Quaternary extinction event

Last updated
Late Pleistocene landscape of northern Spain, by Mauricio Anton. From left to right: wild horse (Equus ferus), woolly mammoth (Mammuthus primigenius) reindeer (Rangifer tarandus) the cave or steppe lion (Panthera spelaea) and the woolly rhinoceros (Coelodonta antiquitatis) Ice age fauna of northern Spain - Mauricio Anton.jpg
Late Pleistocene landscape of northern Spain, by Mauricio Antón. From left to right: wild horse ( Equus ferus), woolly mammoth (Mammuthus primigenius) reindeer (Rangifer tarandus) the cave or steppe lion ( Panthera spelaea ) and the woolly rhinoceros (Coelodonta antiquitatis)
Mural of La Brea Tar Pits by Charles R. Knight, including sabertooth cats (Smilodon fatalis, left) ground sloths (Paramylodon harlani, right) and Columbian mammoths (Mammuthus columbi, background) La Brea Tar Pits.jpg
Mural of La Brea Tar Pits by Charles R. Knight, including sabertooth cats ( Smilodon fatalis , left) ground sloths ( Paramylodon harlani , right) and Columbian mammoths (Mammuthus columbi, background)

The latter half of the Late Pleistocene to the beginning of the Holocene (~50,000-10,000 years Before Present) [1] saw extinctions of numerous predominantly megafaunal (large) animal species (the Pleistocene megafauna), which resulted in a collapse in faunal density and diversity across the globe. The extinctions during the Late Pleistocene are differentiated from previous extinctions by the widespread absence of ecological succession to replace these extinct megafaunal species, [2] and the regime shift of previously established faunal relationships and habitats as a consequence. The timing and severity of the extinctions varied by region and are thought to have been driven by varying combinations of human and climatic factors. [2] Human impact on megafauna populations is thought to have been driven by hunting ("overkill") as well as possibly environmental alteration. [3] [4] [5] The relative importance of human vs climatic factors in the extinctions has been the subject of long-running controversy. [2]

Contents

Major extinctions occurred in Australia-New Guinea (Sahul) beginning approximately 50,000 years ago and in the Americas about 13,000 years ago, coinciding in time with the early human migrations into these regions. [6] [7] [8] Extinctions in northern Eurasia were staggered over tens of thousands of years between 50,000 and 10,000 years ago, [1] while extinctions in the Americas were virtually simultaneous, spanning only 3000 years at most. [3] [9] Overall, during Late Pleistocene about 65% of all megafaunal species worldwide became extinct, [10] rising to 72% in North America, 83% in South America and 88% in Australia, with Africa, South Asia and Southeast Asia having much lower extinctions than other regions. [11]

Extinctions by biogeographic realm

Summary

Extinctions range of the continental large and medium-sized mammals from 40,000 to 4,000 years BP in different biogeographic realms [12]
Biogeographic realm Giants
(over 1,000 kg)
Very large
(400–1,000 kg)
Large
(150–400 kg)
Moderately large
(50–150 kg)
Medium
(10–50 kg)
TotalRegions included
StartLoss%StartLoss%StartLoss%StartLoss%StartLoss%StartLoss%
Afrotropic 6−116.6%4−125%25−312%3200%69−22.9%136-75.1% Trans-Saharan Africa and Arabia
Indomalaya 5−240%6−116.7%10−110%20−315%56−11.8%97-88.2% Indian subcontinent, Southeast Asia, and southern China
Palearctic 8−8100%10−550%14−535.7%23−315%41−12.4%96-2222.9% Eurasia and North Africa
Nearctic 5−5100%10−880%26−2284.6%20−1365%25−936%86-5766% North America
Neotropic 9−9100%12−12100%17−1482%20−1155%35−514.3%93-5154% South America, Central America, South Florida, and the Caribbean
Australasia 4−4100%5−5100%6−6100%16−1381.2%25−1040%56-3867% Australia, New Guinea, New Zealand, and neighbouring islands.
Global33−2678.8%46−3167.4%86−4754.7%113−4136.3%215−2310.1%493-16834%

Introduction

The proportion of extinct large mammal species (more than or equal to 10 kg) in each country during the last 132,000 years, only counting extinctions earlier than 1000 years BP The proportion of extinct large mammal species (more than or equal to 10 kg) in each TDWG country during the last 132 000 years, only counting extinctions earlier than 1000 years BP.jpg
The proportion of extinct large mammal species (more than or equal to 10 kg) in each country during the last 132,000 years, only counting extinctions earlier than 1000 years BP

The Late Pleistocene saw the extinction of many mammals weighing more than 40 kilograms (88 lb). The proportion of megafauna extinctions is progressively larger the further the human migratory distance from Africa, with the highest extinction rates in Australia, and North and South America.

