeogeography, Palaeoclimatology, Palaeoecology. 474: 152–163. Bibcode:2017PPP...474..152X. doi:10.1016/j.palaeo.2016.07.036. Retrieved 23 December 2022.
  427. ^ Wang, Chunjiang (January 2007). "Anomalous hopane distributions at the Permian–Triassic boundary, Meishan, China – Evidence for the end-Permian marine ecosystem collapse". Organic Geochemistry. 38 (1): 52–66. Bibcode:2007OrGeo..38...52W. doi:10.1016/j.orggeochem.2006.08.014. Retrieved 23 May 2023.
  428. ^ Borruel-Abadía, Violeta; Barrenechea, José F.; Galán-Abellán, Ana Belén; De la Horra, Raúl; López-Gómez, José; Ronchi, Ausonio; Luque, Francisco Javier; Alonso-Azcárate, Jacinto; Marzo, Mariano (20 June 2019). "Could acidity be the reason behind the Early Triassic biotic crisis on land?". Chemical Geology. 515: 77–86. Bibcode:2019ChGeo.515...77B. doi:10.1016/j.chemgeo.2019.03.035. S2CID 134704729. Retrieved 18 December 2022.
  429. ^ Wang, Han; He, Weihong; Xiao, Yifan; Yang, Tinglu; Zhang, Kexin; Wu, Huiting; Huang, Yafei; Peng, Xingfang; Wu, Shunbao (1 July 2023). "Stagewise collapse of biotic communities and its relations to oxygen depletion along the north margin of Nanpanjiang Basin during the Permian–Triassic transition". Palaeogeography, Palaeoclimatology, Palaeoecology. 621: 111569. Bibcode:2023PPP...62111569W. doi:10.1016/j.palaeo.2023.111569. Retrieved 31 May 2023.
  430. ^ Hülse, Dominik; Lau, Kimberly V.; Van de Velde, Sebastiaan J.; Arndt, Sandra; Meyer, Katja M.; Ridgwell, Andy (28 October 2021). "End-Permian marine extinction due to temperature-driven nutrient recycling and euxinia". Nature Geoscience. 14 (11): 862–867. Bibcode:2021NatGe..14..862H. doi:10.1038/s41561-021-00829-7. S2CID 240076553. Retrieved 8 January 2023.
  431. ^ Benton, Michael James (January 2008). "Presidential Address 2007: The end-Permian mass extinction — events on land in Russia". Proceedings of the Geologists' Association. 119 (2): 119–136. doi:10.1016/S0016-7878(08)80313-6. Retrieved 18 June 2024 – via Elsevier Science Direct.
  432. ^ a b Wignall, Paul Barry; Twitchett, Richard J. (2002). "Extent, duration, and nature of the Permian-Triassic superanoxic event". In Koeberl, Christian; MacLeod, Kenneth G. (eds.). Catastrophic Events and Mass Extinction: Impacts and Beyond. Geological Society of America Special Papers No. 356. pp. 395–413. Bibcode:2002GSASP.356..395W. doi:10.1130/0-8137-2356-6.395. ISBN 9780813723563. OL 11351081M. Retrieved 17 February 2024.
  433. ^ Wei, Hengye; Algeo, Thomas J.; Yu, Hao; Wang, Jiangguo; Guo, Chuan; Shi, Guo (15 April 2015). "Episodic euxinia in the Changhsingian (late Permian) of South China: Evidence from framboidal pyrite and geochemical data". Sedimentary Geology. 319: 78–97. doi:10.1016/j.sedgeo.2014.11.008. Retrieved 18 June 2024 – via Elsevier Science Direct.
  434. ^ a b Brennecka, Gregory A.; Herrmann, Achim D.; Algeo, Thomas J.; Anbar, Ariel D. (10 October 2011). "Rapid expansion of oceanic anoxia immediately before the end-Permian mass extinction". Proceedings of the National Academy of Sciences of the United States of America. 108 (43): 17631–17634. doi:10.1073/pnas.1106039108. PMC 3203792. PMID 21987794.
  435. ^ Wang, Wen-qian; Zhang, Feifei; Shen, Shu-zhong; Bizzarro, Martin; Garbelli, Claudio; Zheng, Quan-feng; Zhang, Yi-chun; Yuan, Dong-xun; Shi, Yu-kun; Cao, Mengchun; Dahl, Tais W. (15 September 2022). "Constraining marine anoxia under the extremely oxygenated Permian atmosphere using uranium isotopes in calcitic brachiopods and marine carbonates". Earth and Planetary Science Letters. 594: 117714. Bibcode:2022E&PSL.59417714W. doi:10.1016/j.epsl.2022.117714. S2CID 250941149. Retrieved 31 May 2023.
  436. ^ Luo, Genming; Wang, Yongbiao; Algeo, Thomas J.; Kump, Lee R.; Bai, Xiao; Yang, Hao; Yao, Le; Xie, Shucheng (1 July 2011). "Enhanced nitrogen fixation in the immediate aftermath of the latest Permian marine mass extinction". Geology. 39 (7): 647–650. Bibcode:2011Geo....39..647L. doi:10.1130/G32024.1. Retrieved 26 May 2023.
  437. ^ Stüeken, Eva E.; Foriel, Julien; Buick, Roger; Schoepfer, Shane D. (2 September 2015). "Selenium isotope ratios, redox changes and biological productivity across the end-Permian mass extinction". Chemical Geology. 410: 28–39. doi:10.1016/j.chemgeo.2015.05.021. Retrieved 13 March 2024 – via Elsevier Science Direct.
