Tuesday, May 31, 2016

General Mills Recalls Flour Amid Possible E.coli Contamination

begin quote from:

General Mills Recalls Flour Amid Possible E.coli Contamination

Wall Street Journal - ‎8 hours ago‎
General Mills Inc. GIS -0.14 % said it is voluntarily recalling certain lots of flour as a precaution amid a 20-state outbreak of a strain of E.coli potentially linked to the products.
 
DOW JONES, A NEWS CORP COMPANY

General Mills Recalls Flour Amid Possible E.coli Contamination

Company is recalling several lots of Gold Medal, Wondra and Signature Kitchen brands

General Mills issued a voluntary recall Tuesday of certain lots of Gold Medal, Wondra and Signature Kitchen flour amid an outbreak of a strain of E.coli potentially linked to the products. ENLARGE
General Mills issued a voluntary recall Tuesday of certain lots of Gold Medal, Wondra and Signature Kitchen flour amid an outbreak of a strain of E.coli potentially linked to the products. Photo: abir sultan/European Pressphoto Agency
General Mills Inc. said it is voluntarily recalling certain lots of flour as a precaution amid a 20-state outbreak of a strain of E.coli potentially linked to the products.
The Minneapolis cereal company said the E.coli strain—called E.coli 0121—hasn’t been found in any of its flour products or its flour manufacturing plant. General Mills says it hasn’t been contacted directly by any consumer reporting a confirmed illness related to its products.
General Mills said it is recalling several lots of its Gold Medal, Wondra and Signature Kitchen brands that were sold in supermarkets such as Safeway, Albertsons, Jewel, Shaws, Vons, United, Randalls and Acme.
According to General Mills, state and federal authorities have been looking into 38 occurrences of illnesses across 20 states related to the E.coli strain. The company stated in a news release that roughly half of the individuals had reported making something with flour at some point before becoming ill, with some reporting they had used a General Mills brand. The 38 cases occurred between Dec. 21, 2015, and May 3.
Consumers were advised to stop using the flour included in the recall.
The company is providing more details at www.generalmills.com/flour.
Write to Tess Stynes at tess.stynes@wsj.com
 

Permian–Triassic extinction event killed 96% of marine species and 70% of all terrestrial vertebrates

begin quote from:

  Permian–Triassic extinction event

Permian–Triassic extinction event

From Wikipedia, the free encyclopedia
Extinction intensity.svg Cambrian Ordovician Silurian Devonian Carboniferous Permian Triassic Jurassic Cretaceous Paleogene Neogene
Marine extinction intensity during the Phanerozoic
%
Millions of years ago
P–Tr
Extinction intensity.svg
Plot of extinction intensity (percentage of genera that are present in each interval of time but do not exist in the following interval) vs time in the past for marine genera.[1] Geological periods are annotated (by abbreviation and colour) above. The Permian–Triassic extinction event is the most significant event for marine genera, with just over 50% (according to this source) failing to survive. (source and image info)
The Permian–Triassic (P–Tr) extinction event, colloquially known as the Great Dying, the End Permian or the Great Permian Extinction,[2][3] occurred about 252 Ma (million years) ago,[4] forming the boundary between the Permian and Triassic geologic periods, as well as the Paleozoic and Mesozoic eras. It is the Earth's most severe known extinction event, with up to 96% of all marine species[5][6] and 70% of terrestrial vertebrate species becoming extinct.[7] It is the only known mass extinction of insects.[8][9] Some 57% of all families and 83% of all genera became extinct. Because so much biodiversity was lost, the recovery of life on Earth took significantly longer than after any other extinction event,[5] possibly up to 10 million years.[10]
There is evidence for between one and three distinct pulses, or phases, of extinction.[7][11][12][13] Suggested mechanisms for the latter include one or more large bolide impact events, massive volcanism, coal or gas fires and explosions from the Siberian Traps,[14] and a runaway greenhouse effect triggered by sudden release of methane from the sea floor due to methane clathrate dissociation or methane-producing microbes known as methanogens;[15] possible contributing gradual changes include sea-level change, increasing anoxia, increasing aridity, and a shift in ocean circulation driven by climate change.

Contents

Dating the extinction

Permian–Triassic extinction
view • discuss • edit
-262 —
-260 —
-258 —
-256 —
-254 —
-252 —
-250 —
-248 —
-246 —
-244 —
-242 —
-240 —
-238 —
Extinction event
δ13C stabilises at +2 ‰. Calcified green algae return.[16]
Full recovery of woody trees[18]
Corals return[19]
Scleractinian
corals & calcified sponges[16]
An approximate timescale of events around the P-Tr boundary.
Axis scale: millions of years ago.

O
S
D
C
P
T
J
K
N
Until 2000, 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.[20] However, it is now possible to date the extinction with remarkable 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 Ma, a duration of 60 ± 48 ka.[21] A large (approximately 0.9%), abrupt global decrease in the ratio of the stable isotope 13C to that of 12C, coincides with this extinction,[18][22][23][24][25] and is sometimes used to identify the Permian–Triassic boundary in rocks that are unsuitable for radiometric dating.[26] Further evidence for environmental change around the P–Tr boundary suggests an 8 °C (14 °F) rise in temperature,[18] and an increase in CO
2
levels by 2000 ppm (by contrast, the concentration immediately before the industrial revolution was 280 ppm.)[18] There is also evidence of increased ultraviolet radiation reaching the earth causing the mutation of plant spores.[18]
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.[27] 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 lack 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 in the earliest Triassic.[27] The very idea of a fungal spike has been criticized on several grounds, including: Reduviasporonites, the most common supposed fungal spore, was actually a fossilized alga;[18][28] the spike did not appear worldwide;[29][30] and in many places it did not fall on the Permian–Triassic boundary.[31] The algae, which were misidentified as fungal spores, 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.[32] Newer chemical evidence agrees better with a fungal origin for Reduviasporonites, diluting these critiques.[33]
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[7] or that the extinction was spread out over a few million years, with a sharp peak in the last million years of the Permian.[31][34] Statistical analyses of some highly fossiliferous strata in Meishan, Zhejiang Province in South China, suggest that the main extinction was clustered around one peak.[11] 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 670 to 1170 thousand years.[35] In a well-preserved sequence in east Greenland, the decline of animals is concentrated in a period 10 to 60 thousand years long, with plants taking several hundred thousand additional years to show the full impact of the event.[36] An older theory, still supported in some recent papers,[7][37] is that there were two major extinction pulses 9.4 million years apart, separated by a period of extinctions well above the background level, and that the final extinction killed off only about 80% of marine species alive at that time while the other losses occurred during the first pulse or the interval between pulses. According to this theory one of these extinction pulses occurred at the end of the Guadalupian epoch of the Permian.[7][38] For example, all but one of the surviving dinocephalian genera died out at the end of the Guadalupian,[39] as did the Verbeekinidae, a family of large-size fusuline foraminifera.[40] 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.[41][42]

