انقراض العصر البرمي-الثلاثي

حدث انقراض العصر البرمي-الثلاثي (إنگليزية: Permian–Triassic extinction event)، ويُعرف أيضاً باسم حدث انقراض العصر البرمي المتأخر (إنگليزية: Late Permian extinction event[2] [3] حدث انقراض نهاية العصر البرمي (إنگليزية: End-Permian extinction event[4][5] ويُعرف بشكل عام باسم الموت العظيم،[6][7] forms the boundary between the Permian and Triassic geologic periods, and with them the Paleozoic and Mesozoic eras respectively, approximately 251.9 million years ago.[8] As the largest of the "Big Five" mass extinctions of the Phanerozoic,[9] it is the Earth's most severe known extinction event,[10][11] with the extinction of 57% of biological families, 83% of genera, 81% of marine species[12][13][14] and 70% of terrestrial vertebrate species.[15] It is the largest known mass extinction of insects.[16] There is evidence for one to three distinct pulses, or phases, of extinction.[17][15][18][19][20]

حدود العصر البرمي-الترياسي في شاطئ فريزر في نيو ساوث ويلز، مع حدث انقراض نهاية العصر البرمي الواقع فوق طبقة الفحم مباشرة.[1]

The scientific consensus is that the main cause of extinction was the large amount of carbon dioxide emitted by the flood basalt volcanic eruptions that created the Siberian Traps, which elevated global temperatures and acidified the oceans.[21] Proposed contributing factors include: the emission of much additional carbon dioxide from the thermal decomposition of hydrocarbon deposits, including oil and coal, triggered by the eruptions; and emissions of methane by novel methanogenic microorganisms, perhaps nourished by minerals dispersed in the eruptions.[22][23]

The speed of recovery from the extinction is disputed. Some scientists estimate that it took 10 million years (until the Middle Triassic), due both to the severity of the extinction and because grim conditions returned periodically over the course of the Early Triassic,[24][25][26] causing further extinction events, such as the Smithian-Spathian boundary extinction.[11][27] However, studies in Bear Lake County, near Paris, Idaho,[28] and nearby sites in Idaho and Nevada[29] showed a relatively quick rebound in a localized Early Triassic marine ecosystem, taking around 3 million years to recover, while an unusually diverse and complex ichnobiota is known from Italy less than a million years after the end-Permian extinction.[30] Additionally, the complex Guiyang biota found near Guiyang, China also indicates life thrived in some places just a million years after the mass extinction,[31][32] as does a fossil assemblage known as the Shanggan fauna found in Shanggan, China[33] and a gastropod fauna from the Al Jil Formation of Oman.[34] Regional differences in the pace of biotic recovery suggest that the impact of the extinction may have been felt less severely in some areas than others, with differential environmental stress and instability.[35] 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;[36] 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.[37]

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

التأريخ

قالب:P-T extinction graphical timeline 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.[38] However, it is now possible to date the extinction with millennial precision. U–Pb zircon dates from five volcanic ash beds from the Global Stratotype Section and Point for the Permian–Triassic boundary at Meishan, China, establish a high-resolution age model for the extinction – allowing exploration of the links between global environmental perturbation, carbon cycle disruption, mass extinction, and recovery at millennial timescales.

The extinction occurred between 251.941 ± 0.037 and 251.880 ± 0.031 million years ago, a duration of 60 ± 48 thousand years.[39] A large (approximately 0.9%), abrupt global decrease in the ratio of the stable isotope carbon-13 to that of carbon-12 coincides with this extinction,[40][41][42][43][44] and is sometimes used to identify the Permian–Triassic boundary in rocks that are unsuitable for radiometric dating.[43]

Further evidence for environmental change around the P–Tr boundary suggests an 8 °C (14 °F) rise in temperature,[40] and an increase in CO 2 levels by 2,000 ppm (for comparison, the concentration immediately before the Industrial Revolution was 280 ppm,[40] and the amount today is about 415 ppm[45]). There is also evidence of increased ultraviolet radiation reaching the earth, causing the mutation of plant spores.[40][46]

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.[47] For a while this "fungal spike" was used by some paleontologists to identify the Permian–Triassic boundary in rocks that are unsuitable for radiometric dating or have a lack of suitable index fossils, but even the proposers of the fungal spike hypothesis pointed out that "fungal spikes" may have been a repeating phenomenon created by the post-extinction ecosystem during the earliest Triassic.[47] 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;[40][48] the spike did not appear worldwide;[49][50][51] and in many places it did not fall on the Permian–Triassic boundary.[52] 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.[53] Newer chemical evidence agrees better with a fungal origin for Reduviasporonites, diluting these critiques.[54][55]

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[15] or that the extinction was long and spread out over a few million years, with a sharp peak in the last million years of the Permian.[52][56] 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,[57] and a study of the Shangsi section showed two extinction pulses with different causes too.[58] 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.[59] 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.[60] 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.[61]

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.[62] Other scientists believe the terrestrial mass extinction began between 60 and 370 thousand years before the onset of the marine mass extinction.[63]

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.[64][65] 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][66] posits that there were two major extinction pulses 9.4 million years apart, separated by a period of extinctions well above the background level, and that the final extinction killed off only about 80% of marine species alive at that time while the other losses occurred during the first pulse or the interval between pulses. According to this theory, one of these extinction pulses occurred at the end of the Guadalupian epoch of the Permian.[67][15][68] For example, all dinocephalian genera died out at the end of the Guadalupian,[66] as did the Verbeekinidae, a family of large-size fusuline foraminifera.[69] 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.[70][71]


أنماط الانقراض

انقراض بحري الأجناس المنقرضة ملاحظات
مفصليات الأرجل
Eurypterids 100% May have become extinct shortly before the P–Tr boundary
Ostracods 59%  
Trilobites 100% In decline since the Devonian; only 2 genera living before the extinction
Brachiopoda
Brachiopods 96% Orthids and productids died out
Bryozoa
Bryozoans 79% Fenestrates, trepostomes, and cryptostomes died out
Chordata
Acanthodians 100% In decline since the Devonian, with only one living family
Cnidaria
Anthozoans 96% Tabulate and rugose corals died out
Echinodermata
Blastoids 100% May have become extinct shortly before the P–Tr boundary
Crinoids 98% Inadunates and camerates died out
Mollusca
Ammonites 97% Goniatites died out
Bivalves 59%  
Gastropods 98%  
Retaria
Foraminiferans 97% Fusulinids died out, but were almost extinct before the catastrophe
Radiolarians 99%[72]

العضيات البحرية

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.[73]

The extinction primarily affected organisms with calcium carbonate skeletons, especially those reliant on stable CO2 levels to produce their skeletons.[74] These organisms were susceptible to the effects of the ocean acidification that resulted from increased atmospheric CO2. There is also evidence that endemism was a strong risk factor influencing a taxon's likelihood of extinction. Bivalve taxa that were endemic and localised to a specific region were more likely to go extinct than cosmopolitan taxa.[75] There was little latitudinal difference in the survival rates of taxa.[76]

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).[77][78] The extinction rate of marine organisms was catastrophic.[18][38][79][62] Bioturbators were extremely severely affected, as evidenced by the loss of the sedimentary mixed layer in many marine facies during the end-Permian extinction.[80]

Surviving marine invertebrate groups included articulate brachiopods (those with a hinge),[81] which had undergone a slow decline in numbers since the P–Tr extinction; the Ceratitida order of ammonites;[82] and crinoids ("sea lilies"),[82] 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.[83][84] In the case of the brachiopods, at least, surviving taxa were generally small, rare members of a formerly diverse community.[85]

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.[86] 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.[87]

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.[88] 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.[89]

Cladodontomorph sharks likely survived the extinction because of their ability to survive in refugia in the deep oceans. This hypothesis is based on the discovery of Early Cretaceous cladodontomorphs in deep, outer shelf environments.[90]

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,[91][92] notably occurring in foraminifera,[93][94] brachiopods,[95][96][97] bivalves,[98][99][100] and ostracods.[101] Though gastropods that survived the cataclysm were smaller in size than those that did not,[102] it remains debated whether the Lilliput effect truly took hold among gastropods.[103][104][105] Some gastropod taxa, termed "Gulliver gastropods", ballooned in size during and immediately following the mass extinction,[106] exemplifying the Lilliput effect's opposite, which has been dubbed the Brobdingnag effect.[107]

اللافقاريات البرية

The Permian had great diversity in insect and other invertebrate species, including the largest insects ever to have existed. The end-Permian is the largest known mass extinction of insects;[16] according to some sources, it may well be the only mass extinction to significantly affect insect diversity.[108][109] 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.[38]

النباتات البرية

The geological record of terrestrial plants is sparse and based mostly on pollen and spore studies. Plants are relatively immune to mass extinction, with the impact of all the major mass extinctions "insignificant" at a family level.[40][محل شك] Even the reduction observed in species diversity (of 50%) may be mostly due to taphonomic processes.[110][40] However, a massive rearrangement of ecosystems does occur, with plant abundances and distributions changing profoundly and all the forests virtually disappearing.[111][40] The dominant floral groups changed, with many groups of land plants entering abrupt decline, such as Cordaites (gymnosperms) and Glossopteris (seed ferns);[112][113] the Palaeozoic flora scarcely survived this extinction.[114]