  438. ^ Grice, Kliti; Nabbefeld, Birgit; Maslen, Ercin (November 2007). "Source and significance of selected polycyclic aromatic hydrocarbons in sediments (Hovea-3 well, Perth Basin, Western Australia) spanning the Permian–Triassic boundary". Organic Geochemistry. 38 (11): 1795–1803. Bibcode:2007OrGeo..38.1795G. doi:10.1016/j.orggeochem.2007.07.001. Retrieved 31 May 2023.
  439. ^ Song, Haijun; Wignall, Paul Barry; Tong, Jinnan; Bond, David P. G.; Song, Huyue; Lai, Xulong; Zhang, Kexin; Wang, Hongmei; Chen, Yanlong (1 November 2012). "Geochemical evidence from bio-apatite for multiple oceanic anoxic events during Permian–Triassic transition and the link with end-Permian extinction and recovery". Earth and Planetary Science Letters. 353–354: 12–21. Bibcode:2012E&PSL.353...12S. doi:10.1016/j.epsl.2012.07.005. Retrieved 13 January 2023.
  440. ^ Müller, J.; Sun, Y. D.; Fang, F.; Regulous, M.; Joachimski, Michael M. (March 2023). "Manganous water column in the Tethys Ocean during the Permian-Triassic transition". Global and Planetary Change. 222: 104067. Bibcode:2023GPC...22204067M. doi:10.1016/j.gloplacha.2023.104067. S2CID 256800036. Retrieved 26 June 2023.
  441. ^ Xiang, Lei; Zhang, Hua; Schoepfer, Shane D.; Cao, Chang-qun; Zheng, Quan-feng; Yuan, Dong-xun; Cai, Yao-feng; Shen, Shu-zhong (15 April 2020). "Oceanic redox evolution around the end-Permian mass extinction at Meishan, South China". Palaeogeography, Palaeoclimatology, Palaeoecology. 544: 109626. doi:10.1016/j.palaeo.2020.109626. Retrieved 1 August 2024 – via Elsevier Science Direct.
  442. ^ Stebbins, Alan; Williams, Jeremy; Brookfield, Michael; Nye Jr., Steven W.; Hannigan, Robyn (15 February 2019). "Frequent euxinia in southern Neo-Tethys Ocean prior to the end-Permian biocrisis: Evidence from the Spiti region, India". Palaeogeography, Palaeoclimatology, Palaeoecology. 516: 1–10. Bibcode:2019PPP...516....1S. doi:10.1016/j.palaeo.2018.11.030. S2CID 134724104.
  443. ^ Zhang, Li-Jun; Zhang, Xin; Buatois, Luis A.; Mángano, M. Gabriela; Shi, Guang R. Shi; Gong, Yi-Ming; Qi, Yong-An (December 2000). "Periodic fluctuations of marine oxygen content during the latest Permian". Global and Planetary Change. 195: 103326. doi:10.1016/j.gloplacha.2020.103326. S2CID 224881713. Retrieved 2 March 2023.
  444. ^ Cao, Changqun; Gordon D. Love; Lindsay E. Hays; Wei Wang; Shuzhong Shen; Roger E. Summons (2009). "Biogeochemical evidence for euxinic oceans and ecological disturbance presaging the end-Permian mass extinction event". Earth and Planetary Science Letters. 281 (3–4): 188–201. Bibcode:2009E&PSL.281..188C. doi:10.1016/j.epsl.2009.02.012.
  445. ^ Schoepfer, Shane D.; Henderson, Charles M.; Garrison, Geoffrey H.; Ward, Peter Douglas (1 January 2012). "Cessation of a productive coastal upwelling system in the Panthalassic Ocean at the Permian–Triassic Boundary". Palaeogeography, Palaeoclimatology, Palaeoecology. 313–314: 181–188. Bibcode:2012PPP...313..181S. doi:10.1016/j.palaeo.2011.10.019. Retrieved 21 December 2022.
  446. ^ Hays, Lindsay; Kliti Grice; Clinton B. Foster; Roger E. Summons (2012). "Biomarker and isotopic trends in a Permian–Triassic sedimentary section at Kap Stosch, Greenland" (PDF). Organic Geochemistry. 43: 67–82. Bibcode:2012OrGeo..43...67H. doi:10.1016/j.orggeochem.2011.10.010. hdl:20.500.11937/26597.
  447. ^ Hays, Lindsay E.; Beatty, Tyler; Henderson, Charles M.; Love, Gordon D.; Summons, Roger E. (January–September 2007). "Evidence for photic zone euxinia through the end-Permian mass extinction in the Panthalassic Ocean (Peace River Basin, Western Canada)". Palaeoworld. 16 (1–3): 39–50. doi:10.1016/j.palwor.2007.05.008. Retrieved 23 May 2023.
  448. ^ Xie, Shucheng; Algeo, Thomas J.; Zhou, Wenfeng; Ruan, Xiaoyan; Luo, Genming; Huang, Junhua; Yan, Jiaxin (15 February 2017). "Contrasting microbial community changes during mass extinctions at the Middle/Late Permian and Permian/Triassic boundaries". Earth and Planetary Science Letters. 460: 180–191. Bibcode:2017E&PSL.460..180X. doi:10.1016/j.epsl.2016.12.015. Retrieved 4 January 2023.