Extinction patterns

Marine extinctions Genera extinct Notes
Arthropoda
Eurypterids 100% May have become extinct shortly before the P–Tr boundary
Ostracods 59%  
Trilobites 100% In decline since the Devonian; only 2 genera living before the extinction
Brachiopoda
Brachiopods 96% Orthids and productids died out
Bryozoa
Bryozoans 79% Fenestrates, trepostomes, and cryptostomes died out
Chordata
Acanthodians 100% In decline since the Devonian, with only one living family
Cnidaria
Anthozoans 96% Tabulate and rugose corals died out
Echinodermata
Blastoids 100% May have become extinct shortly before the P–Tr boundary
Crinoids 98% Inadunates and camerates died out
Mollusca
Ammonites 97%  
Bivalves 59%  
Gastropods 98%  
Retaria
Foraminiferans 97% Fusulinids died out, but were almost extinct before the catastrophe
Radiolarians 99%[43]

Marine organisms

Marine invertebrates suffered the greatest losses during the P–Tr extinction. Evidence of this was found in samples from south China sections at the P–Tr boundary. Here, 286 out of 329 marine invertebrate genera disappear within the final 2 sedimentary zones containing conodonts from the Permian.[11] The decrease in diversity was probably caused by a sharp increase in extinctions, rather than a decrease in speciation.[44]
The extinction primarily affected organisms with calcium carbonate skeletons, especially those reliant on stable CO2 levels to produce their skeletons.[45] These organisms were susceptible to the effects of the ocean acidification that resulted from increased atmospheric CO2.
Among benthic organisms, the extinction event multiplied background extinction rates, and therefore caused most damage to taxa that had a high background extinction rate (by implication, taxa with a high turnover).[46][47] The extinction rate of marine organisms was catastrophic.[11][48][49][50]
Surviving marine invertebrate groups include: articulate brachiopods (those with a hinge),[51] which have suffered a slow decline in numbers since the P–Tr extinction; the Ceratitida order of ammonites;[52] and crinoids ("sea lilies"),[52] 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 apparatus were the worst hit.[16][53] In the case of the brachiopods at least, surviving taxa were generally small, rare members of a diverse community.[54]
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, that is the range of different ecological guilds, environmental factors were apparently responsible. Diversity and disparity fell further until the P–Tr boundary; the extinction here was non-selective, consistent with a catastrophic initiator. During the Triassic, diversity rose rapidly, but disparity remained low.[55]
The range of morphospace occupied by the ammonoids, that is the range of possible forms, shape or structure, became more restricted as the Permian progressed. Just 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.[56]

Terrestrial invertebrates

The Permian had great diversity in insect and other invertebrate species, including the largest insects ever to have existed. The end-Permian is the only known mass extinction of insects,[8] with eight or nine insect orders becoming extinct and ten more 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.[48]
Most fossil insect groups found after the Permian–Triassic boundary differ significantly from those that lived prior to the P–Tr extinction. With the exception of the Glosselytrodea, Miomoptera, and Protorthoptera, Paleozoic insect groups have not been discovered in deposits dating to after the P–Tr boundary. The caloneurodeans, monurans, paleodictyopteroids, protelytropterans, and protodonates became extinct by the end of the Permian. In well-documented Late Triassic deposits, fossils overwhelmingly consist of modern fossil insect groups.[8]

Terrestrial plants

Plant ecosystem response

The geological record of terrestrial plants is sparse and based mostly on pollen and spore studies. Interestingly, plants are relatively immune to mass extinction, with the impact of all the major mass extinctions "insignificant" at a family level.[18] Even the reduction observed in species diversity (of 50%) may be mostly due to taphonomic processes.[18] However, a massive rearrangement of ecosystems does occur, with plant abundances and distributions changing profoundly and all the forests virtually disappearing;[18][57] the Palaeozoic flora scarcely survived this extinction.[58]
At the P–Tr boundary, the dominant floral groups changed, with many groups of land plants entering abrupt decline, such as Cordaites (gymnosperms) and Glossopteris (seed ferns).[59] Dominant gymnosperm genera were replaced post-boundary by lycophytes—extant lycophytes are recolonizers of disturbed areas.[60]
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. 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 afterwards. These fluctuations of the dominant flora between woody and herbaceous taxa indicate chronic environmental stress resulting in a loss of most large woodland plant species. The successions and extinctions of plant communities do not coincide with the shift in δ13C values, but occurred many years after.[30] The recovery of gymnosperm forests took 4–5 million years.[18]

Coal gap

No coal deposits are known from the Early Triassic, and those in the Middle Triassic are thin and low-grade.[19] 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.[19] 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.[19] On the other hand, abiotic factors (not caused by organisms), such as decreased rainfall or increased input of clastic sediments, may also be to blame.[18] Finally, it is also true that there are very few sediments of any type known from the Early Triassic, and the lack of coal may simply reflect this scarcity. This opens the possibility that coal-producing ecosystems may have responded to the changed conditions by relocating, perhaps to areas where we have no sedimentary record for the Early Triassic.[18] For example, in eastern Australia a cold climate had been the norm for a long period of time, with a peat mire ecosystem adapted to these conditions. Approximately 95% of these peat-producing plants went locally extinct at the P–Tr boundary;[61] Interestingly, coal deposits in Australia and Antarctica disappear significantly before the P–Tr boundary.[18]

Terrestrial vertebrates

There is enough evidence to indicate that over two-thirds of terrestrial labyrinthodont amphibians, sauropsid ("reptile") and therapsid ("mammal-like reptile") families became extinct. Large herbivores suffered the heaviest losses. All Permian anapsid reptiles died out except the procolophonids (testudines have anapsid skulls but are most often thought to have evolved later, 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).[62][63] Even the groups that survived suffered extremely heavy losses of species, and some terrestrial vertebrate groups very nearly became extinct at the end-Permian. Some of the surviving groups did not persist for long past this period, while others that barely survived went on to produce diverse and long-lasting lineages. Yet it took 30 million years for the terrestrial vertebrate fauna to fully recover both numerically and ecologically.[64]

Possible explanations of these patterns

An analysis of marine fossils from the Permian's final Changhsingian stage found that marine organisms with low tolerance for hypercapnia (high concentration of carbon dioxide) had high extinction rates, while the most tolerant organisms had very slight losses.
The most vulnerable marine organisms were those that produced calcareous hard parts (i.e., 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, for example 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.[65]
This pattern is consistent with what is known about the effects of hypoxia, a shortage but not a 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. 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—the decline in oxygen levels does not match the temporal pattern of the extinction.[65]
Marine organisms are more sensitive to changes in CO2 levels than are 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.[65] 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.[66]
It is difficult to analyze extinction and survival rates of land organisms in detail, because few terrestrial fossil beds span the Permian–Triassic boundary. Triassic insects are very different from those of the Permian, but a gap in the insect fossil record spans approximately 15 million years from the late Permian to early Triassic. 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.[65] However, analysis of the fossil river deposits of the floodplains indicate a shift from meandering to braided river patterns, indicating an abrupt drying of the climate.[67] The climate change may have taken as little as 100,000 years, prompting the extinction of the unique Glossopteris flora and its herbivores, followed by the carnivorous guild.[68] End-Permian extinctions did not occur at an instantaneous time horizon; particularly, floral extinction was delayed in time[69]