The increased extent of extinction mirrors the migration pattern of modern humans: the further away from Africa, the more recently humans inhabited the area, the less time those environments (including its megafauna) had to become accustomed to humans (and vice versa).

There are two main hypotheses to explain this extinction:

There are some inconsistencies between the current available data and the prehistoric overkill hypothesis. For instance, there are ambiguities around the timing of Australian megafauna extinctions. [13] Evidence supporting the prehistoric overkill hypothesis includes the persistence of megafauna on some islands for millennia past the disappearance of their continental cousins. For instance, ground sloths survived on the Antilles long after North and South American ground sloths were extinct, woolly mammoths died out on remote Wrangel Island 6,000 years after their extinction on the mainland, while Steller's sea cows persisted off the isolated and uninhabited Commander Islands for thousands of years after they had vanished from the continental shores of the north Pacific. [14] The later disappearance of these island species correlates with the later colonization of these islands by humans.

The original debates as to whether human arrival times or climate change constituted the primary cause of megafaunal extinctions necessarily were based on paleontological evidence coupled with geological dating techniques. Recently, genetic analyses of surviving megafaunal populations have contributed new evidence, leading to the conclusion: "The inability of climate to predict the observed population decline of megafauna, especially during the past 75,000 years, implies that human impact became the main driver of megafauna dynamics around this date." [15]

An alternative hypothesis to the theory of human responsibility is climate change associated with the last glacial period. Discredited explanations include the Younger Dryas impact hypothesis [16] and Tollmann's hypothesis that extinctions resulted from bolide impacts.

Recent research indicates that each species responded differently to environmental changes, and no one factor by itself explains the large variety of extinctions. The causes may involve the interplay of climate change, competition between species, unstable population dynamics, and human predation. [17]

Africa

Although Africa was one of the least affected regions, the region still suffered extinctions, particularly around the Late Pleistocene-Holocene transition. These extinctions were likely predominantly climatically driven by changes to grassland habitats. [18]

South Asia, Southeast Asia and East Asia

Giant tapir (Tapirus augustus) restoration MegatapirusDB.jpg
Giant tapir (Tapirus augustus) restoration
Life-sized models of Stegodon ExtinctElephant-SiwalikHills.JPG
Life-sized models of Stegodon
Palaeoloxodon namadicus fossil at Indian Museum, Kolkata, India Palaeoloxodon namadicus.JPG
Palaeoloxodon namadicus fossil at Indian Museum, Kolkata, India
Fossil jaw (Xiahe mandible) of a denisovan Xiahe mandible.jpg
Fossil jaw (Xiahe mandible) of a denisovan

The timing of extinctions on the Indian subcontinent is uncertain due to a lack of reliable dating. [20] Similar issues have been reported for Chinese sites, though there is no evidence for any of the megafaunal taxa having survived into the Holocene in that region. [21] Extinctions in Southeast Asia and South China have been proposed to be the result of environmental shift from open to closed forested habitats. [22]