  449. ^ Luo, Genming; Huang, Junhuang; Xie, Shucheng; Wignall, Paul Barry; Tang, Xinyan; Huang, Xianyu; Yin, Hongfu (13 February 2009). "Relationships between carbon isotope evolution and variation of microbes during the Permian–Triassic transition at Meishan Section, South China". International Journal of Earth Sciences. 99 (4): 775–784. doi:10.1007/s00531-009-0421-9. ISSN 1437-3254. Retrieved 1 August 2024 – via Springer Link.
  450. ^ Sun, Ruoyu; Liu, Yi; Sonke, Jeroen E.; Feifei, Zhang; Zhao, Yaqiu; Zhang, Yonggen; Chen, Jiubin; Liu, Cong-Qiang; Shen, Shuzhong; Anbar, Ariel D.; Zheng, Wang (8 May 2023). "Mercury isotope evidence for marine photic zone euxinia across the end-Permian mass extinction". Communications Earth & Environment. 4 (1): 159. Bibcode:2023ComEE...4..159S. doi:10.1038/s43247-023-00821-6. S2CID 258577845.
  451. ^ Penn, Justin L.; Deutsch, Curtis; Payne, Jonathan L.; Sperling, Erik A. (7 December 2018). "Temperature-dependent hypoxia explains biogeography and severity of end-Permian marine mass extinction". Science. 362 (6419): 1–6. Bibcode:2018Sci...362.1327P. doi:10.1126/science.aat1327. PMID 30523082. S2CID 54456989.
  452. ^ Shen, Jun; Chen, Jiubin; Algeo, Thomas J.; Yuan, Shengliu; Feng, Qinglai; Yu, Jianxin; Zhou, Lian; O'Connell, Brennan; Planavsky, Noah J. (5 April 2019). "Evidence for a prolonged Permian–Triassic extinction interval from global marine mercury records". Nature Communications. 10 (1): 1563. Bibcode:2019NatCo..10.1563S. doi:10.1038/s41467-019-09620-0. PMC 6450928. PMID 30952859.
  453. ^ He, Weihong; Weldon, Elizabeth A.; Yang, Tinglu; Wang, Han; Xiao, Yifan; Zhang, Kexin; Peng, Xingfang; Feng, Qinglai (1 September 2024). "An end-Permian two-stage extinction pattern in the deep-water Dongpan Section, and its relationship to the migration and vertical expansion of the oxygen minimum zone in the South China Basin". Palaeogeography, Palaeoclimatology, Palaeoecology. 649: 112307. doi:10.1016/j.palaeo.2024.112307. Retrieved 13 October 2024 – via Elsevier Science Direct.
  454. ^ Meyers, Katja; L.R. Kump; A. Ridgwell (September 2008). "Biogeochemical controls on photic-zone euxinia during the end-Permian mass extinction". Geology. 36 (9): 747–750. Bibcode:2008Geo....36..747M. doi:10.1130/g24618a.1.
  455. ^ a b Saunders, Andrew D. (9 June 2015). "Two LIPs and two Earth-system crises: the impact of the North Atlantic Igneous Province and the Siberian Traps on the Earth-surface carbon cycle". Geological Magazine. 153 (2): 201–222. doi:10.1017/S0016756815000175. hdl:2381/32095. S2CID 131273374. Retrieved 26 May 2023.
  456. ^ Kidder, David L.; Worsley, Thomas R. (15 February 2004). "Causes and consequences of extreme Permo-Triassic warming to globally equable climate and relation to the Permo-Triassic extinction and recovery". Palaeogeography, Palaeoclimatology, Palaeoecology. 203 (3–4): 207–237. Bibcode:2004PPP...203..207K. doi:10.1016/S0031-0182(03)00667-9. Retrieved 23 May 2023.
  457. ^ Sephton, Mark A.; Looy, Cindy V.; Brinkhuis, Henk; Wignall, Paul Barry; De Leeuw, Jan W.; Visscher, Henk (1 December 2005). "Catastrophic soil erosion during the end-Permian biotic crisis". Geology. 33 (12): 941–944. Bibcode:2005Geo....33..941S. doi:10.1130/G21784.1. Retrieved 26 May 2023.
  458. ^ Duan, Xiong; Shi, Zhiqiang (30 May 2024). "Sedimentary records of sea level fall during the end-Permian in the upper Yangtze region (southern China): Implications for the mass extinction". Heliyon. 10 (10): e31226. doi:10.1016/j.heliyon.2024.e31226. PMC 11126861. PMID 38799747.
  459. ^ Algeo, Thomas J.; Henderson, Charles M.; Tong, Jinnan; Feng, Qinglai; Yin, Hongfu; Tyson, Richard V. (June 2013). "Plankton and productivity during the Permian–Triassic boundary crisis: An analysis of organic carbon fluxes". Global and Planetary Change. 105: 52–67. Bibcode:2013GPC...105...52A. doi:10.1016/j.gloplacha.2012.02.008. Retrieved 3 July 2023.
  460. ^ Grasby, Stephen E.; Ardakani, Omid H.; Liu, Xiaojun; Bond, David P. G.; Wignall, Paul Barry; Strachan, Lorna J. (29 November 2023). "Marine snowstorm during the Permian−Triassic mass extinction". Geology. 52 (2): 120–124. doi:10.1130/G51497.1. ISSN 0091-7613.
  461. ^ Schobben, Martin; Foster, William J.; Sleveland, Arve R. N.; Zuchuat, Valentin; Svensen, Henrik H.; Planke, Sverre; Bond, David P. G.; Marcelis, Fons; Newton, Robert J.; Wignall, Paul Barry; Poulton, Simon W. (17 August 2020). "A nutrient control on marine anoxia during the end-Permian mass extinction". Nature Geoscience. 13 (9): 640–646. Bibcode:2020NatGe..13..640S. doi:10.1038/s41561-020-0622-1. hdl:1874/408736. S2CID 221146234. Retrieved 8 January 2023.