Biotic recovery

Earlier analyses indicated that life on Earth recovered quickly after the Permian extinctions, but this was mostly in the form of disaster taxa, opportunist organisms such as the hardy Lystrosaurus. Research published in 2006 indicates that the specialized animals that formed complex ecosystems, with high biodiversity, complex food webs and a variety of niches, took much longer to recover. It is thought that this long recovery was due to the successive waves of extinction, which inhibited recovery, and prolonged environmental stress to organisms, which continued into the Early Triassic. Research indicates that recovery did not begin until the start of the mid-Triassic, 4 to 6 million years after the extinction;[70] and some writers estimate that the recovery was not complete until 30 Ma after the P–Tr extinction, i.e. in the late Triassic.[7]
A study published in the journal Science[71] found that during the Great Extinction the oceans' surface temperatures reached 40 °C (104 °F), which explains why recovery took so long: it was simply too hot for life to survive.[72] Of course, not all of the Earth's surface was 40°, and if it was simply too hot for life to survive then nothing would have survived. Perhaps anoxia provides another aspect of what delayed the recovery.[73]
During the early Triassic (4 to 6 million years after the P–Tr extinction), the plant biomass was insufficient to form coal deposits, which implies a limited food mass for herbivores.[19] River patterns in the Karoo changed from meandering to braided, indicating that vegetation there was very sparse for a long time.[74]
Each major segment of the early Triassic ecosystem—plant and animal, marine and terrestrial—was dominated by a small number of genera, which appeared virtually worldwide, for example: the herbivorous therapsid Lystrosaurus (which accounted for about 90% of early Triassic land vertebrates) and the bivalves Claraia, Eumorphotis, Unionites and Promylina. A healthy ecosystem has a much larger number of genera, each living in a few preferred types of habitat.[59][75]
Disaster taxa took advantage of the devastated ecosystems and enjoyed a temporary population boom and increase in their territory. Microconchids are the dominant component of otherwise impoverished Early Triassic encrusting assemblages. For example: Lingula (a brachiopod); stromatolites, which had been confined to marginal environments since the Ordovician; Pleuromeia (a small, weedy plant); Dicroidium (a seed fern).[75][76][77]

Changes in marine ecosystems

Sessile filter feeders like this crinoid 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 sea floor but, 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.[78]
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,[78] and the increase in predation pressure led to the Mesozoic Marine Revolution.
Bivalves were fairly rare before the P–Tr extinction but became numerous and diverse in the Triassic, and one group, the rudist clams, became the Mesozoic's main reef-builders. Some researchers think much of this change happened in the 5 million years between the two major extinction pulses.[79]
Crinoids ("sea lilies") suffered a selective extinction, resulting in a decrease in the variety of their forms.[80] Their ensuing adaptive radiation was brisk, and resulted in forms possessing flexible arms becoming widespread; motility, predominantly a response to predation pressure, also became far more prevalent.[81]

Land vertebrates

Lystrosaurus was by far the most abundant early Triassic land vertebrate.
Lystrosaurus, a pig-sized herbivorous dicynodont therapsid, constituted as much as 90% of some earliest Triassic land vertebrate fauna. Smaller carnivorous cynodont therapsids also survived, including the ancestors of mammals. In the Karoo region of southern Africa, the therocephalians Tetracynodon, Moschorhinus and Ictidosuchoides survived, but do not appear to have been abundant in the Triassic.[82]
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.[83] 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 and higher metabolic rates,[84] while losing part of the differential color-sensitive retinal receptors reptilians and birds preserved.
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.[85]
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, in which dinosaurs, pterosaurs, crocodiles, archosaurs, amphibians, and mammaliforms were abundant and diverse.[5]

Causes

Pinpointing the exact cause or causes of the Permian–Triassic extinction event is difficult, mostly because the catastrophe occurred over 250 million years ago, and much of the evidence that would have pointed to the cause has been destroyed by now or is concealed deep within the Earth under many layers of rock. The sea floor is also completely recycled every 200 million years by the ongoing process of plate tectonics and seafloor spreading, leaving no useful indications beneath the ocean. With the fairly significant evidence that scientists have accumulated, several mechanisms have been proposed for the extinction event, including both catastrophic and gradual processes (similar to those theorized for the Cretaceous–Paleogene extinction event). The former group includes one or more large bolide impact events, increased volcanism, and sudden release of methane from the sea floor, either due to dissociation of methane hydrate deposits or metabolism of organic carbon deposits by methanogenic microbes. The latter group includes sea level change, increasing anoxia, 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.[45]

Impact event

Artist's impression of a major impact event: A collision between Earth and an asteroid a few kilometres in diameter would release as much energy as several million nuclear weapons detonating.
Evidence that an impact event may have caused the Cretaceous–Paleogene extinction event (Cretaceous-Tertiary) 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 and for 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;[86][87] fullerenes trapping extraterrestrial noble gases;[88] meteorite fragments in Antarctica;[89] and grains rich in iron, nickel and silicon, which may have been created by an impact.[90] However, the accuracy of most of these claims has been challenged.[91][92][93][94] Quartz from Graphite Peak in Antarctica, for example, once considered "shocked", has been re-examined by optical and transmission electron microscopy. The observed features were concluded to be not due to shock, but rather to plastic deformation, consistent with formation in a tectonic environment such as volcanism.[95]
An impact crater on the sea floor 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 ocean as it is to hit land. However, Earth has no ocean-floor crust more than 200 million years old because the "conveyor belt" process of seafloor spreading and subduction destroys it within that time. Craters produced by very large impacts may be masked by extensive flood basalting from below after the crust is punctured or weakened.[96] Subduction should not, however, be entirely accepted as an explanation of why no firm evidence can be found: as with the K-T event, an ejecta blanket stratum rich in siderophilic elements (such as iridium) would be expected to be seen in formations from the time.
One attraction of large impact theories is that theoretically they could trigger other cause-considered extinction-paralleling phenomena,[clarification needed][97] such as the Siberian Traps eruptions (see below) as being either an impact site[98] or the antipode of an impact site.[97][99] 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.

Possible impact sites

Several possible impact craters have been proposed as the site of an impact causing the P–Tr extinction, including the Bedout structure off the northwest coast of Australia[87] and the hypothesized Wilkes Land crater of East Antarctica.[100][101] In each case, the idea that an impact was responsible has not been proven and has been widely criticized. In the case of Wilkes Land, the age of this sub-ice geophysical feature is very uncertain – it may be later than the Permian–Triassic extinction.
The Araguainha crater in Brazil has been most recently dated to 254.7 ± 2.5 million years ago, overlapping with estimates for the Permo-Triassic boundary.[102] Much of the local rock was oil shale. The estimated energy released by the Araguainha impact is insufficient to be a direct cause of the global mass extinction, but the colossal local earth tremors would have released huge amounts of oil and gas from the shattered rock. The resulting sudden global warming might have precipitated the Permian–Triassic extinction event.[103]