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.[115]

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.[51]

The Cordaites flora, which dominated the Angaran floristic realm corresponding to Siberia, collapsed over the course of the extinction.[116] 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.[117]

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.[118]

In South China, the subtropical Cathaysian gigantopterid dominated rainforests abruptly collapsed.[119][120][121] 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.[122]

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.[123]

الفقاريات البرية

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).[124][10]

The groups that survived suffered extremely heavy losses of species and some terrestrial vertebrate groups very nearly became extinct at the end of the Permian. Some of the surviving groups did not persist for long past this period, but others that barely survived went on to produce diverse and long-lasting lineages. However, it took 30 million years for the terrestrial vertebrate fauna to fully recover both numerically and ecologically.[125]

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.[126]


. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Biotic recovery

In the wake of the extinction event, the ecological structure of present-day biosphere evolved from the stock of surviving taxa. In the sea, the "Palaeozoic Evolutionary Fauna" declined while the "Modern Evolutionary Fauna" achieved greater dominance;[127] the Permian-Triassic mass extinction marked a key turning point in this ecological shift that began after the Capitanian mass extinction[128] and culminated in the Late Jurassic.[129] Typical taxa of shelly benthic faunas were now bivalves, snails, sea urchins and Malacostraca, whereas bony fishes[130] and marine reptiles[131] 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).[132][133] However, recovery was also differential between taxa. Some survivors became extinct some million years after the extinction event without having rediversified (dead clade walking,[134] e.g. the snail family Bellerophontidae),[135] whereas others rose to dominance over geologic times (e.g., bivalves).[136][137]

الأنظمة البيئية البحرية

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

Marine post-extinction faunas were mostly species-poor and were dominated by few disaster taxa such as the bivalves Claraia, Unionites, Eumorphotis, and Promyalina,[138] the foraminifera Rectocornuspira kalhori and Earlandia,[139] and the inarticulate brachiopod Lingularia.[138] Their guild diversity was also low.[140] Seafloor communities maintained a comparatively low diversity until the end of the Early Triassic, approximately 4 million years after the extinction event.[141] 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,[142] and conodonts, which diversified considerably over the first two million years of the Early Triassic.[143] The relative delay in the recovery of benthic organisms has been attributed to widespread anoxia,[144] but high abundances of benthic species contradict this explanation.[145] Epifaunal benthos took longer to recover than infaunal benthos.[146] More recent work suggests that the pace of recovery was intrinsically driven by the intensity of competition among species, which drives rates of niche differentiation and speciation.[147] That recovery was slow in the Early Triassic can be explained by low levels of biological competition due to the paucity of taxonomic diversity,[148] 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.[149] Accordingly, low levels of interspecific competition in seafloor communities that are dominated by primary consumers correspond to slow rates of diversification and high levels of interspecific competition among nektonic secondary and tertiary consumers to high diversification rates. Other studies point to the continual episodes of extremely hot climatic conditions during the Early Triassic for the delayed recovery of oceanic life,[150][151] in particular skeletonised taxa that are most vulnerable to high carbon dioxide concentrations.[152] The 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.[35] Recurrent Early Triassic environmental stresses also acted as a ceiling limiting the maximum ecological complexity of marine ecosystems until the Spathian.[153] Recovery biotas appear to have been ecologically uneven and unstable into the Anisian, making them vulnerable to environmental stresses.[154] Whereas most marine communities were fully recovered by the Middle Triassic,[155][156] global marine diversity reached pre-extinction values no earlier than the Middle Jurassic, approximately 75 million years after the extinction event.[157]

 
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.[158] 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,[158] and the increase in predation pressure led to the Mesozoic Marine Revolution.

Bivalves rapidly recolonised many marine environments in the wake of the catastrophe.[159] 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,[160] and one group, the rudist clams, became the Mesozoic's main reef-builders. 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 competed with.[129] 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.[161] Some researchers think the change was attributable not only to the end-Permian extinction but also the ecological restructuring that began as a result of the Capitanian extinction.[162]

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.[163] 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.[164]

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.[165]

Crinoids ("sea lilies") suffered a selective extinction, resulting in a decrease in the variety of their forms.[166] 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.[167] 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.[168] Though their taxonomic diversity remained relatively low, crinoids regained much of their ecological dominance by the Middle Triassic epoch.[161]

Microbial-metazoan reefs dominated surviving communities in the immediate wake of the mass extinction. Metazoan-built reefs reemerged during the Olenekian, mainly being composed of sponge biostrome and bivalve builups. "Tubiphytes"-dominated reefs appeared at the end of the Olenekian, representing the earliest platform-margin reefs of the Triassic, though they did not become abundant until the late Anisian, when reefs' species richness increased. The first scleractinian corals appear in the late Anisian as well, although they would not become the dominant reef builders until the end of the Triassic period.[169] Metazoan reefs became common again during the Anisian because the oceans cooled down then from their overheated state during the Early Triassic.[170]

Ichnocoenoses show that marine ecosystems recovered to pre-extinction levels of ecological complexity by the late Olenekian.[171] 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.[172]

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.[173][174] 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.[175] 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.[176]

النباتات البرية

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.[177] Dominant gymnosperm genera were replaced post-boundary by lycophytes – extant lycophytes are recolonizers of disturbed areas.[178] 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.[116] The worldwide recovery of gymnosperm forests took approximately 4–5 million years.[40] 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,[179] and floral recovery likely didn't follow a congruent, globally universal trend but instead varied by region according to local environmental conditions.[121]

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 قالب:Delta values but occurred many years after.[51]

In what is now the Barents Sea of the coast of Norway, the post-extinction fauna is dominated by pteridophytes and lycopods, which were suited for primary succession and recolonisation of devastated areas, although gymnosperms made a rapid recovery, with the lycopod dominated flora not persisting across most of the Early Triassic as postulated in other regions.[179]

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.[180]

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.[115]

The gigantopterid-dominated forests of South China were replaced by low-lying herbaceous vegetation dominated by the isoetalean Tomiostrobus following the collapse of the former. In contrast to the highly biodiverse gigantopterid rainforests, the post-extinction landscape of South China was near-barren and had vastly lower diversity.[121]


. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

فجوة الفحم

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.[181] Abiotic factors (factors not caused by organisms), such as decreased rainfall or increased input of clastic sediments, may also be to blame.[40]

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.[40] For example, in eastern Australia a cold climate had been the norm for a long period, with a peat mire ecosystem adapted to these conditions. Approximately 95% of these peat-producing plants went locally extinct at the P–Tr boundary;[182] coal deposits in Australia and Antarctica disappear significantly before the P–Tr boundary.[40]

الفقاريات البرية

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] Overall, terrestrial faunas after the extinction event tended to be more variable and heterogeneous across space than those of the Late Permian, which exbibited less provincialism, being much more geographically homogeneous.[183]

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.[184]

 
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. Ichnological evidence from the earliest Triassic of the Karoo Basin shows dicynodonts were abundant in the immediate aftermath of the biotic crisis.[185] Smaller carnivorous cynodont therapsids also survived, including the ancestors of mammals. In the Karoo region of southern Africa, the therocephalians Tetracynodon, Moschorhinus and Ictidosuchoides survived, but do not appear to have been abundant in the Triassic.[186] 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.[187][188]

Archosaurs (which included the ancestors of dinosaurs and crocodilians) were initially rarer than therapsids, but they began to displace therapsids in the mid-Triassic. In the mid to late Triassic, the dinosaurs evolved from one group of archosaurs, and went on to dominate terrestrial ecosystems during the Jurassic and Cretaceous.[189] 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,[190] 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.[191] The archosaur dominance would end again due to the Cretaceous–Paleogene extinction event, after which both birds (only extant dinosaurs) and mammals (only extant synapsids) would diversify and share the world.

Some temnospondyl amphibians made a relatively quick recovery, in spite of nearly becoming extinct. Mastodonsaurus and trematosaurians were the main aquatic and semiaquatic predators during most of the Triassic, some preying on tetrapods and others on fish.[192]

اللافقاريات البرية

Most fossil insect groups found after the Permian–Triassic boundary differ significantly from those before: Of Paleozoic insect groups, only the Glosselytrodea, Miomoptera, and Protorthoptera have been discovered in deposits from after the extinction. The caloneurodeans, monurans, paleodictyopteroids, protelytropterans, and protodonates became extinct by the end of the Permian. Though Triassic insects are very different from those of the Permian, a gap in the insect fossil record spans approximately 15 million years from the late Permian to early Triassic. In well-documented Late Triassic deposits, fossils overwhelmingly consist of modern fossil insect groups.[108]

Microbially induced sedimentary structures

Microbially induced sedimentary structures (MISS) dominated North Chinese terrestrial fossil assemblages in the Early Triassic.[193][194] In Arctic Canada as well, MISS became a common occurrence following the Permian-Triassic extinction.[195] 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.[194]

فرضيات حول السبب

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

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

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

Any hypothesis about the cause must explain the selectivity of the event, which affected organisms with calcium carbonate skeletons most severely; the long period (4 to 6 million years) before recovery started, and the minimal extent of biological mineralization (despite inorganic carbonates being deposited) once the recovery began.[74]