  462. ^ Liao, Wei; Bond, David P. G.; Wang, Yongbiao; He, Lei; Yang, Hao; Weng, Zeting; Li, Guoshan (15 November 2017). "An extensive anoxic event in the Triassic of the South China Block: A pyrite framboid study from Dajiang and its implications for the cause(s) of oxygen depletion". Palaeogeography, Palaeoclimatology, Palaeoecology. 486: 86–95. Bibcode:2017PPP...486...86L. doi:10.1016/j.palaeo.2016.11.012. Retrieved 8 May 2023.
  463. ^ Fio, Karmen; Spangenberg, Jorge E.; Vlahović, Igor; Sremac, Jasenka; Velić, Ivo; Mrinjek, Ervin (1 November 2010). "Stable isotope and trace element stratigraphy across the Permian–Triassic transition: A redefinition of the boundary in the Velebit Mountain, Croatia". Chemical Geology. 278 (1–2): 38–57. Bibcode:2010ChGeo.278...38F. doi:10.1016/j.chemgeo.2010.09.001. Retrieved 24 April 2023.
  464. ^ Newby, Sean M.; Owens, Jeremy D.; Schoepfer, Shane D.; Algeo, Thomas J. (2 August 2021). "Transient ocean oxygenation at end-Permian mass extinction onset shown by thallium isotopes". Nature Geoscience. 14 (9): 678–683. Bibcode:2021NatGe..14..678N. doi:10.1038/s41561-021-00802-4. ISSN 1752-0908. S2CID 236780878. Retrieved 27 December 2023.
  465. ^ Shen, Jun; Schoepfer, Shane D.; Feng, Qinglai; Zhou, Lian; Yu, Jianxin; Song, Huyue; Wei, Hengye; Algeo, Thomas J. (October 2015). "Marine productivity changes during the end-Permian crisis and Early Triassic recovery". Earth-Science Reviews. 149: 136–162. Bibcode:2015ESRv..149..136S. doi:10.1016/j.earscirev.2014.11.002. Retrieved 20 January 2023.
  466. ^ Wignall, Paul Barry; Hallam, Anthony (May 1992). "Anoxia as a cause of the Permian/Triassic mass extinction: facies evidence from northern Italy and the western United States". Palaeogeography, Palaeoclimatology, Palaeoecology. 93 (1–2): 21–46. Bibcode:1992PPP....93...21W. doi:10.1016/0031-0182(92)90182-5. Retrieved 20 January 2023.
  467. ^ Chen, Zhong-Qiang; Yang, Hao; Luo, Mao; Benton, Michael James; Kaiho, Kunio; Zhao, Laishi; Huang, Yuangeng; Zhang, Kexing; Fang, Yuheng; Jiang, Haishui; Qiu, Huan; Li, Yang; Tu, Chengyi; Shi, Lei; Zhang, Lei; Feng, Xueqian; Chen, Long (October 2015). "Complete biotic and sedimentary records of the Permian–Triassic transition from Meishan section, South China: Ecologically assessing mass extinction and its aftermath". Earth-Science Reviews. 149: 67–107. Bibcode:2015ESRv..149...67C. doi:10.1016/j.earscirev.2014.10.005. hdl:1983/d2b89cc3-b0a8-41b5-a220-b3d7d75687e0. Retrieved 14 January 2023.
  468. ^ Pietsch, Carlie; Mata, Scott A.; Bottjer, David J. (1 April 2014). "High temperature and low oxygen perturbations drive contrasting benthic recovery dynamics following the end-Permian mass extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 399: 98–113. Bibcode:2014PPP...399...98P. doi:10.1016/j.palaeo.2014.02.011. Retrieved 2 April 2023.
  469. ^ Algeo, Thomas J.; Chen, Zhong Qiang; Fraiser, Margaret L.; Twitchett, Richard J. (15 July 2011). "Terrestrial–marine teleconnections in the collapse and rebuilding of Early Triassic marine ecosystems". Palaeogeography, Palaeoclimatology, Palaeoecology. Permian - Triassic ecosystems: collapse and rebuilding. 308 (1): 1–11. Bibcode:2011PPP...308....1A. doi:10.1016/j.palaeo.2011.01.011. ISSN 0031-0182. Retrieved 24 November 2023.
  470. ^ Tian, Li; Tong, Jinnan; Algeo, Thomas J.; Song, Haijun; Song, Huyue; Chu, Daoliang; Shi, Lei; Bottjer, David J. (15 October 2014). "Reconstruction of Early Triassic ocean redox conditions based on framboidal pyrite from the Nanpanjiang Basin, South China". Palaeogeography, Palaeoclimatology, Palaeoecology. 412: 68–79. Bibcode:2014PPP...412...68T. doi:10.1016/j.palaeo.2014.07.018. Retrieved 8 May 2023.
  471. ^ Song, Haijun; Wignall, Paul Barry; Tong, Jinnan; Song, Huyue; Chen, Jing; Chu, Daoliang; Tian, Li; Luo, Mao; Zong, Keqing; Chen, Yanlong; Lai, Xulong; Zhang, Kexin; Wang, Hongmei (15 August 2015). "Integrated Sr isotope variations and global environmental changes through the Late Permian to early Late Triassic". Earth and Planetary Science Letters. 424: 140–147. Bibcode:2015E&PSL.424..140S. doi:10.1016/j.epsl.2015.05.035. ISSN 0012-821X. Retrieved 24 November 2023.