Volcanism

The final stages of the Permian had two flood basalt events. A small one, the Emeishan Traps in China, occurred at the same time as the end-Guadalupian extinction pulse, in an area close to the equator at the time.[104][105] 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.[106][107][108] The date of the Siberian Traps eruptions and the extinction event are in good agreement.[21][109]
The Emeishan and Siberian Traps eruptions may have caused dust clouds and 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. The eruptions may also have caused acid rain when the aerosols washed out of the atmosphere. That may have killed land plants and molluscs and planktonic organisms which had calcium carbonate shells. The eruptions would also have emitted carbon dioxide, causing global warming. When all of the dust clouds and aerosols washed out of the atmosphere, the excess carbon dioxide would have remained and the warming would have proceeded without any mitigating effects.[97]
The Siberian Traps had unusual features that made them even more dangerous. 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.[110] 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.[97]
In January 2011, a team, led by Stephen Grasby of the Geological Survey of Canada—Calgary, 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 Buchanan Lake. 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."[111][112] 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."[113] In 2013, QY Tang reported the total amounts of important volatiles emitted from the Siberian Traps are 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.[114]
In 2015, evidence and a timeline indicated the extinction was caused by events in the Large Igneous Province of the Siberian Traps.[115][116][117][118][119]

Methane hydrate gasification

Further information: Arctic methane release
Scientists have found worldwide evidence of a swift decrease of about 1% in the 13C/12C isotope ratio in carbonate rocks from the end-Permian.[50][120] This is the first, largest, and most rapid of a series of negative and positive excursions (decreases and increases in 13C/12C ratio) that continues until the isotope ratio abruptly stabilised in the middle Triassic, followed soon afterwards by the recovery of calcifying life forms (organisms that use calcium carbonate to build hard parts such as shells).[16]
A variety of factors may have contributed to this drop in the 13C/12C ratio, but most turn out to be insufficient to account fully for the observed amount:[121]
  • Gases from volcanic eruptions have a 13C/12C ratio about 0.5 to 0.8% below standard (δ13C about −0.5 to −0.8%), but an assessment made in 1995 concluded that the amount required to produce a reduction of about 1.0% worldwide requires eruptions greater by orders of magnitude than any for which evidence has been found.[122] (However, this analysis addressed only CO2 produced by the magma itself, not from interactions with carbon bearing sediments, as later proposed.)
  • A reduction in organic activity would extract 12C more slowly from the environment and leave more of it to be incorporated into sediments, thus reducing the 13C/12C ratio. Biochemical processes preferentially use the lighter isotopes since chemical reactions are ultimately driven by electromagnetic forces between atoms and lighter isotopes respond more quickly to these forces, but a study of a smaller drop of 0.3 to 0.4% in 13C/12C (δ13C −3 to −4 ‰) at the Paleocene-Eocene Thermal Maximum (PETM) concluded that even transferring all the organic carbon (in organisms, soils, and dissolved in the ocean) into sediments would be insufficient: even such a large burial of material rich in 12C would not have produced the 'smaller' drop in the 13C/12C ratio of the rocks around the PETM.[122]
  • Buried sedimentary organic matter has a 13C/12C ratio 2.0 to 2.5% below normal (δ13C −2.0 to −2.5%). Theoretically, if the sea level fell sharply, shallow marine sediments would be exposed to oxidization. But 6500–8400 gigatons (1 gigaton = 109 metric tons) of organic carbon would have to be oxidized and returned to the ocean-atmosphere system within less than a few hundred thousand years to reduce the 13C/12C ratio by 1.0%, which is not thought to be a realistic possibility.[48] Moreover, sea levels were rising rather than falling at the time of the extinction.[123]
  • Rather than a sudden decline in sea level, intermittent periods of ocean-bottom hyperoxia and anoxia (high-oxygen and low- or zero-oxygen conditions) may have caused the 13C/12C ratio fluctuations in the Early Triassic;[16] and global anoxia may have been responsible for the end-Permian blip. The continents of the end-Permian and early Triassic were more clustered in the tropics than they are now, and large tropical rivers would have dumped sediment into smaller, partially enclosed ocean basins at low latitudes. Such conditions favor oxic and anoxic episodes; oxic/anoxic conditions would result in a rapid release/burial, respectively, of large amounts of organic carbon, which has a low 13C/12C ratio because biochemical processes use the lighter isotopes more.[124] That or another organic-based reason may have been responsible for both that and a late Proterozoic/Cambrian pattern of fluctuating 13C/12C ratios.[16]
Other hypotheses include mass oceanic poisoning releasing vast amounts of CO2[125] and a long-term reorganisation of the global carbon cycle.[121]
Prior to consideration of the inclusion of roasting carbonate sediments by volcanism, the only proposed mechanism sufficient to cause a global 1% reduction in the 13C/12C ratio was the release of methane from methane clathrates,.[48] Carbon-cycle models confirm that it would have had enough effect to produce the observed reduction.[121][125] Methane clathrates, also known as methane hydrates, consist of methane molecules trapped in cages of water molecules. The methane, produced by methanogens (microscopic single-celled organisms), has a 13C/12C ratio about 6.0% below normal (δ13C −6.0%). At the right combination of pressure and temperature, it gets trapped in clathrates fairly close to the surface of permafrost and in much larger quantities at continental margins (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.[126]
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.[127] A vast release of methane might cause significant global warming since methane is a very powerful greenhouse gas. Strong evidence suggests the global temperatures increased by about 6 °C (10.8 °F) near the equator and therefore by more at higher latitudes: a sharp decrease in oxygen isotope ratios (18O/16O);[128] 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.[129]
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,[16] but would it also have to be reburied at an unrealistically high rate to account for the rapid increases in the 13C/12C ratio (episodes of high positive δ13C) throughout the early Triassic before it was released again several times.[16]

Anoxia

See also: Anoxic event
Evidence for widespread ocean anoxia (severe deficiency of oxygen) and euxinia (presence of hydrogen sulfide) is found from the Late Permian to the Early Triassic. 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.[130] However, in some sites, including Meishan, China, and eastern Greenland, evidence for anoxia precedes the extinction.[131][132] 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 hydrogen sulfide was present even in shallow waters.
This spread of toxic, oxygen-depleted water would have been devastating for marine life, producing widespread die-offs. Models of ocean chemistry show that anoxia and euxinia would have been closely associated with hypercapnia (high levels of carbon dioxide).[133] 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 after the extinction. Models also show that anoxic events can cause catastrophic hydrogen sulfide emissions into the atmosphere (see below).[134]
The sequence of events leading to anoxic oceans may have been triggered by carbon dioxide emissions from the eruption of the Siberian Traps.[134] 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.

Hydrogen sulfide emissions

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. 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.[134] 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:[citation needed] many show deformities that could have been caused by ultraviolet radiation, which would have been more intense after hydrogen sulfide emissions weakened the ozone layer.

The supercontinent Pangaea

Map of Pangaea showing where today's continents were at the Permian–Triassic boundary
About halfway through the Permian (in the Kungurian age of the Permian's Cisuralian epoch), all the continents joined to form the supercontinent Pangaea, surrounded by the superocean Panthalassa, although blocks that are now parts of Asia did not join the supercontinent until very late in the Permian.[135] The configuration severely decreased the extent of shallow aquatic environments, the most productive part of the seas, and it exposed formerly isolated organisms of the rich continental shelves to competition from invaders. Pangaea's formation would also have altered both oceanic circulation and atmospheric weather patterns, creating seasonal monsoons near the coasts and an arid climate in the vast continental interior.[citation needed]
Marine life suffered very high but not catastrophic rates of extinction after the formation of Pangaea (see the diagram "Marine genus biodiversity" at the top of this article), almost as high as in some of the "Big Five" mass extinctions. The formation of Pangaea seems not to have caused a significant rise in extinction levels on land, and, in fact, most of the advance of the therapsids and increase in their diversity seems to have occurred in the late Permian, after Pangaea was almost complete. Thus, it seems likely that Pangaea initiated a long period of increased marine extinctions but was not directly responsible for the "Great Dying" and the end of the Permian.