نشاط بركاني

الفخاخ السيبيرية

The flood basalt eruptions that produced the Siberian Traps constituted one of the largest known volcanic events on Earth and covered over 2,000,000 square kilometres (770,000 sq mi) with lava.[196][197][198] Such a vast aerial extent of the flood basalts may have contributed to their exceptionally catastrophic impact.[199] The date of the Siberian Traps eruptions and the extinction event are in good agreement.[39][200] The timeline of the extinction event strongly indicates it was caused by events in the large igneous province of the Siberian Traps.[201][202][203][204] A study of the Norilsk and Maymecha-Kotuy regions of the northern Siberian platform indicates that volcanic activity occurred during a small number of high intensity pulses that exuded enormous volumes of magma, as opposed to flows emplaced at regular intervals.[205] The rate of carbon dioxide release from the Siberian Traps represented one of the most rapid rises of carbon dioxide levels in the geologic record,[206] with the rate of carbon dioxide emissions being estimated by one study to be five times faster than the rate during the already catastrophic Capitanian mass extinction event,[207] which occurred as a result of the activity of the Emeishan Traps in southwestern China at the end of the Middle Permian.[208][209][210] Carbon dioxide levels prior to and after the eruptions are poorly constrained, but may have jumped from between 500 and 4,000 PPM prior to the extinction event to around 8,000 PPM after the extinction.[211] As carbon dioxide levels shot up, extreme temperature rise would have followed.[212] 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.[213][8] Further evidence – based on paired coronene-mercury spikes – for a volcanic combustion cause of the mass extinction was published in 2020.[214][22]

The Siberian Traps had unusual features that made them even more dangerous. The Siberian lithosphere is significantly enriched in halogens, whose properties render them extremely destructive to the ozone layer, and evidence from subcontinental lithospheric xenoliths indicates that as much as 70% of the halogen content was released into the atmosphere from sections of the lithosphere intruded into by the Siberian Traps.[215] The Siberian Traps eruptions released sulphur-rich volatiles that caused dust clouds and the formation of acid aerosols, which would have blocked out sunlight and thus disrupted photosynthesis both on land and in the photic zone of the ocean, causing food chains to collapse. These volcanic outbursts of sulphur also induced brief but severe global cooling that interrupted the broader trend of rapid global warming,[216] leading to glacio-eustatic sea level fall.[215][217] The eruptions may also have caused acid rain as the aerosols washed out of the atmosphere. That may have killed land plants and mollusks and planktonic organisms which had calcium carbonate shells. Pure flood basalts produce fluid, low-viscosity lava, and do not hurl debris into the atmosphere. It appears, however, that 20% of the output of the Siberian Traps eruptions was pyroclastic (consisted of ash and other debris thrown high into the atmosphere), increasing the short-term cooling effect.[218] When all of the dust and ash clouds and aerosols washed out of the atmosphere, the excess carbon dioxide emitted by the Siberian Traps would have remained and global warming would have proceeded without any mitigating effects.[212]

The Siberian Traps are underlain by thick sequences of Early-Mid Paleozoic aged carbonate and evaporite deposits, as well as Carboniferous-Permian aged coal bearing clastic rocks. When heated, such as by igneous intrusions, these rocks are capable of emitting large amounts of greenhouse and toxic gases. The unique setting of the Siberian Traps over these deposits is likely the reason for the severity of the extinction.[219][220] The basalt lava erupted or intruded into carbonate rocks and into sediments that were in the process of forming large coal beds, both of which would have emitted large amounts of carbon dioxide, leading to stronger global warming after the dust and aerosols settled.[212] The timing of the change of the Siberian Traps from flood basalt dominated emplacement to sill dominated emplacement, the latter of which would have liberated the largest amounts of trapped hydrocarbon deposits, coincides with the onset of the main phase of the mass extinction[221] and is linked to a major negative قالب:Delta excursion.[222] Venting of coal-derived methane was not the only mechanism of carbon release; there is also extensive evidence of explosive combustion of coal and discharge of coal-fly ash.[21] A 2011 study led by Stephen E. Grasby reported evidence that volcanism caused massive coal beds to ignite, possibly releasing more than 3 trillion tons of carbon. The team found ash deposits in deep rock layers near what is now the Buchanan Lake Formation. According to their article, "coal ash dispersed by the explosive Siberian Trap eruption would be expected to have an associated release of toxic elements in impacted water bodies where fly ash slurries developed. ... Mafic megascale eruptions are long-lived events that would allow significant build-up of global ash clouds."[223][224] In a statement, Grasby said, "In addition to these volcanoes causing fires through coal, the ash it spewed was highly toxic and was released in the land and water, potentially contributing to the worst extinction event in earth history."[225] A 2013 study led by Q.Y. Yang reported that the total amounts of important volatiles emitted from the Siberian Traps consisted of 8.5 × 107 Tg CO2, 4.4 × 106 Tg CO, 7.0 × 106 Tg H2S, and 6.8 × 107 Tg SO2. The data support a popular notion that the end-Permian mass extinction on the Earth was caused by the emission of enormous amounts of volatiles from the Siberian Traps into the atmosphere.[226]

Mercury anomalies corresponding to the time of Siberian Traps activity have been found in many geographically disparate sites,[227] evidencing that these volcanic eruptions released significant quantities of toxic mercury into the atmosphere and ocean, causing even further large scale die-offs of terrestrial and marine life.[228][229][230] A series of surges in mercury emissions raised environmental mercury concentrations to levels orders of magnitude above normal background levels and caused intervals of extreme environmental toxicity, each lasting for over a thousand years, in both terrestrial and marine ecosystems.[231] Immense volumes of nickel aerosols were also released by Siberian Traps volcanic activity,[232][233] further contributing to metal poisoning.[234] Cobalt and arsenic emissions from the Siberian Traps caused further still environmental stress.[228] A major volcanogenic influx of isotopically light zinc from the Siberian Traps has also been recorded.[235]

Choiyoi Silicic Large Igneous Province

A second flood basalt event that emplaced what is now known as the Choiyoi Silicic Large Igneous Province in southwestern Gondwana between around 286 Ma and 247 Ma has also been suggested as a possible extinction mechanism.[115] Being about 1,300,000 cubic kilometres in volume[236] and 1,680,000 square kilometres in area, this flood basalt event was approximately 40% the size of the Siberian Traps and thus may have been a significant additional factor explaining the severity of the end-Permian extinction.[115]

القوس البركاني لمنطقة اندساس الهند الصينية-جنوب الصين

Mercury anomalies preceding the end-Permian extinction have been discovered in what was then the boundary between the South China Craton and the Indochinese plate, which was home to a subduction zone and a corresponding volcanic arc. Hafnium isotopes from syndepositional magmatic zircons found in ash beds created by this pulse of volcanic activity confirm its origin in subduction-zone volcanism rather than large igneous province activity.[237] The enrichment of copper samples from these deposits in isotopically light copper provide additional confirmation of the felsic nature of this volcanism and that its origin was not a large igneous province.[238] This volcanism has been speculated to have caused local episodes of biotic stress among radiolarians, sponges, and brachiopods that took place over the 60,000 years preceding the end-Permian marine extinction, as well as an ammonoid crisis manifested in their decreased morphological complexity and size and their increased rate of turnover that began in the lower C. yini biozone, around 200,000 years prior to the end-Permian extinction.[237]

غزغزة كلاثرات الميثان

Methane clathrates, also known as methane hydrates, consist of methane molecules trapped in cages of water molecules. The methane, produced by methanogens (microscopic single-celled organisms), has a 13 C 12 C ratio about 6.0% below normal (قالب:Delta −6.0%). At the right combination of pressure and temperature, the methane is trapped in clathrates fairly close to the surface of permafrost and, in much larger quantities, on continental shelves and the deeper seabed close to them. Oceanic methane hydrates are usually found buried in sediments where the seawater is at least 300 m (980 ft) deep. They can be found up to about 2,000 m (6,600 ft) below the sea floor, but usually only about 1,100 m (3,600 ft) below the sea floor.[239]

The release of methane from the clathrates has been considered as a cause because scientists have found worldwide evidence of a swift decrease of about 1% in the 13 C 12 C isotope ratio in carbonate rocks from the end-Permian.[62][240] This is the first, largest, and most rapid of a series of negative and positive excursions (decreases and increases in 13 C 12 C ratio) that continues until the isotope ratio abruptly stabilised in the middle Triassic, followed soon afterwards by the recovery of calcifying life forms (organisms that use calcium carbonate to build hard parts such as shells).[83] While a variety of factors may have contributed to this drop in the 13 C 12 C ratio, a 2002 review found most of them to be insufficient to account fully for the observed amount:[241]