  472. ^ Muto, Shun; Takahashi, Satoshi; Yamakita, Satoshi; Suzuki, Noritoshi; Suzuki, Nozomi; Aita, Yoshiaki (15 January 2018). "High sediment input and possible oceanic anoxia in the pelagic Panthalassa during the latest Olenekian and early Anisian: Insights from a new deep-sea section in Ogama, Tochigi, Japan". Palaeogeography, Palaeoclimatology, Palaeoecology. 490: 687–707. Bibcode:2018PPP...490..687M. doi:10.1016/j.palaeo.2017.11.060. ISSN 0031-0182. Retrieved 24 November 2023.
  473. ^ Woods, Adam D.; Zonneveld, John-Paul; Wakefield, Ryan (13 December 2023). "Hyperthermal-driven anoxia and reduced productivity in the aftermath of the Permian-Triassic mass extinction: a case study from Western Canada". Frontiers in Earth Science. 11. Bibcode:2023FrEaS..1123413W. doi:10.3389/feart.2023.1323413. ISSN 2296-6463.
  474. ^ Li, Guoshan; Wang, Yongbiao; Shi, Guang R.; Liao, Wei; Yu, Lixue (15 April 2016). "Fluctuations of redox conditions across the Permian–Triassic boundary—New evidence from the GSSP section in Meishan of South China". Palaeogeography, Palaeoclimatology, Palaeoecology. 448: 48–58. Bibcode:2016PPP...448...48L. doi:10.1016/j.palaeo.2015.09.050. Retrieved 26 May 2023.
  475. ^ Schobben, Martin; Stebbins, Alan; Algeo, Thomas J.; Strauss, Harald; Leda, Lucyna; Haas, János; Struck, Ulrich; Korn, Dieter; Korte, Christoph (15 November 2017). "Volatile earliest Triassic sulfur cycle: A consequence of persistent low seawater sulfate concentrations and a high sulfur cycle turnover rate?". Palaeogeography, Palaeoclimatology, Palaeoecology. 486: 74–85. Bibcode:2017PPP...486...74S. doi:10.1016/j.palaeo.2017.02.025. Retrieved 26 May 2023.
  476. ^ Zhang R, Follows MJ, Grotzinger JP, Marshall J (2001). "Could the Late Permian deep ocean have been anoxic?". Paleoceanography and Paleoclimatology. 16 (3): 317–329. Bibcode:2001PalOc..16..317Z. doi:10.1029/2000PA000522.
  477. ^ Mays, Chris; McLoughlin, Stephen; Frank, Tracy D.; Fielding, Christopher R.; Slater, Sam M.; Vajda, Vivi (17 September 2021). "Lethal microbial blooms delayed freshwater ecosystem recovery following the end-Permian extinction". Nature Communications. 12 (1): 5511. Bibcode:2021NatCo..12.5511M. doi:10.1038/s41467-021-25711-3. PMC 8448769. PMID 34535650.
  478. ^ a b Smith, Roger M. H.; Botha, Jennifer; Viglietti, Pia A. (15 October 2022). "Taphonomy of drought afflicted tetrapods in the Early Triassic Karoo Basin, South Africa". Palaeogeography, Palaeoclimatology, Palaeoecology. 604: 111207. Bibcode:2022PPP...60411207S. doi:10.1016/j.palaeo.2022.111207. S2CID 251781291. Retrieved 23 May 2023.
  479. ^ Song, Yi; Tian, Yuan; Yu, Jianxin; Algeo, Thomas J.; Luo, Genming; Chu, Daoliang; Xie, Shucheng (August 2022). "Wildfire response to rapid climate change during the Permian-Triassic biotic crisis". Global and Planetary Change. 215: 103872. Bibcode:2022GPC...21503872S. doi:10.1016/j.gloplacha.2022.103872. S2CID 249857664. Retrieved 23 May 2023.
  480. ^ Benton, Michael James; Newell, Andrew J. (May 2014). "Impacts of global warming on Permo-Triassic terrestrial ecosystems". Gondwana Research. 25 (4): 1308–1337. Bibcode:2014GondR..25.1308B. doi:10.1016/j.gr.2012.12.010. Retrieved 26 May 2023.
  481. ^ Smith, R.M.H. (16 November 1999). "Changing fluvial environments across the Permian–Triassic boundary in the Karoo Basin, South Africa and possible causes of tetrapod extinctions". Palaeogeography, Palaeoclimatology, Palaeoecology. 117 (1–2): 81–104. Bibcode:1995PPP...117...81S. doi:10.1016/0031-0182(94)00119-S.
  482. ^ Chinsamy-Turan (2012). Anusuya (ed.). Forerunners of mammals : radiation, histology, biology. Bloomington: Indiana University Press. ISBN 978-0-253-35697-0.
  483. ^ Gastaldo, Robert A.; Kus, Kaci; Tabor, Neil; Neveling, Johann (26 June 2020). "Calcic Vertisols in the upper Daptocephalus Assemblage Zone, Balfour Formation, Karoo Basin, South Africa: Implications for Late Permian Climate". Journal of Sedimentary Research. 90 (6): 609–628. Bibcode:2020JSedR..90..609G. doi:10.2110/jsr.2020.32. S2CID 221865490. Retrieved 31 May 2023.