Microbes

A hypothesis published in 2014 posits that a genus of anaerobic methanogenic archaea known as Methanosarcina was responsible for the event.[136] 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 an enzymes involved in producing methane.[136] On the other hand, in the canonical Meishan sections, the Nickel concentration increases somewhat after the δ13C concentrations have begun to fall.[137]

Combination of causes

Possible causes supported by strong evidence appear to describe a sequence of catastrophes, each worse than the last: the Siberian Traps eruptions were bad enough alone, but because they occurred near coal beds and the continental shelf, they also triggered very large releases of carbon dioxide and methane.[65] 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.[65]
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;[16] 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.[138]

See also

References


  • 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.
    1. Zhang R, Follows, MJ, Grotzinger, JP, & Marshall J (2001). "Could the Late Permian deep ocean have been anoxic?". Paleoceanography 16 (3): 317–329. Bibcode:2001PalOc..16..317Z. doi:10.1029/2000PA000522.

    Further reading

    • Over, Jess (editor), Understanding Late Devonian and Permian–Triassic Biotic and Climatic Events, (Volume 20 in series Developments in Palaeontology and Stratigraphy (2006). The state of the inquiry into the extinction events.
    • Sweet, Walter C. (editor), Permo–Triassic Events in the Eastern Tethys : Stratigraphy Classification and Relations with the Western Tethys (in series World and Regional Geology)