  • Gases from volcanic eruptions have a 13 C 12 C ratio about 0.5 to 0.8% below standard (قالب:Delta about −0.5 to −0.8%), but an assessment made in 1995 concluded that the amount required to produce a reduction of about 1.0% worldwide requires eruptions greater by orders of magnitude than any for which evidence has been found.[242] (However, this analysis addressed only CO2 produced by the magma itself, not from interactions with carbon bearing sediments, as later proposed.)
  • A reduction in organic activity would extract 12 C more slowly from the environment and leave more of it to be incorporated into sediments, thus reducing the 13 C 12 C 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 13 C 12 C (قالب:Delta −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 12 C would not have produced the 'smaller' drop in the 13 C 12 C ratio of the rocks around the PETM.[242]
  • Buried sedimentary organic matter has a 13 C 12 C ratio 2.0 to 2.5% below normal (قالب:Delta −2.0 to −2.5%). Theoretically, if the sea level fell sharply, shallow marine sediments would be exposed to oxidation. But 6,500–8,400 gigatons (1 gigaton = 109 metric tons) of organic carbon would have to be oxidized and returned to the ocean-atmosphere system within less than a few hundred thousand years to reduce the 13 C 12 C ratio by 1.0%, which is not thought to be a realistic possibility.[38] Moreover, sea levels were rising rather than falling at the time of the extinction.[212]
  • 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 13 C 12 C ratio fluctuations in the Early Triassic;[83] 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 13 C 12 C ratio because biochemical processes use the lighter isotopes more.[243] That or another organic-based reason may have been responsible for both that and a late Proterozoic/Cambrian pattern of fluctuating 13 C 12 C ratios.[83]

Prior to consideration of the inclusion of roasting carbonate sediments by volcanism, the only proposed mechanism sufficient to cause a global 1% reduction in the 13 C 12 C ratio was the release of methane from methane clathrates.[38] Carbon-cycle models confirm that it would have had enough effect to produce the observed reduction.[241][244] It was also suggested that a large-scale release of methane and other greenhouse gases from the ocean into the atmosphere was connected to the anoxic events and euxinic (i.e. sulfidic) events at the time, with the exact mechanism compared to the 1986 Lake Nyos disaster [245]

The area covered by lava from the Siberian Traps eruptions is about twice as large as was originally thought, and most of the additional area was shallow sea at the time. The seabed probably contained methane hydrate deposits, and the lava caused the deposits to dissociate, releasing vast quantities of methane.[246] 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 (18 O 16 O);[247] 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.[248]

However, the pattern of isotope shifts expected to result from a massive release of methane does not match the patterns seen throughout the Early Triassic. Not only would such a cause require the release of five times as much methane as postulated for the PETM,[83] but would it also have to be reburied at an unrealistically high rate to account for the rapid increases in the 13 C 12 C ratio (episodes of high positive قالب:Delta) throughout the early Triassic before it was released several times again.[83] The latest research suggests that greenhouse gas release during the extinction event was dominated by volcanic carbon dioxide,[249] 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.[250]

Hypercapnia and acidification

Marine organisms are more sensitive to changes in CO2 (carbon dioxide) levels than terrestrial organisms for a variety of reasons. CO2 is 28 times more soluble in water than is oxygen. Marine animals normally function with lower concentrations of CO2 in their bodies than land animals, as the removal of CO2 in air-breathing animals is impeded by the need for the gas to pass through the respiratory system's membranes (lungs' alveolus, tracheae, and the like), even when CO2 diffuses more easily than oxygen. In marine organisms, relatively modest but sustained increases in CO2 concentrations hamper the synthesis of proteins, reduce fertilization rates, and produce deformities in calcareous hard parts.[126] Higher concentrations of CO2 also result in decreased activity levels in many active marine animals, hindering their ability to obtain food.[251] 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.[126]

In addition, an increase in CO2 concentration is inevitably linked to ocean acidification, consistent with the preferential extinction of heavily calcified taxa and other signals in the rock record that suggest a more acidic ocean.[252] The decrease in ocean pH is calculated to be up to 0.7 units.[253] 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.[2] 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.[254] 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.[255]

On land, the increasing acidification of rainwater caused increased soil erosion as a result of the increased acidity of forest soils, evidenced by the increased influx of terrestrially derived organic sediments found in marine sedimentary deposits during the end-Permian extinction.[256] Further evidence of an increase in soil acidity comes from elevated Ba/Sr ratios in earliest Triassic soils.[257] 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.[258] 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.[259] As with many other environmental stressors, acidity on land episodically persisted well into the Triassic, stunting the recovery of terrestrial ecosystems.[260]

Anoxia and euxinia

Evidence for widespread ocean anoxia (severe deficiency of oxygen) and euxinia (presence of hydrogen sulfide) is found from the Late Permian to the Early Triassic.[261][262] Throughout most of the Tethys and Panthalassic Oceans, evidence for anoxia, including fine laminations in sediments, small pyrite framboids, high uranium/thorium ratios, and biomarkers for green sulfur bacteria, appear at the extinction event.[263] However, evidence for anoxia precedes the extinction at some other sites, including Spiti, India,[264] Shangsi, China,[265] Meishan, China,[266] Opal Creek, Alberta,[267] and Kap Stosch, Greenland.[268] 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.[269] 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.[270] 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.[271]

This spread of toxic, oxygen-depleted water would have devastated marine life, causing widespread die-offs. Models of ocean chemistry suggest that anoxia and euxinia were closely associated with hypercapnia (high levels of carbon dioxide).[272] This suggests that poisoning from hydrogen sulfide, anoxia, and hypercapnia acted together as a killing mechanism. Hypercapnia best explains the selectivity of the extinction, but anoxia and euxinia probably contributed to the high mortality of the event. The persistence of anoxia through the Early Triassic may explain the slow recovery of marine life and low levels of biodiversity after the extinction,[273][274][275] particularly that of benthic organisms.[276] Models also show that anoxic events can cause catastrophic hydrogen sulfide emissions into the atmosphere (see below).[277]

The sequence of events leading to anoxic oceans may have been triggered by carbon dioxide emissions from the eruption of the Siberian Traps.[277] In that scenario, warming from the enhanced greenhouse effect would reduce the solubility of oxygen in seawater, causing the concentration of oxygen to decline. Increased weathering of the continents due to warming and the acceleration of the water cycle would increase the riverine flux of phosphate to the ocean. The phosphate would have supported greater primary productivity in the surface oceans. The increase in organic matter production would have caused more organic matter to sink into the deep ocean, where its respiration would further decrease oxygen concentrations. Once anoxia became established, it would have been sustained by a positive feedback loop because deep water anoxia tends to increase the recycling efficiency of phosphate, leading to even higher productivity.[278] Convective overturn helped facilitate the expansion of anoxia throughout the water column.[279]

A severe anoxic event at the end of the Permian would have allowed sulfate-reducing bacteria to thrive, causing the production of large amounts of hydrogen sulfide in the anoxic ocean, turning it euxinic. Upwelling of this water may have released massive hydrogen sulfide emissions into the atmosphere and would poison terrestrial plants and animals and severely weaken the ozone layer, exposing much of the life that remained to fatal levels of UV radiation.[277] Indeed, biomarker evidence for anaerobic photosynthesis by Chlorobiaceae (green sulfur bacteria) from the Late-Permian into the Early Triassic indicates that hydrogen sulfide did upwell into shallow waters because these bacteria are restricted to the photic zone and use sulfide as an electron donor. The hypothesis has the advantage of explaining the mass extinction of plants, which would have added to the methane levels and should otherwise have thrived in an atmosphere with a high level of carbon dioxide. Fossil spores from the end-Permian further support the theory; many spores show deformities that could have been caused by ultraviolet radiation, which would have been more intense after hydrogen sulfide emissions weakened the ozone layer.[280] Fossilised plants 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.[281]

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.[282]

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.[283]

التقحل

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.[284] 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.[285] 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.[188][187]

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.[286] 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.[287]

In the Kuznetsk Basin of southwestern Siberia, an increase in aridity led to the demise of the humid-adapted cordaites forests in the region a few hundred thousand years before the Permian-Triassic boundary. This has been attributed to a broader poleward shift of drier, more arid climates during the late Changhsingian before the more abrupt main phase of the extinction at the Permian-Triassic boundary that disproportionately affected tropical and subtropical species.[117]

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.[288]

ارتطام كويكب

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

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

Reported evidence for an impact event from the P–Tr boundary level includes rare grains of shocked quartz in Australia and Antarctica;[289][290] fullerenes trapping extraterrestrial noble gases;[291] meteorite fragments in Antarctica;[292] and grains rich in iron, nickel, and silicon, which may have been created by an impact.[293] However, the accuracy of most of these claims has been challenged.[294][295][296][297] 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.[298]

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.[299] 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,[212] such as the Siberian Traps eruptions at either an impact site[300] or the antipode of an impact site.[212][301] 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.[302] Additionally, many sites spanning the Permian-Triassic boundary display a complete lack of evidence of an impact event.[303]

مواقع الارتطام المحتملة

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[290] and the hypothesized 480 km (300 mi) Wilkes Land crater of East Antarctica.[304][305] 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,[306] precipitated the mass extinction.[307] 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.[308] The large earthquakes generated by the impact would have additionally generated massive tsunamis across much of the globe.[307][309] 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.[308]

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),[310] 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.[311]

الميكروبات ميثانية الأصل

A hypothesis published in 2014 posits that a genus of anaerobic methanogenic archaea known as Methanosarcina was responsible for the event.[23] Three lines of evidence suggest that these microbes acquired a new metabolic pathway via gene transfer at about that time, enabling them to efficiently metabolize acetate into methane. That would have led to their exponential reproduction, allowing them to rapidly consume vast deposits of organic carbon that had accumulated in the marine sediment. The result would have been a sharp buildup of methane and carbon dioxide in the Earth's oceans and atmosphere, in a manner that may be consistent with the 13C/12C isotopic record. Massive volcanism facilitated this process by releasing large amounts of nickel, a scarce metal which is a cofactor for enzymes involved in producing methane.[23] On the other hand, in the canonical Meishan sections, the nickel concentration increases somewhat after the قالب:Delta concentrations have begun to fall.[312]