  484. ^ Fielding, Christopher R.; Frank, Tracy D.; Tevyaw, Allen P.; Savatic, Katarina; Vajda, Vivi; McLoughlin, Stephen; Mays, Chris; Nicoll, Robert S.; Bocking, Malcolm; Crowley, James L. (19 July 2020). "Sedimentology of the continental end-Permian extinction event in the Sydney Basin, eastern Australia". Sedimentology. 68 (1): 30–62. doi:10.1111/sed.12782. S2CID 225605914. Retrieved 30 October 2022.
  485. ^ Fielding, Christopher R.; Frank, Tracy D.; McLoughlin, Stephen; Vajda, Vivi; Mays, Chris; Tevyaw, Allen P.; Winguth, Arne; Winguth, Cornelia; Nicoll, Robert S.; Bocking, Malcolm; Crowley, James L. (23 January 2019). "Age and pattern of the southern high-latitude continental end-Permian extinction constrained by multiproxy analysis". Nature Communications. 10 (385): 385. Bibcode:2019NatCo..10..385F. doi:10.1038/s41467-018-07934-z. PMC 6344581. PMID 30674880.
  486. ^ Baucon, Andrea; Ronchi, Ausonio; Felletti, Fabrizio; Neto de Carvalho, Carlos (15 September 2014). "Evolution of Crustaceans at the edge of the end-Permian crisis: Ichnonetwork analysis of the fluvial succession of Nurra (Permian–Triassic, Sardinia, Italy)". Palaeogeography, Palaeoclimatology, Palaeoecology. 410: 74–103. Bibcode:2014PPP...410...74B. doi:10.1016/j.palaeo.2014.05.034. Retrieved 8 April 2023.
  487. ^ Black, Benjamin A.; Lamarque, Jean-François; Shields, Christine A.; Elkins-Tanton, Linda T.; Kiehl, Jeffrey T. (1 January 2014). "Acid rain and ozone depletion from pulsed Siberian Traps magmatism". Geology. 42 (1): 67–70. doi:10.1130/G34875.1. ISSN 1943-2682. Retrieved 18 June 2024 – via GeoScienceWorld.
  488. ^ Van de Schootbrugge, Bas; Wignall, Paul Barry (26 October 2015). "A tale of two extinctions: converging end-Permian and end-Triassic scenarios". Geological Magazine. 153 (2): 332–354. doi:10.1017/S0016756815000643. hdl:1874/329922. S2CID 131750128. Retrieved 26 May 2023.
  489. ^ Benton, Michael James (3 September 2018). "Hyperthermal-driven mass extinctions: killing models during the Permian–Triassic mass extinction". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 376 (2130): 1–19. Bibcode:2018RSPTA.37670076B. doi:10.1098/rsta.2017.0076. PMC 6127390. PMID 30177561.
  490. ^ a b Foster, C. B.; Afonin, S. A. (July 2005). "Abnormal pollen grains: an outcome of deteriorating atmospheric conditions around the Permian–Triassic boundary". Journal of the Geological Society. 162 (4): 653–659. Bibcode:2005JGSoc.162..653F. doi:10.1144/0016-764904-047. S2CID 128829042. Retrieved 26 May 2023.
  491. ^ Li, Rucao; Shen, Shu-Zhong; Xia, Xiao-Ping; Xiao, Bing; Feng, Yuzhou; Chen, Huayong (5 March 2024). "Atmospheric ozone destruction and the end-Permian crisis: Evidence from multiple sulfur isotopes". Chemical Geology. 647: 121936. doi:10.1016/j.chemgeo.2024.121936. Retrieved 21 May 2024 – via Elsevier Science Direct.
  492. ^ Dal Corso, Jacopo; Newton, Robert J.; Zerkle, Aubrey L.; Chu, Daoliang; Song, Haijun; Song, Huyue; Tian, Li; Tong, Jinnan; Di Rocco, Tommaso; Claire, Mark W.; Mather, Tamsin A.; He, Tianchen; Gallagher, Timothy; Shu, Wenchao; Wu, Yuyang; Bottrell, Simon H.; Metcalfe, Ian; Cope, Helen A.; Novak, Martin; Jamieson, Robert A.; Wignall, Paul Barry (2 September 2024). "Repeated pulses of volcanism drove the end-Permian terrestrial crisis in northwest China". Nature Communications. 15 (1): 7628. doi:10.1038/s41467-024-51671-5. ISSN 2041-1723. PMC 11368959. PMID 39223125.
  493. ^ Lamarque, J.-F.; Kiehl, J. T.; Shields, C. A.; Boville, B. A.; Kinnison, D. E. (9 August 2006). "Modeling the response to changes in tropospheric methane concentration: Application to the Permian-Triassic boundary". Paleoceanography and Paleoclimatology. 21 (3): 1–15. Bibcode:2006PalOc..21.3006L. doi:10.1029/2006PA001276.
  494. ^ Beerling, David J; Harfoot, Michael; Lomax, Barry; Pyle, John A (15 July 2007). "The stability of the stratospheric ozone layer during the end-Permian eruption of the Siberian Traps". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 365 (1856): 1843–1866. doi:10.1098/rsta.2007.2046. ISSN 1364-503X. Retrieved 18 June 2024.
  495. ^ Black, Benjamin A.; Lamarque, Jean-François; Shields, Christine A.; Elkins-Tanton, Linda T.; Kiehl, Jeffrey T. (1 January 2014). "Acid rain and ozone depletion from pulsed Siberian Traps magmatism". Geology. 42 (1): 67–70. Bibcode:2014Geo....42...67B. doi:10.1130/G34875.1. Retrieved 31 May 2023.