    External links

  • ""Great Dying" Lasted 200,000 Years". National Geographic. 23 November 2011. Retrieved 1 April 2014.
  • "How a Single Act of Evolution Nearly Wiped Out All Life on Earth". ScienceDaily. 1 April 2014. Retrieved 1 April 2014.
  • Shen S.-Z.; et al. (2011). "Calibrating the End-Permian Mass Extinction". Science 334 (6061): 1367–1372. Bibcode:2011Sci...334.1367S. doi:10.1126/science.1213454. PMID 22096103.
  • Benton M J (2005). When life nearly died: the greatest mass extinction of all time. London: Thames & Hudson. ISBN 0-500-28573-X.
  • Carl T. Bergstrom; Lee Alan Dugatkin (2012). Evolution. Norton. p. 515. ISBN 978-0-393-92592-0.
  • Sahney, S.; Benton, M. J. (2008). "Recovery from the most profound mass extinction of all time". Proceedings of the Royal Society: Biological 275 (1636): 759–65. doi:10.1098/rspb.2007.1370. PMC 2596898. PMID 18198148.
  • Labandeira CC, Sepkoski JJ (1993). "Insect diversity in the fossil record". Science 261 (5119): 310–315. Bibcode:1993Sci...261..310L. doi:10.1126/science.11536548. PMID 11536548.
  • Sole RV, Newman M (2003). "Extinctions and Biodiversity in the Fossil Record". In Canadell JG, Mooney, HA. Encyclopedia of Global Environmental Change, The Earth System – Biological and Ecological Dimensions of Global Environmental Change (Volume 2). New York: Wiley. pp. 297–391. ISBN 0-470-85361-1.
  • "It Took Earth Ten Million Years to Recover from Greatest Mass Extinction". ScienceDaily. 27 May 2012. Retrieved 28 May 2012.
  • Jin YG, Wang Y, Wang W, Shang QH, Cao CQ, Erwin DH (2000). "Pattern of Marine Mass Extinction Near the Permian–Triassic Boundary in South China". Science 289 (5478): 432–436. Bibcode:2000Sci...289..432J. doi:10.1126/science.289.5478.432. PMID 10903200.
  • Yin H, Zhang K, Tong J, Yang Z, Wu S. The Global Stratotype Section and Point (GSSP) of the Permian-Triassic Boundary. Episodes 24. pp. 102–114.
  • Yin HF, Sweets WC, Yang ZY, Dickins JM (1992). "Permo-Triassic events in the eastern Tethys–an overview". In Sweet WC. Permo-Triassic events in the eastern Tethys: stratigraphy, classification, and relations with the western Tethys. Cambridge, UK: Cambridge University Press. pp. 1–7. ISBN 0-521-54573-0.
  • Darcy E. Ogdena and 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.
  • David L. Chandler – MIT News Office (31 March 2014). "Ancient whodunit may be solved: The microbes did it!". MIT News.
  • 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". Science 305 (5683): 506–9. doi:10.1126/science.1097023. PMID 15273391.
  • Benton, M. J. (2012). "No gap in the Middle Permian record of terrestrial vertebrates". Geology 40. doi:10.1130/g32669.1.
  • McElwain, J. C.; Punyasena, S. W. (2007). "Mass extinction events and the plant fossil record". Trends in Ecology & Evolution 22 (10): 548–557. doi:10.1016/j.tree.2007.09.003. PMID 17919771.
  • Retallack, G. J.; Veevers, J. J.; Morante, R. (1996). "Global coal gap between Permian–Triassic extinctions and middle Triassic recovery of peat forming plants". GSA Bulletin 108 (2): 195–207. doi:10.1130/0016-7606(1996)108<0195:gcgbpt>2.3.CO;2.
  • Erwin, D.H (1993). The Great Paleozoic Crisis: Life and Death in the Permian. New York: Columbia University Press. ISBN 0-231-07467-0.
  • Burgess, SD (2014). "High-precision timeline for Earth's most severe extinction". Nature 111 (9): 3316–3321. doi:10.1073/pnas.1317692111. PMC 3948271. PMID 24516148.
  • Magaritz M (1989). "13C minima follow extinction events: a clue to faunal radiation". Geology 17 (4): 337–340. Bibcode:1989Geo....17..337M. doi:10.1130/0091-7613(1989)017<0337:cmfeea>2.3.CO;2.
  • Krull SJ, Retallack JR (2000). "13C depth profiles from paleosols across the Permian–Triassic boundary: Evidence for methane release". GSA Bulletin 112 (9): 1459–1472. Bibcode:2000GSAB..112.1459K. doi:10.1130/0016-7606(2000)112<1459:cdpfpa>2.0.CO;2. ISSN 0016-7606.
  • 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): 175–190. doi:10.1016/S0009-2541(00)00368-5.
  • Musashi M, Isozaki Y, Koike T, Kreulen R (2001). "Stable carbon isotope signature in mid-Panthalassa shallow-water carbonates across the Permo–Triassic boundary: evidence for 13C-depleted ocean". Earth Planet. Sci. Lett. 193: 9–20. Bibcode:2001E&PSL.191....9M. doi:10.1016/S0012-821X(01)00398-3.
  • 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: 175–190. doi:10.1016/S0009-2541(00)00368-5.
  • H Visscher, H Brinkhuis, D L Dilcher, W C Elsik, Y Eshet, C V Looy, M R Rampino, and A Traverse (1996). "The terminal Paleozoic fungal event: Evidence of terrestrial ecosystem destabilization and collapse". Proceedings of the National Academy of Sciences 93 (5): 2155–2158. Bibcode:1996PNAS...93.2155V. doi:10.1073/pnas.93.5.2155. PMC 39926. PMID 11607638.
  • Foster, C.B.; Stephenson, M.H.; Marshall, C.; Logan, G.A.; Greenwood, P.F. (2002). "A Revision Of Reduviasporonites Wilson 1962: Description, Illustration, Comparison And Biological Affinities". Palynology 26 (1): 35–58. doi:10.2113/0260035.
  • López-Gómez, J. and 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.
  • 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 162 (4): 7879–7883. Bibcode:2001PNAS...98.7879L. doi:10.1073/pnas.131218098. PMC 35436. PMID 11427710. See image 2
  • 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". Science 307 (5710): 709–714. Bibcode:2005Sci...307..709W. doi:10.1126/science.1107068. PMID 15661973.
  • Retallack, G.J.; Smith, R.M.H.; Ward, P.D. (2003). "Vertebrate extinction across Permian-Triassic boundary in Karoo Basin, South Africa". Bulletin of the Geological Society of America 115 (9): 1133–1152. Bibcode:2003GSAB..115.1133R. doi:10.1130/B25215.1.
  • Sephton, M. A.; Visscher, H.; Looy, C. V.; Verchovsky, A. B.; Watson, J. S. (2009). "Chemical constitution of a Permian-Triassic disaster species". Geology 37 (10): 875–878. doi:10.1130/G30096A.1.
  • 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.
  • Wang, S.C.; Everson, P.J. (2007). "Confidence intervals for pulsed mass extinction events". Paleobiology 33 (2): 324–336. doi:10.1666/06056.1.
  • 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.
  • Retallack, G.J.; Metzger, C.A.; Greaver, T.; Jahren, A.H.; Smith, R.M.H.; Sheldon, N.D. (2006). "Middle-Late Permian mass extinction on land". Bulletin of the Geological Society of America 118 (11–12): 1398–1411. Bibcode:2006GSAB..118.1398R. doi:10.1130/B26011.1.
  • Stanley SM & Yang X (1994). "A Double Mass Extinction at the End of the Paleozoic Era". Science 266 (5189): 1340–1344. Bibcode:1994Sci...266.1340S. doi:10.1126/science.266.5189.1340. PMID 17772839.
  • Retallack, G.J., Metzger, C.A., Jahren, A.H., Greaver, T., Smith, R.M.H., and Sheldon, N.D (November–December 2006). "Middle-Late Permian mass extinction on land". GSA Bulletin 118 (11/12): 1398–1411. Bibcode:2006GSAB..118.1398R. doi:10.1130/B26011.1.
  • Ota, A, and Isozaki, Y. (March 2006). "Fusuline biotic turnover across the Guadalupian–Lopingian (Middle–Upper Permian) boundary in mid-oceanic carbonate buildups: Biostratigraphy of accreted limestone in Japan". Journal of Asian Earth Sciences 26 (3–4): 353–368. Bibcode:2006JAESc..26..353O. doi:10.1016/j.jseaes.2005.04.001.
  • Shen, S., and 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.
  • Wang, X-D, and Sugiyama, T. (December 2000). "Diversity and extinction patterns of Permian coral faunas of China". Lethaia 33 (4): 285–294. doi:10.1080/002411600750053853.
  • Racki G (1999). "Silica-secreting biota and mass extinctions: survival processes and patterns". Palaeogeography, Palaeoclimatology, Palaeoecology 154 (1–2): 107–132. doi:10.1016/S0031-0182(99)00089-9.
  • Bambach, R.K.; Knoll, A.H.; Wang, S.C. (December 2004). "Origination, extinction, and mass depletions of marine diversity". Paleobiology 30 (4): 522–542. doi:10.1666/0094-8373(2004)030<0522:oeamdo>2.0.CO;2. ISSN 0094-8373.