القارة العظمى پانگايا

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

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

Oceanic circulation and atmospheric weather patterns during the mid-Permian produced seasonal monsoons near the coasts and an arid climate in the vast continental interior.[313]

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

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

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

غبار ما بين النجوم

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.[314]

تضافر الأسباب

Possible causes supported by strong evidence appear to describe a sequence of catastrophes, each worse than the last: the Siberian Traps eruptions were bad enough alone, but because they occurred near coal beds and the continental shelf, they also triggered very large releases of carbon dioxide and methane. The resultant global warming may have caused perhaps the most severe anoxic event in the oceans' history: according to this theory, the oceans became so anoxic, anaerobic sulfur-reducing organisms dominated the chemistry of the oceans and caused massive emissions of toxic hydrogen sulfide.[126]

However, there may be some weak links in this chain of events: the changes in the 13C/12C ratio expected to result from a massive release of methane do not match the patterns seen throughout the early Triassic;[83] and the types of oceanic thermohaline circulation that may have existed at the end of the Permian are not likely to have supported deep-sea anoxia.[282]

انظر أيضاً

المراجع

  1. ^ 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.
  2. ^ أ ب Beauchamp, Benoit; Grasby, Stephen E. (15 September 2012). "Permian lysocline shoaling and ocean acidification along NW Pangea led to carbonate eradication and chert expansion". Palaeogeography, Palaeoclimatology, Palaeoecology. 350–352: 73–90. Bibcode:2012PPP...350...73B. doi:10.1016/j.palaeo.2012.06.014. Retrieved 21 December 2022.
  3. ^ 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.
  4. ^ 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: Virtual Field Trips Through the Nature of the Past. Springer Cham. pp. 253–267. doi:10.1007/978-3-030-35058-1_10. ISBN 978-3-030-35060-4. S2CID 226405085.
  5. ^ ""Great Dying" lasted 200,000 years". National Geographic. 23 November 2011. Retrieved 1 April 2014.
  6. ^ 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.
  7. ^ 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)
  8. ^ أ ب 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 (in الإنجليزية). 13 (11): 745–750. Bibcode:2020NatGe..13..745J. doi:10.1038/s41561-020-00646-4. ISSN 1752-0908. S2CID 224783993. Retrieved 8 November 2020.
  9. ^ 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. Retrieved 5 March 2023.
  10. ^ أ ب 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. ^ أ ب 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 (in الإنجليزية). 113 (42): E6325–E6334. Bibcode:2016PNAS..113E6325S. doi:10.1073/pnas.1613094113. ISSN 0027-8424. PMC 5081622. PMID 27698119.
  13. ^ أ ب 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. ^ أ ب ت ث ج 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. ^ أ ب 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. ^ 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.
  18. ^ أ ب ت ث 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. ^ 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.
  20. ^ Yin HF, Sweets WC, Yang ZY, Dickins JM (1992). "Permo–Triassic events in the eastern Tethys – an overview". In Sweet WC (ed.). Permo–Triassic Events in the Eastern Tethys: Stratigraphy, classification, and relations with the western Tethys. Cambridge: Cambridge University Press. pp. 1–7. ISBN 978-0-521-54573-0.
  21. ^ أ ب 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.
  22. ^ أ ب 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 (in الإنجليزية). 49 (3): 289–293. doi:10.1130/G48022.1. ISSN 0091-7613.   Available under CC BY 4.0.
  23. ^ أ ب ت 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.
  24. ^ 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.
  25. ^ 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.
  26. ^ 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.
  27. ^ "It took Earth ten million years to recover from greatest mass extinction". ScienceDaily. 27 May 2012. Retrieved 28 May 2012.
  28. ^ 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.
  29. ^ 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.
  30. ^ 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.
  31. ^ 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.
  32. ^ 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.
  33. ^ 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.
  34. ^ 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.
  35. ^ أ ب 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.
  36. ^ 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.
  37. ^ Song, Haijun; Wignall, Paul B.; 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.
  38. ^ أ ب ت ث ج Erwin, D.H. (1993). The great Paleozoic crisis; Life and death in the Permian. Columbia University Press. ISBN 978-0-231-07467-4.
  39. ^ أ ب Burgess, S.D. (2014). "High-precision timeline for Earth's most severe extinction". Proceedings of the National Academy of Sciences of the United States of America. 111 (9): 3316–3321. Bibcode:2014PNAS..111.3316B. doi:10.1073/pnas.1317692111. PMC 3948271. PMID 24516148.
  40. ^ أ ب ت ث ج ح خ د ذ ر ز س خطأ استشهاد: وسم <ref> غير صحيح؛ لا نص تم توفيره للمراجع المسماة McElwain2007
  41. ^ 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.
  42. ^ Krull, S. J.; Retallack, J. R. (2000). "13C depth profiles from paleosols across the Permian–Triassic boundary: Evidence for methane release". GSA Bulletin. 112 (9): 1459–1472. Bibcode:2000GSAB..112.1459K. doi:10.1130/0016-7606(2000)112<1459:CDPFPA>2.0.CO;2. ISSN 0016-7606.
  43. ^ أ ب 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.
  44. ^ Musashi, M.; Isozaki, Y.; Koike, T.; Kreulen, R. (2001). "Stable carbon isotope signature in mid-Panthalassa shallow-water carbonates across the Permo–Triassic boundary: Evidence for 13C-depleted ocean". Earth and Planetary Science Letters. 193 (1–2): 9–20. Bibcode:2001E&PSL.191....9M. doi:10.1016/S0012-821X(01)00398-3.
  45. ^ "Daily CO2". Mauna Loa Observatory.
  46. ^ Visscher, Henk; Looy, Cindy V.; Collinson, Margaret E.; Brinkhuis, Henk; Cittert, Johanna H. A. van Konijnenburg-van; Kürschner, Wolfram M.; Sephton, Mark A. (2004-08-31). "Environmental mutagenesis during the end-Permian ecological crisis". Proceedings of the National Academy of Sciences of the United States of America. 101 (35): 12952–12956. Bibcode:2004PNAS..10112952V. doi:10.1073/pnas.0404472101. ISSN 0027-8424. PMC 516500. PMID 15282373.
  47. ^ أ ب 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.
  48. ^ Foster, C. B.; Stephenson, M. H.; Marshall, C.; Logan, G. A.; Greenwood, P. F. (2002). "A revision of Reduviasporonites Wilson 1962: Description, illustration, comparison and biological affinities". Palynology. 26 (1): 35–58. doi:10.2113/0260035.
  49. ^ 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.
  50. ^ López-Gómez, J.; Taylor, E. L. (2005). "Permian–Triassic transition in Spain: A multidisciplinary approach". Palaeogeography, Palaeoclimatology, Palaeoecology. 229 (1–2): 1–2. doi:10.1016/j.palaeo.2005.06.028.
  51. ^ أ ب ت 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
  52. ^ أ ب 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.
  53. ^ Retallack, G.J.; Smith, R.M.H.; Ward, P.D. (2003). "Vertebrate extinction across Permian-Triassic boundary in Karoo Basin, South Africa". Bulletin of the Geological Society of America. 115 (9): 1133–1152. Bibcode:2003GSAB..115.1133R. doi:10.1130/B25215.1.
  54. ^ Sephton, M.A.; Visscher, H.; Looy, C.V.; Verchovsky, A.B.; Watson, J.S. (2009). "Chemical constitution of a Permian-Triassic disaster species". Geology. 37 (10): 875–878. Bibcode:2009Geo....37..875S. doi:10.1130/G30096A.1.
  55. ^ 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.
  56. ^ 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.
  57. ^ 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.
  58. ^ 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.
  59. ^ Wang, S.C.; Everson, P.J. (2007). "Confidence intervals for pulsed mass extinction events". Paleobiology. 33 (2): 324–336. Bibcode:2007Pbio...33..324W. doi:10.1666/06056.1. S2CID 2729020.
  60. ^ Fielding, Christopher R.; Frank, Tracy D.; Savatic, Katarina; Mays, Chris; McLoughlin, Stephen; Vajda, Vivi; Nicoll, Robert S. (15 May 2022). "Environmental change in the late Permian of Queensland, NE Australia: The warmup to the end-Permian Extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 594: 110936. Bibcode:2022PPP...594k0936F. doi:10.1016/j.palaeo.2022.110936. S2CID 247514266. Retrieved 23 December 2022.
  61. ^ 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. Retrieved 5 March 2023.
  62. ^ أ ب ت 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.
  63. ^ Dal Corso, Jacopo; Song, Haijun; Callegaro, Sara; Chu, Daoliang; Sun, Yadong; Hilton, Jason; Grasby, Stephen E.; Joachimski, Michael M.; Wignall, Paul B. (22 February 2022). "Environmental crises at the Permian–Triassic mass extinction". Nature Reviews Earth & Environment. 3 (3): 197–214. Bibcode:2022NRvEE...3..197D. doi:10.1038/s43017-021-00259-4. hdl:10852/100010. S2CID 247013868. Retrieved 20 December 2022.
  64. ^ Rohde, R. A. & Muller, R. A. (2005). "Cycles in fossil diversity". Nature. 434 (7030): 209–210. Bibcode:2005Natur.434..208R. doi:10.1038/nature03339. PMID 15758998. S2CID 32520208. Retrieved 14 January 2023.
  65. ^ Bond, D.P.G.; Wignall, P.B.; Wang, W.; Izon, G.; Jiang, H. S.; Lai, X. L.; et al. (2010). "The mid-Capitanian (Middle Permian) mass extinction and carbon isotope record of South China". Palaeogeography, Palaeoclimatology, Palaeoecology. 292 (1–2): 282–294. Bibcode:2010PPP...292..282B. doi:10.1016/j.palaeo.2010.03.056.