  496. ^ Lamarque, J.-F.; Kiehl, J. T.; Orlando, J. J. (16 January 2007). "Role of hydrogen sulfide in a Permian-Triassic boundary ozone collapse". Geophysical Research Letters. 34 (2): 1–4. Bibcode:2007GeoRL..34.2801L. doi:10.1029/2006GL028384. S2CID 55812439.
  497. ^ Kaiho, Kunio; Koga, Seizi (August 2013). "Impacts of a massive release of methane and hydrogen sulfide on oxygen and ozone during the late Permian mass extinction". Global and Planetary Change. 107: 91–101. doi:10.1016/j.gloplacha.2013.04.004. Retrieved 18 June 2024 – via Elsevier Science Direct.
  498. ^ Kump, Lee; Pavlov, Alexander; Arthur, Michael A. (1 May 2005). "Massive release of hydrogen sulfide to the surface ocean and atmosphere during intervals of oceanic anoxia". Geology. 33 (5): 397–400. Bibcode:2005Geo....33..397K. doi:10.1130/G21295.1. Retrieved 2 April 2023.
  499. ^ Ward, Peter Douglas (17 April 2007). "The Mother of All Extinctions". Under a Green Sky: Global Warming, the Mass Extinctions of the Past, and What They Can Tell Us About Our Future. New York: HarperCollins. pp. 61–86. ISBN 978-0-06-113791-4.
  500. ^ Algeo, Thomas J; Shen, Jun (8 September 2023). "Theory and classification of mass extinction causation". National Science Review. 11 (1): nwad237. doi:10.1093/nsr/nwad237. ISSN 2095-5138. PMC 10727847. PMID 38116094.
  501. ^ Retallack GJ, Seyedolali A, Krull ES, Holser WT, Ambers CP, Kyte FT (1998). "Search for evidence of impact at the Permian–Triassic boundary in Antarctica and Australia". Geology. 26 (11): 979–982. Bibcode:1998Geo....26..979R. doi:10.1130/0091-7613(1998)026<0979:SFEOIA>2.3.CO;2.
  502. ^ a b Becker L, Poreda RJ, Basu AR, Pope KO, Harrison TM, Nicholson C, Iasky R (2004). "Bedout: a possible end-Permian impact crater offshore of northwestern Australia". Science. 304 (5676): 1469–1476. Bibcode:2004Sci...304.1469B. doi:10.1126/science.1093925. PMID 15143216. S2CID 17927307.
  503. ^ Becker L, Poreda RJ, Hunt AG, Bunch TE, Rampino M (2001). "Impact event at the Permian–Triassic boundary: Evidence from extraterrestrial noble gases in fullerenes". Science. 291 (5508): 1530–1533. Bibcode:2001Sci...291.1530B. doi:10.1126/science.1057243. PMID 11222855. S2CID 45230096.
  504. ^ Basu AR, Petaev MI, Poreda RJ, Jacobsen SB, Becker L (2003). "Chondritic meteorite fragments associated with the Permian–Triassic boundary in Antarctica". Science. 302 (5649): 1388–1392. Bibcode:2003Sci...302.1388B. doi:10.1126/science.1090852. PMID 14631038. S2CID 15912467.
  505. ^ Kaiho K, Kajiwara Y, Nakano T, Miura Y, Kawahata H, Tazaki K, Ueshima M, Chen Z, Shi GR (2001). "End-Permian catastrophe by a bolide impact: Evidence of a gigantic release of sulfur from the mantle". Geology. 29 (9): 815–818. Bibcode:2001Geo....29..815K. doi:10.1130/0091-7613(2001)029<0815:EPCBAB>2.0.CO;2. ISSN 0091-7613.
  506. ^ Farley KA, Mukhopadhyay S, Isozaki Y, Becker L, Poreda RJ (2001). "An extraterrestrial impact at the Permian–Triassic boundary?". Science. 293 (5539): 2343a–2343. doi:10.1126/science.293.5539.2343a. PMID 11577203.
  507. ^ Koeberl C, Gilmour I, Reimold WU, Philippe Claeys P, Ivanov B (2002). "End-Permian catastrophe by bolide impact: Evidence of a gigantic release of sulfur from the mantle: Comment and Reply". Geology. 30 (9): 855–856. Bibcode:2002Geo....30..855K. doi:10.1130/0091-7613(2002)030<0855:EPCBBI>2.0.CO;2. ISSN 0091-7613.
  508. ^ Isbell JL, Askin RA, Retallack GR (1999). "Search for evidence of impact at the Permian–Triassic boundary in Antarctica and Australia; discussion and reply". Geology. 27 (9): 859–860. Bibcode:1999Geo....27..859I. doi:10.1130/0091-7613(1999)027<0859:SFEOIA>2.3.CO;2.
  509. ^ Koeberl K, Farley KA, Peucker-Ehrenbrink B, Sephton MA (2004). "Geochemistry of the end-Permian extinction event in Austria and Italy: No evidence for an extraterrestrial component". Geology. 32 (12): 1053–1056. Bibcode:2004Geo....32.1053K. doi:10.1130/G20907.1.
  510. ^ Langenhorst F, Kyte FT, Retallack GJ (2005). "Reexamination of quartz grains from the Permian–Triassic boundary section at Graphite Peak, Antarctica" (PDF). Lunar and Planetary Science Conference XXXVI. Retrieved 2007-07-13.
  511. ^ Zhou, Lei; Kyte, Frank T. (25 November 1988). "The Permian-Triassic boundary event: a geochemical study of three Chinese sections". Earth and Planetary Science Letters. 90 (4): 411–421. Bibcode:1988E&PSL..90..411L. doi:10.1016/0012-821X(88)90139-2. Retrieved 31 May 2023.