  • Knoll, A.H. (2004). "Biomineralization and evolutionary history. In: P.M. Dove, J.J. DeYoreo and S. Weiner (Eds), Reviews in Mineralogy and Geochemistry," (PDF).
  • Stanley, S.M. (2008). "Predation defeats competition on the seafloor". Paleobiology 34 (1): 1–21. doi:10.1666/07026.1. Retrieved 2008-05-13.
  • Stanley, S.M. (2007). "An Analysis of the History of Marine Animal Diversity". Paleobiology 33 (sp6): 1–55. doi:10.1666/06020.1.
  • Erwin DH (1993). The great Paleozoic crisis; Life and death in the Permian. Columbia University Press. ISBN 0-231-07467-0.
  • 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.
  • 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.
  • "Permian : The Marine Realm and The End-Permian Extinction". paleobiology.si.edu. Retrieved 2016-01-26.
  • "Permian extinction". Encyclopedia Britannica. Retrieved 2016-01-26.
  • 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.
  • 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.
  • Villier, L.; Korn, D. (Oct 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.
  • Saunders, W. B.; Greenfest-Allen, E.; Work, D. M.; Nikolaeva, S. V. (2008). "Morphologic and taxonomic history of Paleozoic ammonoids in time and morphospace". Paleobiology 34 (1): 128–154. doi:10.1666/07053.1.
  • "The Dino Directory – Natural History Museum".
  • Cascales-Miñana, B.; Cleal, C. J. (2011). "Plant fossil record and survival analyses". Lethaia 45: no–no. doi:10.1111/j.1502-3931.2011.00262.x.
  • Retallack, GJ (1995). "Permian–Triassic life crisis on land". Science 267 (5194): 77–80. Bibcode:1995Sci...267...77R. doi:10.1126/science.267.5194.77. PMID 17840061.
  • Looy, CV Brugman WA Dilcher DL & Visscher H (1999). "The delayed resurgence of equatorial forests after the Permian–Triassic ecologic crisis". Proceedings of the National Academy of Sciences of the United States of America 96 (24): 13857–13862. Bibcode:1999PNAS...9613857L. doi:10.1073/pnas.96.24.13857. PMC 24155. PMID 10570163.
  • 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. doi:10.1016/S0031-0182(01)00413-8.
  • Maxwell, W. D. (1992). "Permian and Early Triassic extinction of non-marine tetrapods". Palaeontology 35: 571–583.
  • Erwin DH (1990). "The End-Permian Mass Extinction". Annual Review of Ecology and Systematics 21: 69–91. doi:10.1146/annurev.es.21.110190.000441.
  • "Bristol University – News – 2008: Mass extinction".
  • Knoll, A.H., Bambach, R.K., Payne, J.L., Pruss, S., and Fischer, W.W. (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 2008-07-04.
  • 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.
  • 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. doi:10.1016/0031-0182(94)00119-S. Retrieved 21 February 2012.
  • Chinsamy-Turan (2012). Anusuya, ed. Forerunners of mammals : radiation, histology, biology. Bloomington: Indiana University Press. ISBN 978-0-253-35697-0.
  • 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. doi:10.1073/pnas.0404472101. ISSN 0027-8424. PMC 516500. PMID 15282373.
  • Lehrmann, Daniel J.; Ramezani, Jahandar; Bowring, Samuel A.; Martin, Mark W.; Montgomery, Paul; Enos, Paul; Payne, Jonathan L.; Orchard, Michael J.; Wang Hongmei; Wei Jiayong (December 2006). "Timing of recovery from the end-Permian extinction: Geochronologic and biostratigraphic constraints from south China". Geology 34 (12): 1053–1056. Bibcode:2006Geo....34.1053L. doi:10.1130/G22827A.1.
  • Yadong Sun1,2,*, Michael M. Joachimski3, Paul B. Wignall2, Chunbo Yan1, Yanlong Chen4, Haishui Jiang1, Lina Wang1, Xulong Lai1 (2012). "Lethally Hot Temperatures During the Early Triassic Greenhouse". Science 338 (6105): 366–370. Bibcode:2012Sci...338..366S. doi:10.1126/science.1224126. PMID 23087244.
  • During the greatest mass extinction in Earth’s history the world’s oceans reached 40 °C (104 °F) – lethally hot.
  • Lau, Kimberly V.; Maher, Kate; Altiner, Demir; Kelley, Brian M.; Kump, Lee R.; Lehrmann, Daniel J.; Silva-Tamayo, Juan Carlos; Weaver, Karrie L.; Yu, Meiyi; Payne, Jonathan L. (2016). "Marine anoxia and delayed Earth system recovery after the end-Permian extinction". Proceedings of the National Academy of Sciences 113 (9): 2360–2365. doi:10.1073/pnas.1515080113.
  • Ward PD, Montgomery DR, & Smith R (2000). "Altered river morphology in South Africa related to the Permian–Triassic extinction". Science 289 (5485): 1740–1743. Bibcode:2000Sci...289.1740W. doi:10.1126/science.289.5485.1740. PMID 10976065.
  • Hallam, A; Wignall, P B (1997). Mass Extinctions and their Aftermath. Oxford University Press. ISBN 978-0-19-854916-1.
  • Rodland, DL & Bottjer, DJ (2001). "Biotic Recovery from the End-Permian Mass Extinction: Behavior of the Inarticulate Brachiopod Lingula as a Disaster Taxon". PALAIOS 16 (1): 95–101. doi:10.1669/0883-1351(2001)016<0095:brftep>2.0.CO;2. ISSN 0883-1351.
  • Zi-qiang W (1996). "Recovery of vegetation from the terminal Permian mass extinction in North China". Review of Palaeobotany and Palynology 91 (1–4): 121–142. doi:10.1016/0034-6667(95)00069-0.
  • 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.
  • Clapham, M.E., Bottjer, D.J. and Shen, S. (2006). "Decoupled diversity and ecology during the end-Guadalupian extinction (late Permian)". Geological Society of America Abstracts with Programs 38 (7): 117. Retrieved 2008-03-28.
  • 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.
  • 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.
  • Botha, J., and Smith, R.M.H. (2007). "Lystrosaurus species composition across the Permo–Triassic boundary in the Karoo Basin of South Africa". Lethaia 40 (2): 125–137. doi:10.1111/j.1502-3931.2007.00011.x. Retrieved 2008-07-02. Full version online at "Lystrosaurus species composition across the Permo–Triassic boundary in the Karoo Basin of South Africa" (PDF). Retrieved 2008-07-02.
  • Benton, M.J. (2004). Vertebrate Paleontology. Blackwell Publishers. xii–452. ISBN 0-632-05614-2.
  • Ruben, J.A., and Jones, T.D. (2000). "Selective Factors Associated with the Origin of Fur and Feathers". American Zoologist 40 (4): 585–596. doi:10.1093/icb/40.4.585.
  • Yates AM & Warren AA (2000). "The phylogeny of the 'higher' temnospondyls (Vertebrata: Choanata) and its implications for the monophyly and origins of the Stereospondyli". Zoological Journal of the Linnean Society 128 (1): 77–121. doi:10.1111/j.1096-3642.2000.tb00650.x. Archived from the original on 2007-10-01. Retrieved 2008-01-18.
  • 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.
  • 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.
  • 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.
  • 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.
  • 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. Retrieved 2007-10-22.
  • 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.
  • 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.
  • 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.
  • 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.
  • Langenhorst F, Kyte FT & Retallack GJ (2005). "Reexamination of quartz grains from the Permian–Triassic boundary section at Graphite Peak, Antarctica" (PDF). Lunar and Planetary Science Conference XXXVI. Retrieved 2007-07-13.
  • 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. doi:10.1016/S0012-821X(02)00824-5.
  • White RV (2002). "Earth's biggest 'whodunnit': unravelling the clues in the case of the end-Permian mass extinction" (PDF). Phil. Trans. Royal Society of London 360 (1801): 2963–2985. Bibcode:2002RSPTA.360.2963W. doi:10.1098/rsta.2002.1097. PMID 12626276. Retrieved 2008-01-12.
  • AHager, Bradford H (2001). "Giant Impact Craters Lead To Flood Basalts: A Viable Model". CCNet 33/2001: Abstract 50470.