  66. ^ أ ب Retallack, G. J.; Metzger, C. A.; Greaver, T.; Jahren, A. H.; Smith, R. M. H.; Sheldon, N. D. (November–December 2006). "Middle-Late Permian mass extinction on land". The Bulletin of the Geological Society of America. 118 (11–12): 1398–1411. Bibcode:2006GSAB..118.1398R. doi:10.1130/B26011.1.
  67. ^ Li, Dirson Jian (18 December 2012). "The tectonic cause of mass extinctions and the genomic contribution to biodiversification". Quantitative Biology. arXiv:1212.4229. Bibcode:2012arXiv1212.4229L.
  68. ^ 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.
  69. ^ 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.
  70. ^ Shen, S.; Shi, G. R. (2002). "Paleobiogeographical extinction patterns of Permian brachiopods in the Asian-western Pacific region". Paleobiology. 28 (4): 449–463. doi:10.1666/0094-8373(2002)028<0449:PEPOPB>2.0.CO;2. ISSN 0094-8373. S2CID 35611701.
  71. ^ Wang, X-D & Sugiyama, T. (December 2000). "Diversity and extinction patterns of Permian coral faunas of China". Lethaia. 33 (4): 285–294. doi:10.1080/002411600750053853.
  72. ^ 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.
  73. ^ Bambach, R. K.; Knoll, A. H.; Wang, S. C. (December 2004). "Origination, extinction, and mass depletions of marine diversity". Paleobiology. 30 (4): 522–542. doi:10.1666/0094-8373(2004)030<0522:OEAMDO>2.0.CO;2. ISSN 0094-8373. S2CID 17279135.
  74. ^ أ ب 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.
  75. ^ 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...611k1362Y. doi:10.1016/j.palaeo.2022.111362. Retrieved 20 February 2023.
  76. ^ Allen, Bethany J.; Clapham, Matthew E.; Saupe, Erin E.; Wignall, Paul B.; Hill, Daniel J.; Dunhill, Alexander M. (10 February 2023). "Estimating spatial variation in origination and extinction in deep time: a case study using the Permian–Triassic marine invertebrate fossil record". Paleobiology: 1–18. doi:10.1017/pab.2023.1. S2CID 256801383. Retrieved 23 March 2023.
  77. ^ 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.
  78. ^ 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.
  79. ^ 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.
  80. ^ Hofmann, R.; Buatois, L. A.; MacNaughton, R. B.; Mángano, M. G. (15 June 2015). "Loss of the sedimentary mixed layer as a result of the end-Permian extinction". Palaeogeography, Palaeoclimatology, Palaeoecology. 428: 1–11. doi:10.1016/j.palaeo.2015.03.036. Retrieved 19 December 2022.
  81. ^ "Permian: The Marine Realm and The End-Permian Extinction". paleobiology.si.edu. Retrieved 2016-01-26.
  82. ^ أ ب "Permian extinction". Encyclopædia Britannica. Retrieved 2016-01-26.
  83. ^ أ ب ت ث ج ح خ 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.
  84. ^ 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.
  85. ^ 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.
  86. ^ 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.
  87. ^ 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.
  88. ^ 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.
  89. ^ 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.
  90. ^ 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. Retrieved 28 March 2023.
  91. ^ 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.
  92. ^ 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.
  93. ^ 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.
  94. ^ 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.
  95. ^ 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.
  96. ^ 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.
  97. ^ 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. Retrieved 2 April 2023.
  98. ^ Huang, Yunfei; Tong, Jinnan; Tian, Li; Song, Haijun; Chu, Daoliang; Miao, Xue; Song, Ting (1 January 2023). "Temporal shell-size variations of bivalves in South China from the Late Permian to the early Middle Triassic". Palaeogeography, Palaeoclimatology, Palaeoecology. 609: 111307. Bibcode:2023PPP...609k1307H. doi:10.1016/j.palaeo.2022.111307. S2CID 253368808. Retrieved 13 January 2023.
  99. ^ 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.
  100. ^ 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.
  101. ^ 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.
  102. ^ 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.
  103. ^ 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.
  104. ^ Fraiser, M. L.; Twitchett, Richard J.; Frederickson, J. A.; Metcalfe, B.; Bottjer, D. J. (1 January 2011). "Gastropod evidence against the Early Triassic Lilliput effect: COMMENT". Geology. 39 (1): e232. Bibcode:2011Geo....39E.232F. doi:10.1130/G31614C.1. Retrieved 13 January 2023.
  105. ^ Brayard, Arnaud; Nützel, Alexander; Kajm, Andrzej; Escarguel, Gilles; Hautmann, Michael; Stephen, Daniel A.; Bylund, Kevin G.; Jenks, Jim; Bucher, Hugo (1 January 2011). "Gastropod evidence against the Early Triassic Lilliput effect: REPLY". Geology. 39 (1): e233. Bibcode:2011Geo....39E.233B. doi:10.1130/G31765Y.1. Retrieved 13 January 2023.
  106. ^ 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.
  107. ^ Atkinson, Jed W.; Wignall, Paul B.; Morton, Jacob D.; Aze, Tracy (9 January 2019). "Body size changes in bivalves of the family Limidae in the aftermath of the end-Triassic mass extinction: the Brobdingnag effect". Palaeontology. 62 (4): 561–582. Bibcode:2019Palgy..62..561A. doi:10.1111/pala.12415. S2CID 134070316. Retrieved 20 January 2023.
  108. ^ أ ب 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.
  109. ^ 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.
  110. ^ 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.
  111. ^ "The Dino Directory – Natural History Museum".
  112. ^ 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.
  113. ^ 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.
  114. ^ 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.
  115. ^ أ ب ت ث Bodnar, Josefina; Coturel, Eliana P.; Falco, Juan Ignacio; Beltrána, Marisol (4 March 2021). "An updated scenario for the end-Permian crisis and the recovery of Triassic land flora in Argentina". Historical Biology. 33 (12): 3654–3672. doi:10.1080/08912963.2021.1884245. S2CID 233810158. Retrieved 24 December 2022.
  116. ^ أ ب 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.
  117. ^ أ ب Davydov, V. I.; Karasev, E. V.; Nurgalieva, N. G.; Schmitz, M. D.; Budnikov, I. V.; Biakov, A. S.; Kuzina, D. M.; Silantiev, V. V.; Urazaeva, M. N.; Zharinova, V. V.; Zorina, S. O.; Gareev, B.; Vasilenko, D. V. (1 July 2021). "Climate and biotic evolution during the Permian-Triassic transition in the temperate Northern Hemisphere, Kuznetsk Basin, Siberia, Russia". Palaeogeography, Palaeoclimatology, Palaeoecology. 573: 110432. Bibcode:2021PPP...573k0432D. doi:10.1016/j.palaeo.2021.110432. S2CID 235530804. Retrieved 19 December 2022.
  118. ^ 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. S2CID 240118558. Retrieved 28 March 2023.
  119. ^ 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.
  120. ^ 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.
  121. ^ أ ب ت 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 (in الإنجليزية). 204: 103153. Bibcode:2020ESRv..20403153F. doi:10.1016/j.earscirev.2020.103153. S2CID 216433847.
  122. ^ 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.
  123. ^ 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. Retrieved 8 April 2023.
  124. ^ Maxwell, W.D. (1992). "Permian and Early Triassic extinction of non-marine tetrapods". Palaeontology. 35: 571–583.
  125. ^ "Bristol University – News – 2008: Mass extinction".
  126. ^ أ ب ت ث Knoll AH, Bambach RK, Payne JL, Pruss S, Fischer WW (2007). "Paleophysiology and end-Permian mass extinction" (PDF). Earth and Planetary Science Letters. 256 (3–4): 295–313. Bibcode:2007E&PSL.256..295K. doi:10.1016/j.epsl.2007.02.018. Retrieved 2021-12-13.
  127. ^ 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.
  128. ^ 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.
  129. ^ أ ب 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.
  130. ^ 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.
  131. ^ 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.
  132. ^ 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.
  133. ^ 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.
  134. ^ 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.
  135. ^ 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.
  136. ^ Hautmann, Michael (29 September 2009). "The first scallop" (PDF). Paläontologische Zeitschrift. 84 (2): 317–322. doi:10.1007/s12542-009-0041-5. S2CID 84457522.
  137. ^ 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.
  138. ^ أ ب 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.
  139. ^ 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.
  140. ^ 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.
  141. ^ 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.
  142. ^ 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.
  143. ^ 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.
  144. ^ 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.
  145. ^ Hofmann, Richard; Hautmann, Michael; Bucher, Hugo (October 2015). "Recovery dynamics of benthic marine communities from the Lower Triassic Werfen Formation, northern Italy". Lethaia. 48 (4): 474–496. doi:10.1111/let.12121.
  146. ^ 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.
  147. ^ 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.
  148. ^ 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.
  149. ^ 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. Retrieved 8 April 2023.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  150. ^ 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.