Europe and northern Asia

Saiga antelope (Saiga spp.) inhabited a range from England and France to Yukon in the Late Pleistocene, diversifying into two species. S. borealis is now extinct and the critically endangered S. tatarica is now limited to the steppe in Kazakhstan and Mongolia Saiga.jpg
Saiga antelope (Saiga spp.) inhabited a range from England and France to Yukon in the Late Pleistocene, diversifying into two species. S. borealis is now extinct and the critically endangered S. tatarica is now limited to the steppe in Kazakhstan and Mongolia
Hippopotamuses (Hippopotamus spp.) inhabited Great Britain until 80,000 BCE, whence due to glacial shifts, hippopotamuses were restricted to southeastern Europe, Mediterranean islands and finally western Asia until 1,000 BCE Hippopotamus amphibius Whipsnade Zoo.jpg
Hippopotamuses (Hippopotamus spp.) inhabited Great Britain until 80,000 BCE, whence due to glacial shifts, hippopotamuses were restricted to southeastern Europe, Mediterranean islands and finally western Asia until 1,000 BCE
Reconstruction of the five phenotypes of Pleistocene wild horse. The coat colours and dimensions are based on genetic evidence and historic descriptions Wild horse reconstruction.jpg
Reconstruction of the five phenotypes of Pleistocene wild horse. The coat colours and dimensions are based on genetic evidence and historic descriptions
Elasmotherium sibiricum reconstruction Elasm062.jpg
Elasmotherium sibiricum reconstruction
Cave paintings of the woolly rhinoceros (Coelodonta antiquitatis) in Chauvet-Pont-d'Arc Cave, France 16 PanneauDesLions(CentreGauche)RhinocerosEnFuite.jpg
Cave paintings of the woolly rhinoceros (Coelodonta antiquitatis) in Chauvet-Pont-d'Arc Cave, France
The 'Gallery of Lions', representations of the Eurasian cave lion in Chauvet-Pont-d'Arc Cave, France Lions painting, Chauvet Cave (museum replica).jpg
The 'Gallery of Lions', representations of the Eurasian cave lion in Chauvet-Pont-d'Arc Cave, France
The leopard (Panthera pardus) inhabited the entire expanse of Afro-Eurasia below the 54th parallel north, from modern day Spain and the UK in the west, to South Africa in the south, and Siberia, Japan and Sundaland in the east during the Late Pleistocene ZOO Olomouc07 Panthera pardus orientalis.JPG
The leopard (Panthera pardus) inhabited the entire expanse of Afro-Eurasia below the 54th parallel north, from modern day Spain and the UK in the west, to South Africa in the south, and Siberia, Japan and Sundaland in the east during the Late Pleistocene
Cave bear (Ursus spelaeus) reconstruction Ursus spelaeus Sergiodlarosa.jpg
Cave bear (Ursus spelaeus) reconstruction
The woolly mammoth became extinct around 10,000 BCE - except for diminutive relict populations on St. Paul Island and Wrangel Island, which humans did not colonise until 3,600 BCE and 2,000 BCE respectively Mamut lanudo.jpg
The woolly mammoth became extinct around 10,000 BCE – except for diminutive relict populations on St. Paul Island and Wrangel Island, which humans did not colonise until 3,600 BCE and 2,000 BCE respectively
Model of the European straight-tusked elephant (Paleoloxodon antiquus) Reproduccion de Palaeoloxodon antiquus.jpg
Model of the European straight-tusked elephant (Paleoloxodon antiquus)

The Palearctic realm spans the entirety of the European continent and stretches into northern Asia, through the Caucasus and central Asia to northern China, Siberia and Beringia. Extinctions were more severe in Northern Eurasia than in Africa or South and Southeast Asia. These extinctions were staggered over tens of thousands of years, spanning from around 50,000 years Before Present (BP) to around 10,000 years BP, with temperate adapted species like the straight-tusked elephant and the narrow-nosed rhinoceros generally going extinct earlier than cold adapted species like the woolly mammoth and woolly rhinoceros. Climate change has been considered a probable major factor in the extinctions, possibly in combination with human hunting. [1]

North America

Extinctions in North America were concentrated at the end of the Late Pleistocene, around 13,800–11,400 years Before Present, which were coincident with the onset of the Younger Dryas cooling period, as well as the emergence of the hunter-gatherer Clovis culture. The relative importance of human and climactic factors in the North American extinctions has been the subject of significant controversy. Extinctions totalled around 35 genera. [3] The radiocarbon record for North America south of the Alaska-Yukon region has been described as "inadequate" to construct a reliable chronology. [52]