  • Hagstrum, Jonathan T (2001). "Large Oceanic Impacts As The Cause Of Antipodal Hotspots And Global Mass Extinctions". CCNet 33/2001: Abstract 50288.
  • von Frese RR, Potts L, Gaya-Pique L, Golynsky AV, Hernandez O, Kim J, Kim H & Hwang J (2006). "Permian–Triassic mascon in Antarctica". Eos Trans. AGU, Jt. Assem. Suppl. 87 (36): Abstract T41A–08. Retrieved 2007-10-22.
  • Von Frese, R.R.B.; L. V. Potts; S. B. Wells; T. E. Leftwich; H. R. Kim; J. W. Kim; A. V. Golynsky; O. Hernandez; L. R. Gaya-Piqué (2009). "GRACE gravity evidence for an impact basin in Wilkes Land, Antarctica". Geochem. Geophys. Geosyst. 10 (2): Q02014. Bibcode:2009GGG....1002014V. doi:10.1029/2008GC002149.
  • Tohver E., Lana C., Cawood P.A., Fletcher I.R., Jourdan F., Sherlock S., Rasmussen B., Trindade R.I.F., Yokoyama E., Filho C.R. Souza, Marangoni Y. "Geochronological constraints on the age of a Permo–Triassic impact event: U–Pb and 40Ar/39Ar results for the 40 km Araguainha structure of central Brazil". Geochimica et Cosmochimica Acta 86: 214–227. Bibcode:2012GeCoA..86..214T. doi:10.1016/j.gca.2012.03.005.
  • Biggest extinction in history caused by climate-changing meteor. University of Western Australia University News Wednesday, 31 July 2013. http://www.news.uwa.edu.au/201307315921/international/biggest-extinction-history-caused-climate-changing-meteor
  • Zhou, M-F., Malpas, J, Song, X-Y, Robinson, PT, Sun, M, Kennedy, AK, Lesher, CM & Keays, RR (2002). "A temporal link between the Emeishan large igneous province (SW China) and the end-Guadalupian mass extinction". Earth and Planetary Science Letters 196 (3–4): 113–122. Bibcode:2002E&PSL.196..113Z. doi:10.1016/S0012-821X(01)00608-2.
  • Wignall, Paul B.; et al. (2009). "Volcanism, Mass Extinction, and Carbon Isotope Fluctuations in the Middle Permian of China". Science 324 (5931): 1179–1182. Bibcode:2009Sci...324.1179W. doi:10.1126/science.1171956. PMID 19478179.
  • Andy Saunders, Marc Reichow (2009). "The Siberian Traps – Area and Volume". Retrieved 2009-10-18.
  • Andy Saunders and Marc Reichow (January 2009). "The Siberian Traps and the End-Permian mass extinction: a critical review" (PDF). Chinese Science Bulletin (Springer) 54 (1): 20–37. doi:10.1007/s11434-008-0543-7.
  • Reichow, MarcK.; Pringle, M.S.; Al'Mukhamedov, A.I.; Allen, M.B.; Andreichev, V.L.; Buslov, M.M.; Davies, C.E.; Fedoseev, G.S.; Fitton, J.G.; Inger, S.; Medvedev, A.Ya.; Mitchell, C.; Puchkov, V.N.; Safonova, I.Yu.; Scott, R.A.; Saunders, A.D. (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: 9–20. Bibcode:2009E&PSL.277....9R. doi:10.1016/j.epsl.2008.09.030.
  • 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: 75–91. Bibcode:2003E&PSL.214...75K. doi:10.1016/S0012-821X(03)00347-9.
  • "Volcanism".
  • Dan Verango (January 24, 2011). "Ancient mass extinction tied to torched coal". USA Today.
  • Stephen E. Grasby, Hamed Sanei & Benoit Beauchamp (January 23, 2011). "Catastrophic dispersion of coal fly ash into oceans during the latest Permian extinction". Nature Geoscience 4 (2): 104–107. Bibcode:2011NatGe...4..104G. doi:10.1038/ngeo1069.
  • "Researchers find smoking gun of world's biggest extinction; Massive volcanic eruption, burning coal and accelerated greenhouse gas choked out life". University of Calgary. January 23, 2011. Retrieved 2011-01-26.
  • Yang, QY (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. doi:10.1016/j.chemgeo.2012.08.031.
  • Burgess, Seth D.; Bowring, Samuel; Shen, Shu-zhong (2014-03-04). "High-precision timeline for Earth’s most severe extinction". Proceedings of the National Academy of Sciences 111 (9): 3316–3321. doi:10.1073/pnas.1317692111. ISSN 0027-8424. PMC 3948271. PMID 24516148.
  • "Earth’s worst extinction "inescapably" tied to Siberian Traps, CO2, and climate change". Skeptical Science. Retrieved 2016-03-11.
  • Black, Benjamin A.; Weiss, Benjamin P.; Elkins-Tanton, Linda T.; Veselovskiy, Roman V.; Latyshev, Anton (2015-04-30). "Siberian Traps volcaniclastic rocks and the role of magma-water interactions". Geological Society of America Bulletin 127: B31108.1. doi:10.1130/B31108.1. ISSN 0016-7606.
  • Burgess, Seth D.; Bowring, Samuel A. (2015-08-01). "High-precision geochronology confirms voluminous magmatism before, during, and after Earth’s most severe extinction". Science Advances 1 (7): e1500470. doi:10.1126/sciadv.1500470. ISSN 2375-2548. PMC 4643808. PMID 26601239.
  • Fischman, Josh. "Giant Eruptions and Giant Extinctions [Video]". Scientific American. Retrieved 2016-03-11.
  • 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.
  • Berner, R.A. (2002). "Examination of hypotheses for the Permo-Triassic boundary extinction by carbon cycle modeling". Proceedings of the National Academy of Sciences 99 (7): 4172–4177. Bibcode:2002PNAS...99.4172B. doi:10.1073/pnas.032095199. PMC 123621. PMID 11917102.
  • Dickens GR, O'Neil JR, Rea DK & Owen RM (1995). "Dissociation of oceanic methane hydrate as a cause of the carbon isotope excursion at the end of the Paleocene". Paleoceanography 10 (6): 965–71. Bibcode:1995PalOc..10..965D. doi:10.1029/95PA02087.
  • White, R. V. (2002). "Earth's biggest 'whodunnit': Unravelling the clues in the case of the end-Permian mass extinction". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 360 (1801): 2963–85. doi:10.1098/rsta.2002.1097. PMID 12626276.
  • Schrag, D.P., Berner, R.A., Hoffman, P.F., and Halverson, G.P. (2002). "On the initiation of a snowball Earth". Geochemistry Geophysics Geosystems 3 (6): 1036. 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.
  • Benton, M.J.; Twitchett, R.J. (2003). "How to kill (almost) all life: the end-Permian extinction event". Trends in Ecology & Evolution 18 (7): 358–365. doi:10.1016/S0169-5347(03)00093-4.
  • Dickens GR (2001). "The potential volume of oceanic methane hydrates with variable external conditions". Organic Geochemistry 32 (10): 1179–1193. doi:10.1016/S0146-6380(01)00086-9.
  • Reichow MK, Saunders AD, White RV, Pringle MS, Al'Muhkhamedov AI, Medvedev AI & Kirda NP (2002). "40Ar/39Ar Dates from the West Siberian Basin: Siberian Flood Basalt Province Doubled". Science 296 (5574): 1846–1849. Bibcode:2002Sci...296.1846R. doi:10.1126/science.1071671. PMID 12052954.
  • Holser WT, Schoenlaub H-P, Attrep Jr M, Boeckelmann K, Klein P, Magaritz M, Orth CJ, Fenninger A, Jenny C, Kralik M, Mauritsch H, Pak E, Schramm J-F, 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.
  • Dobruskina IA (1987). "Phytogeography of Eurasia during the early Triassic". Palaeogeography, Palaeoclimatology, Palaeoecology 58 (1–2): 75–86. doi:10.1016/0031-0182(87)90007-1.
  • Wignall, P.B.; Twitchett, R.J. (2002). "Extent, duration, and nature of the Permian-Triassic superanoxic event". Geological Society of America Special Papers 356: 395–413. doi:10.1130/0-8137-2356-6.395. ISBN 0-8137-2356-6.
  • 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.
  • 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". Organic Geochemistry 43: 67–82. doi:10.1016/j.orggeochem.2011.10.010.
  • 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. doi:10.1130/G24618A (inactive 2016-03-11).
  • Kump, Lee; Alexander Pavlov; Michael A. Arthur (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.
  • The Permian – Palaeos
  • Rothman, D. H.; Fournier, G. P.; French, K. L.; Alm, E. J.; Boyle, E. A.; Cao, C.; Summons, R. E. (2014-03-31). "Methanogenic burst in the end-Permian carbon cycle". Proceedings of the National Academy of Sciences 111 (15): 5462–7. doi:10.1073/pnas.1318106111. PMC 3992638. PMID 24706773. — Lay summary: Chandler, David L.; Massachusetts Institute of Technology (March 31, 2014). "Ancient whodunit may be solved: Methane-producing microbes did it!". Science Daily.
  • 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.
  •