  151. ^ 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.
  152. ^ 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. PMID 29034568. S2CID 207103048. Retrieved 20 April 2023.
  153. ^ 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.
  154. ^ 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.
  155. ^ 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.
  156. ^ 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.
  157. ^ 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.
  158. ^ أ ب 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.
  159. ^ 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.
  160. ^ 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.
  161. ^ أ ب 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.
  162. ^ Clapham ME, Bottjer DJ, Shen S (2006). "Decoupled diversity and ecology during the end-Guadalupian extinction (late Permian)". Geological Society of America Abstracts with Programs. 38 (7): 117. Archived from the original on 2015-12-08. Retrieved 2008-03-28.
  163. ^ 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. Retrieved 14 April 2023.
  164. ^ 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.
  165. ^ 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.
  166. ^ 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.
  167. ^ 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.
  168. ^ 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.
  169. ^ 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.
  170. ^ 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.
  171. ^ 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.
  172. ^ 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.
  173. ^ 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.
  174. ^ 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.
  175. ^ 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.
  176. ^ 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.
  177. ^ Aftabuzzaman, Md.; Kaiho, Kunio; Biswas, Raman Kumar; Liu, Yuqing; Saito, Ryosuke; Tian, Li; Bhat, Ghulam M.; Chen, Zhong-Qiang (October 2021). "End-Permian terrestrial disturbance followed by the complete plant devastation, and the vegetation proto-recovery in the earliest-Triassic recorded in coastal sea sediments". Global and Planetary Change. 205: 103621. Bibcode:2021GPC...20503621A. doi:10.1016/j.gloplacha.2021.103621. Retrieved 24 December 2022.
  178. ^ 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.
  179. ^ أ ب 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.
  180. ^ 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.
  181. ^ خطأ استشهاد: وسم <ref> غير صحيح؛ لا نص تم توفيره للمراجع المسماة Retallack1996
  182. ^ 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.
  183. ^ 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.
  184. ^ Yao, Mingtao; Sun, Zuoyu; Meng, Qingqiang; Li, Jiachun; Jiang, Dayong (15 August 2022). "Vertebrate coprolites from Middle Triassic Chang 7 Member in Ordos Basin, China: Palaeobiological and palaeoecological implications". Palaeogeography, Palaeoclimatology, Palaeoecology. 600: 111084. Bibcode:2022PPP...600k1084M. doi:10.1016/j.palaeo.2022.111084. S2CID 249186414. Retrieved 8 January 2023.
  185. ^ 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. doi:10.1016/j.gr.2019.03.009. Retrieved 24 April 2023.
  186. ^ 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. doi:10.1111/j.1502-3931.2007.00011.x. Archived from the original (PDF) on 2008-09-10. Retrieved 2008-07-02.
  187. ^ أ ب 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.
  188. ^ أ ب Yu, Yingyue; Tian, Li; Chu, Daoliang; Song, Huyue; Guo, Wenwei; Tong, Jinnan (1 January 2022). "Latest Permian–Early Triassic paleoclimatic reconstruction by sedimentary and isotopic analyses of paleosols from the Shichuanhe section in central North China Basin". Palaeogeography, Palaeoclimatology, Palaeoecology. 585: 110726. Bibcode:2022PPP...585k0726Y. doi:10.1016/j.palaeo.2021.110726. S2CID 239498183. Retrieved 6 November 2022.
  189. ^ Benton, M.J. (2004). Vertebrate Paleontology. Blackwell Publishers. xii–452. ISBN 978-0-632-05614-9.
  190. ^ 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.
  191. ^ 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. Retrieved 8 April 2023.
  192. ^ 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.
  193. ^ 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. doi:10.1002/gj.4657. S2CID 254207982. Retrieved 8 April 2023.
  194. ^ أ ب 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.
  195. ^ Wignall, Paul B.; 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. Retrieved 23 December 2022.
  196. ^ Andy Saunders; Marc Reichow (2009). "The Siberian Traps – area and volume". Retrieved 2009-10-18.
  197. ^ 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.
  198. ^ 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.
  199. ^ 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.
  200. ^ 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.
  201. ^ 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 (in الإنجليزية). 111 (9): 3316–3321. Bibcode:2014PNAS..111.3316B. doi:10.1073/pnas.1317692111. ISSN 0027-8424. PMC 3948271. PMID 24516148.
  202. ^ Black, Benjamin A.; Weiss, Benjamin P.; Elkins-Tanton, Linda T.; Veselovskiy, Roman V.; Latyshev, Anton (2015-04-30). "Siberian Traps volcaniclastic rocks and the role of magma-water interactions". Geological Society of America Bulletin (in الإنجليزية). 127 (9–10): B31108.1. Bibcode:2015GSAB..127.1437B. doi:10.1130/B31108.1. ISSN 0016-7606.
  203. ^ Burgess, Seth D.; Bowring, Samuel A. (2015-08-01). "High-precision geochronology confirms voluminous magmatism before, during, and after Earth's most severe extinction". Science Advances (in الإنجليزية). 1 (7): e1500470. Bibcode:2015SciA....1E0470B. doi:10.1126/sciadv.1500470. ISSN 2375-2548. PMC 4643808. PMID 26601239.
  204. ^ Fischman, Josh. Giant eruptions and giant extinctions. Scientific American (video). Retrieved 2016-03-11.
  205. ^ 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.
  206. ^ 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.
  207. ^ Wang, Wen-qian; Zheng, Feifei; Zhang, Shuang; Cui, Ying; Zheng, Quan-feng; Zhang, Yi-chun; Chang, Dong-xun; Zhang, Hua; Xu, Yi-gang; Shen, Shu-zhong (15 January 2023). "Ecosystem responses of two Permian biocrises modulated by CO2 emission rates". Earth and Planetary Science Letters. 602: 117940. Bibcode:2023E&PSL.60217940W. doi:10.1016/j.epsl.2022.117940. S2CID 254660567. Retrieved 20 February 2023.
  208. ^ Wignall, Paul B.; Sun, Yadong; Bond, David P. G.; Izon, Gareth; Newton, Robert J.; Védrine, Stéphanie; Widdowson, Mike; Ali, Jason R.; Lai, Xulong; Jiang, Haishui; Cope, Helen; Bottrell, Simon H. (2009). "Volcanism, Mass Extinction, and Carbon Isotope Fluctuations in the Middle Permian of China". Science. 324 (5931): 1179–1182. Bibcode:2009Sci...324.1179W. doi:10.1126/science.1171956. PMID 19478179. S2CID 206519019. Retrieved 20 February 2023.
  209. ^ Jerram, Dougal A.; Widdowson, Mike; Wignall, Paul B.; Sun, Yadong; Lai, Xulong; Bond, David P. G.; Torsvik, Trond H. (1 January 2016). "Submarine palaeoenvironments during Emeishan flood basalt volcanism, SW China: Implications for plume–lithosphere interaction during the Capitanian, Middle Permian ('end Guadalupian') extinction event". Palaeogeography, Palaeoclimatology, Palaeoecology. 441: 65–73. Bibcode:2016PPP...441...65J. doi:10.1016/j.palaeo.2015.06.009. Retrieved 19 December 2022.
  210. ^ 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.
  211. ^ 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.
  212. ^ أ ب ت ث ج ح White, R. V. (2002). "Earth's biggest 'whodunnit': Unravelling the clues in the case of the end-Permian mass extinction" (PDF). Philosophical Transactions of the Royal Society of London. 360 (1801): 2963–2985. Bibcode:2002RSPTA.360.2963W. doi:10.1098/rsta.2002.1097. PMID 12626276. S2CID 18078072. Retrieved 2008-01-12.
  213. ^ "Driver of the largest mass extinction in the history of the Earth identified". phys.org (in الإنجليزية). Retrieved 8 November 2020.
  214. ^ "Large volcanic eruption caused the largest mass extinction". phys.org (in الإنجليزية). Retrieved 8 December 2020.
  215. ^ أ ب 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.
  216. ^ 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.
  217. ^ 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.
  218. ^ "Volcanism". Hooper Museum. hoopermuseum.earthsci.carleton.ca. Ottawa, Ontario, Canada: Carleton University.
  219. ^ 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.
  220. ^ 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 (in الإنجليزية). 394: 242–253. Bibcode:2014E&PSL.394..242K. doi:10.1016/j.epsl.2014.03.029.
  221. ^ 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.
  222. ^ Dal Corso, Jacopo; Mills, Benjamin J. W.; Chu, Daoling; Newton, Robert J.; Mather, Tamsin A.; Shu, Wenchao; Wu, Yuyang; Tong, Jinnan; Wignall, Paul B. (11 June 2020). "Permo–Triassic boundary carbon and mercury cycling linked to terrestrial ecosystem collapse". Nature Communications. 11 (1): 2962. Bibcode:2020NatCo..11.2962D. doi:10.1038/s41467-020-16725-4. PMC 7289894. PMID 32528009.
  223. ^ Verango, Dan (24 January 2011). "Ancient mass extinction tied to torched coal". USA Today.
  224. ^ 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.
  225. ^ "Researchers find smoking gun of world's biggest extinction: Massive volcanic eruption, burning coal and accelerated greenhouse gas choked out life" (Press release). University of Calgary. January 23, 2011. Retrieved 2011-01-26.
  226. ^ 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.
  227. ^ 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.