  512. ^ Jones AP, Price GD, Price NJ, DeCarli PS, Clegg RA (2002). "Impact induced melting and the development of large igneous provinces". Earth and Planetary Science Letters. 202 (3): 551–561. Bibcode:2002E&PSL.202..551J. CiteSeerX 10.1.1.469.3056. doi:10.1016/S0012-821X(02)00824-5.
  513. ^ AHager, Bradford H. (2001). "Giant Impact Craters Lead To Flood Basalts: A Viable Model". CCNet 33/2001: Abstract 50470. Archived from the original on 2008-04-22. Retrieved 2008-04-06.
  514. ^ Hagstrum, Jonathan T. (2001). "Large Oceanic Impacts As The Cause Of Antipodal Hotspots And Global Mass Extinctions". CCNet 33/2001: Abstract 50288. Archived from the original on 2008-04-22. Retrieved 2008-04-06.
  515. ^ Romano, Marco; Bernardi, Massimo; Petti, Fabio Massimo; Rubidge, Bruce; Hancox, John; Benton, Michael James (November 2020). "Early Triassic terrestrial tetrapod fauna: a review". Earth-Science Reviews. 210: 103331. Bibcode:2020ESRv..21003331R. doi:10.1016/j.earscirev.2020.103331. S2CID 225066013. Retrieved 4 January 2023.
  516. ^ Burger, Benjamin J.; Vargas Estrada, Margarita; Gustin, Mae Sexauer (June 2019). "What caused Earth's largest mass extinction event? New evidence from the Permian-Triassic boundary in northeastern Utah". Global and Planetary Change. 177: 81–100. Bibcode:2019GPC...177...81B. doi:10.1016/j.gloplacha.2019.03.013. S2CID 134324242. Retrieved 2 April 2023.
  517. ^ Frese, Ralph R. B. von; Potts, Laramie V.; Wells, Stuart B.; Gaya-Piqué, Luis-Ricardo; Golynsky, Alexander V.; Hernandez, Orlando; Kim, Jeong Woo; Kim, Hyung Rae; Hwang, Jong Sun (2006). "Permian–Triassic mascon in Antarctica". American Geophysical Union, Fall Meeting 2007. 2007: Abstract T41A–08. Bibcode:2006AGUSM.T41A..08V.
  518. ^ Frese, Ralph R. B. von; Potts, Laramie V.; Wells, Stuart B.; Leftwich, Timothy E.; Kim, Hyung Rae; Kim, Jeong Woo; Golynsky, Alexander V.; Hernandez, Orlando; Gaya-Piqué, Luis-Ricardo (25 February 2009). "GRACE gravity evidence for an impact basin in Wilkes Land, Antarctica". Geochemistry, Geophysics, Geosystems. 10 (2). Bibcode:2009GGG....10.2014V. doi:10.1029/2008GC002149. ISSN 1525-2027.
  519. ^ a b "Biggest extinction in history caused by climate-changing meteor". University News (Press release). University of Western Australia. 31 July 2013.
  520. ^ Rocca, M.; Rampino, M.; Baez Presser, J. (2017). "Geophysical evidence for a la impact structure on the Falkland (Malvinas) Plateau". Terra Nova. 29 (4): 233–237. Bibcode:2017TeNov..29..233R. doi:10.1111/ter.12269. S2CID 134484465.
  521. ^ McCarthy, Dave; Aldiss, Don; Arsenikos, Stavros; Stone, Phil; Richards, Phil (24 August 2017). "Comment on "Geophysical evidence for a large impact structure on the Falkland (Malvinas) Plateau"" (PDF). Terra Nova. 29 (6): 411–415. Bibcode:2017TeNov..29..411M. doi:10.1111/ter.12285. ISSN 0954-4879. S2CID 133781924. Retrieved 26 March 2023.
  522. ^ Shen, Shu-Zhong; Bowring, Samuel A. (2014). "The end-Permian mass extinction: A still unexplained catastrophe". National Science Review. 1 (4): 492–495. doi:10.1093/nsr/nwu047.
  523. ^ Gribbin, John (2012). Alone in the Universe: Why our Planet is Unique. Hoboken, NJ: John Wiley & Sons. pp. 72–73. ISBN 978-1-118-14797-9.
  524. ^ Payne, Jonathan L.; Clapham, Matthew E. (May 2012). "End-Permian Mass Extinction in the Oceans: An Ancient Analog for the Twenty-First Century?". Annual Review of Earth and Planetary Sciences. 40 (1): 89–111. Bibcode:2012AREPS..40...89P. doi:10.1146/annurev-earth-042711-105329. Retrieved 31 May 2023.
  525. ^ Cariglino, Bárbara; Moisan, Philippe; Lara, María Belén (November 2021). "The fossil record of plant-insect interactions and associated entomofaunas in Permian and Triassic floras from southwestern Gondwana: A review and future prospects". Journal of South American Earth Sciences. 111: 103512. doi:10.1016/j.jsames.2021.103512. Retrieved 18 June 2024 – via Elsevier Science Direct.
  526. ^ "Greatest mass extinction driven by acidic oceans, study finds". ScienceDaily. Retrieved 2024-11-24.
  527. ^ Brannen, Peter (July 29, 2017). "Opinion | When Life on Earth Was Nearly Extinguished". The New York Times – via NYTimes.com.

Further reading

[edit]
[edit]