  228. ^ أ ب Grasby, Stephen E.; Beauchamp, Benoit; Bond, David P. G.; Wignall, Paul B.; Talavera, Cristina; Galloway, Jennifer M.; Piepjohn, Karsten; Reinhardt, Lutz; Blomeier, Dirk (1 September 2015). "Progressive environmental deterioration in northwestern Pangea leading to the latest Permian extinction". Geological Society of America Bulletin. 127 (9–10): 1331–1347. Bibcode:2015GSAB..127.1331G. doi:10.1130/B31197.1. Retrieved 14 January 2023.
  229. ^ Grasby, Stephen E.; Beauchamp, Benoit; Bond, David P. G.; Wignall, Paul B.; Sanei, Hamed (2016). "Mercury anomalies associated with three extinction events (Capitanian Crisis, Latest Permian Extinction and the Smithian/Spathian Extinction) in NW Pangea". Geological Magazine. 153 (2): 285–297. Bibcode:2016GeoM..153..285G. doi:10.1017/S0016756815000436. S2CID 85549730. Retrieved 16 September 2022.
  230. ^ 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.
  231. ^ 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.
  232. ^ 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.
  233. ^ 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.
  234. ^ 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.
  235. ^ 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.
  236. ^ Nelson, D. A.; Cottle, J. M. (29 March 2019). "Tracking voluminous Permian volcanism of the Choiyoi Province into central Antarctica". Lithosphere. 11 (3): 386–398. Bibcode:2019Lsphe..11..386N. doi:10.1130/L1015.1. S2CID 135130436. Retrieved 14 January 2023.
  237. ^ أ ب 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.
  238. ^ 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.
  239. ^ 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.
  240. ^ 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.
  241. ^ أ ب 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.
  242. ^ أ ب 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.
  243. ^ Schrag DP, Berner RA, Hoffman PF, Halverson GP (2002). "On the initiation of a snowball Earth". Geochemistry, Geophysics, Geosystems. 3 (6): 1–21. Bibcode:2002GGG....3fQ...1S. doi:10.1029/2001GC000219. Preliminary abstract at Schrag, D.P. (June 2001). "On the initiation of a snowball Earth". Geological Society of America. Archived from the original on 2018-04-25. Retrieved 2008-04-20.
  244. ^ 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.
  245. ^ 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.
  246. ^ Reichow MK, Saunders AD, White RV, Pringle MS, Al'Muhkhamedov AI, Medvedev AI, Kirda NP (2002). " 40 Ar 39 Ar 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.
  247. ^ 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.
  248. ^ 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.
  249. ^ 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.
  250. ^ Wu, Yuyang; Chu, Daoliang; Tong, Jinnan; Song, Haijun; Dal Corso, Jacopo; Wignall, Paul B.; Song, Huyue; Du, Yong; Cui, Ying (9 April 2021). "Six-fold increase of atmospheric pCO2 during the Permian–Triassic mass extinction". Nature Communications. 12 (1): 2137. Bibcode:2021NatCo..12.2137W. doi:10.1038/s41467-021-22298-7. PMC 8035180. PMID 33837195. S2CID 233200774.
  251. ^ 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.
  252. ^ 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.
  253. ^ 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.
  254. ^ Wignall, Paul B.; 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. Retrieved 7 November 2022.
  255. ^ 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.
  256. ^ Sephton, Mark A.; Jiao, Dan; Engel, Michael H.; Looy, Cindy V.; Visscher, Henk (1 February 2015). "Terrestrial acidification during the end-Permian biosphere crisis?". Geology. 43 (2): 159–162. Bibcode:2015Geo....43..159S. doi:10.1130/G36227.1. Retrieved 23 December 2022.
  257. ^ Sheldon, Nathan D. (28 February 2006). "Abrupt chemical weathering increase across the Permian–Triassic boundary". Palaeogeography, Palaeoclimatology, Palaeoecology. 231 (3–4): 315–321. doi:10.1016/j.palaeo.2005.09.001. Retrieved 24 April 2023.
  258. ^ 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.
  259. ^ 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.
  260. ^ 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.
  261. ^ Song, Haijun; Wignall, Paul B.; Tong, Jinnan; Bond, David P. G.; Song, Huyue; Lai, Xulong; Zhang, Kexin; Wang, Hongmei; Chen, Yanlong (1 November 2012). "Geochemical evidence from bio-apatite for multiple oceanic anoxic events during Permian–Triassic transition and the link with end-Permian extinction and recovery". Earth and Planetary Science Letters. 353–354: 12–21. Bibcode:2012E&PSL.353...12S. doi:10.1016/j.epsl.2012.07.005. Retrieved 13 January 2023.
  262. ^ 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.
  263. ^ Wignall, P.B.; Twitchett, R.J. (2002). Extent, duration, and nature of the Permian-Triassic superanoxic event. Vol. 356. pp. 395–413. Bibcode:2002GSASP.356..679O. doi:10.1130/0-8137-2356-6.395. ISBN 978-0-8137-2356-3. {{cite book}}: |journal= ignored (help)
  264. ^ 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. Retrieved 21 December 2022.
  265. ^ 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.
  266. ^ 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.
  267. ^ 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.
  268. ^ 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.
  269. ^ 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.
  270. ^ 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. Retrieved 29 October 2022.
  271. ^ 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.
  272. ^ 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.
  273. ^ 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.
  274. ^ Wignall, Paul B.; Hallam, A. (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.
  275. ^ 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.
  276. ^ 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.
  277. ^ أ ب ت 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.
  278. ^ Schobben, Martin; Foster, William J.; Sleveland, Arve R. N.; Zuchuat, Valentin; Svensen, Henrik H.; Planke, Sverre; Bond, David P. G.; Marcelis, Fons; Newton, Robert J.; Wignall, Paul B.; Poulton, Simon W. (17 August 2020). "A nutrient control on marine anoxia during the end-Permian mass extinction". Nature Geoscience. 13 (9): 640–646. Bibcode:2020NatGe..13..640S. doi:10.1038/s41561-020-0622-1. S2CID 221146234. Retrieved 8 January 2023.
  279. ^ 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. doi:10.1016/j.chemgeo.2010.09.001. Retrieved 24 April 2023.
  280. ^ Visscher, H.; Looy, C.V.; Collinson, M.E.; Brinkhuis, H.; van Konijnenburg-van Cittert, J.H.A.; Kurschner, W.M.; Sephton, M.A. (2004-07-28). "Environmental mutagenesis during the end-Permian ecological crisis". Proceedings of the National Academy of Sciences of the United States of America (in الإنجليزية). 101 (35): 12952–12956. Bibcode:2004PNAS..10112952V. doi:10.1073/pnas.0404472101. ISSN 0027-8424. PMC 516500. PMID 15282373.
  281. ^ 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.
  282. ^ أ ب 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.
  283. ^ 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.
  284. ^ 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.
  285. ^ Chinsamy-Turan (2012). Anusuya (ed.). Forerunners of mammals : radiation, histology, biology. Bloomington: Indiana University Press. ISBN 978-0-253-35697-0.
  286. ^ 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.
  287. ^ 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.
  288. ^ 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.
  289. ^ 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.
  290. ^ أ ب 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.
  291. ^ 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.
  292. ^ 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.
  293. ^ 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.
  294. ^ 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.
  295. ^ 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.
  296. ^ 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.
  297. ^ 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.
  298. ^ (2005) "Reexamination of quartz grains from the Permian–Triassic boundary section at Graphite Peak, Antarctica".. 
  299. ^ 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.
  300. ^ AHager, Bradford H. (2001). "Giant Impact Craters Lead To Flood Basalts: A Viable Model".. 
  301. ^ Hagstrum, Jonathan T. (2001). "Large Oceanic Impacts As The Cause Of Antipodal Hotspots And Global Mass Extinctions".. 
  302. ^ 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.
  303. ^ 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.
  304. ^ 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.
  305. ^ 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.
  306. ^ 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 40 Ar / 39 Ar 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.
  307. ^ أ ب "Biggest extinction in history caused by climate-changing meteor". University News (Press release). University of Western Australia. 31 July 2013.
  308. ^ أ ب 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 29 November 2022.
  309. ^ 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 29 November 2022.
  310. ^ 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.
  311. ^ 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 (in الإنجليزية). 29 (6): 411–415. Bibcode:2017TeNov..29..411M. doi:10.1111/ter.12285. ISSN 0954-4879. S2CID 133781924. Retrieved 26 March 2023.
  312. ^ 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.
  313. ^ Ziegler, A.; Eshel, G.; Rees, P.; Rothfus, T.; Rowley, D.; Sunderlin, D. (2003). "Tracing the tropics across land and sea: Permian to present". Lethaia. 36 (3): 227–254. CiteSeerX 10.1.1.398.9447. doi:10.1080/00241160310004657.
  314. ^ 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.

للاستزادة

وصلات خارجية

قالب:ExtEvent nav