Long-horned/Giant bison (Bos latifrons), fossil bison skeleton (public display, Cincinnati Museum of Natural History & Science, Cincinnati, Ohio, United States) Bison latifrons fossil buffalo (Pleistocene; North America) 1 (15257877377).jpg
Long-horned/Giant bison (Bos latifrons), fossil bison skeleton (public display, Cincinnati Museum of Natural History & Science, Cincinnati, Ohio, United States)
Mounted skeleton of a shrub-ox (Euceratherium collinum) Preptoceras sinclairi.jpg
Mounted skeleton of a shrub-ox (Euceratherium collinum)
Life restoration of Cervalces scotti Life restoration cervalces scotti.jpg
Life restoration of Cervalces scotti
Tetrameryx shuleri restoration Tetrameryx shuleri.png
Tetrameryx shuleri restoration
A Chacoan peccary (Catagonus wagneri), believed to be the closest surviving relative of the extinct Platygonus Catagonus wagneri closeup.jpg
A Chacoan peccary (Catagonus wagneri), believed to be the closest surviving relative of the extinct Platygonus
Western camel (Camelops hesternus) reconstruction Camelops hesternus Sergiodlarosa.jpg
Western camel (Camelops hesternus) reconstruction
Life restoration of the Yukon horse (Equus lambei) Equus ferus lambei.jpg
Life restoration of the Yukon horse (Equus lambei)
Reconstruction of Mixotoxodon larenis, a toxodontid notoungulate Mixotoxodon.jpg
Reconstruction of Mixotoxodon larenis , a toxodontid notoungulate
Saber-toothed cat (Smilodon fatalis) reconstruction Smilodon fatalis Sergiodlarosa.jpg
Saber-toothed cat (Smilodon fatalis) reconstruction
Scimitar cat (Homotherium serum) reconstruction Homotherium serum.jpg
Scimitar cat (Homotherium serum) reconstruction
American lion (Panthera atrox) reconstruction Panthera leo atrox Sergiodlarosa.jpg
American lion (Panthera atrox) reconstruction
The dhole (Cuon alpinus), now restricted to the southern portions of Asia, was present from Iberia to Mexico during the Late Pleistocene 2010-kabini-dhole.jpg
The dhole (Cuon alpinus), now restricted to the southern portions of Asia, was present from Iberia to Mexico during the Late Pleistocene
Giant short-faced bear (Arctodus simus) reconstruction Arctodus simus Sergiodlarosa.jpg
Giant short-faced bear ( Arctodus simus ) reconstruction
American mastodon (Mammut americanum) reconstruction Mammut americanum Sergiodlarosa.jpg
American mastodon (Mammut americanum) reconstruction
Columbian mammoth (Mammuthus columbi) reconstruction Mammuthus columbi Sergiodlarosa.jpg
Columbian mammoth (Mammuthus columbi) reconstruction
Giant beaver (Castoroides ohioensis) skeleton displayed at the Field Museum of Natural History, Chicago, Illinois, United States Giant-beaver-fieldmuseum.jpg
Giant beaver (Castoroides ohioensis) skeleton displayed at the Field Museum of Natural History, Chicago, Illinois, United States
Skull of Paralouatta marianae, one of the two Cuban members of the extinct Antilles monkeys (Xenotrichini) Paralouatta marianae skull.jpg
Skull of Paralouatta marianae , one of the two Cuban members of the extinct Antilles monkeys (Xenotrichini)
Eremotherium laurillardi skeleton displayed at the Houston Museum of Natural Science WLA hmns Giant ground sloth 2.jpg
Eremotherium laurillardi skeleton displayed at the Houston Museum of Natural Science
Life restoration of Nothrotheriops texanus Nothrotheriops.jpg
Life restoration of Nothrotheriops texanus
Glyptotherium reconstruction Glyptotheriumm.jpg
Glyptotherium reconstruction
Californian turkey (Meleagris californica) and megafaunal Californian condor (Gymnogyps amplus) fossil displays at La Brea Tar Pits Gymnogyps amplus skeleton.jpg
Californian turkey (Meleagris californica) and megafaunal Californian condor (Gymnogyps amplus) fossil displays at La Brea Tar Pits
Teratornis merriami skeleton from the La Brea Tar Pits in flight pose Teratornis fossil.jpg
Teratornis merriami skeleton from the La Brea Tar Pits in flight pose
Reconstruction of the Cuban giant owl (Ornimegalonyx oteroi), of Pleistocene Cuba, with the carcass of a large solenodon Ornimegalonyx oteroi.jpg
Reconstruction of the Cuban giant owl (Ornimegalonyx oteroi), of Pleistocene Cuba, with the carcass of a large solenodon

North American extinctions (noted as herbivores (H) or carnivores (C)) included:

The survivors are in some ways as significant as the losses: bison (H), grey wolf (C), lynx (C), grizzly bear (C), American black bear (C), deer (e.g. caribou, moose, wapiti (elk), Odocoileus spp.) (H), pronghorn (H), white-lipped peccary (H), muskox (H), bighorn sheep (H), and mountain goat (H); the list of survivors also include species which were extirpated during the Quaternary extinction event, but recolonised at least part of their ranges during the mid-Holocene from South American relict populations, such as the cougar (C), jaguar (C), giant anteater (C), collared peccary (H), ocelot (C) and jaguarundi (C). All save the pronghorns and giant anteaters were descended from Asian ancestors that had evolved with human predators. [78] Pronghorns are the second-fastest land mammal (after the cheetah), which may have helped them elude hunters. More difficult to explain in the context of overkill is the survival of bison, since these animals first appeared in North America less than 240,000 years ago and so were geographically removed from human predators for a sizeable period of time. [79] [80] [81] Because ancient bison evolved into living bison, [82] [83] there was no continent-wide extinction of bison at the end of the Pleistocene (although the genus was regionally extirpated in many areas). The survival of bison into the Holocene and recent times is therefore inconsistent with the overkill scenario. By the end of the Pleistocene, when humans first entered North America, these large animals had been geographically separated from intensive human hunting for more than 200,000 years. Given this enormous span of geologic time, bison would almost certainly have been very nearly as naive as native North American large mammals.

The culture that has been connected with the wave of extinctions in North America is the paleo-American culture associated with the Clovis people (q.v.), who were thought to use spear throwers to kill large animals. The chief criticism of the "prehistoric overkill hypothesis" has been that the human population at the time was too small and/or not sufficiently widespread geographically to have been capable of such ecologically significant impacts. This criticism does not mean that climate change scenarios explaining the extinction are automatically to be preferred by default, however, any more than weaknesses in climate change arguments can be taken as supporting overkill. Some form of a combination of both factors could be plausible, and overkill would be a lot easier to achieve large-scale extinction with an already stressed population due to climate change.

South America

Fossil of Hippidion, a genus of horse native to South America. Hippidion 23.JPG
Fossil of Hippidion , a genus of horse native to South America.
Reconstruction of the mother and calf of Macrauchenia, a member of the extinct order Litopterna Macrauchenia (reconstruction).jpg
Reconstruction of the mother and calf of Macrauchenia , a member of the extinct order Litopterna
Skeleton of Toxodon , a member of the extinct order Notoungulata Em - Toxodon platensis - 1.jpg
Skeleton of Toxodon , a member of the extinct order Notoungulata
Reconstruction of the Dire wolf (Aenocyon dirus) The American Museum journal (c1900-(1918)) (Aenocyon dirus).jpg
Reconstruction of the Dire wolf (Aenocyon dirus)
Life restoration of Arctotherium bonariense. Arctotherium.jpg
Life restoration of Arctotherium bonariense.
Reconstruction of the gomphothere Cuvieronius Cuvieronius.jpg
Reconstruction of the gomphothere Cuvieronius
Skeleton of the giant ground sloth Megatherium Megatherium americanum complete.JPG
Skeleton of the giant ground sloth Megatherium
Reconstruction of the glytodont Doedicurus clavicaudatus , distributed in the temperate savannah and woodland of South America. Doedicurus BW.jpg
Reconstruction of the glytodont Doedicurus clavicaudatus , distributed in the temperate savannah and woodland of South America.
Fossil reconstruction of Panochthus frenzelianus with metal model. Panochthus frenzelianus.jpg
Fossil reconstruction of Panochthus frenzelianus with metal model.

South America suffered among the worst losses of the continents, with around 83% of its megafauna going extinct. [11] Extinctions are thought to have occurred in the interval 13,000–10,000 years Before Present, coincident with the end of the Antarctic Cold Reversal (a cooling period earlier and less severe than the Northern Hemisphere Younger Dryas) and the emergence of Fishtail projectile points, which became widespread across South America. Fishtail projectile points are thought to have been used in big game hunting, though direct evidence of exploitation of extinct megafauna by humans is rare. Fishtail points rapidly disappeared after the extinction of the megafauna, and were replaced by other styles more suited to hunting smaller prey. Humans have traditionally been less cited as a causal factor in the extinctions than in North America, though some recent scholarship is beginning to challenge this. [84]

The Pacific (Australasia and Oceania)

Reconstruction of a hippopotamus-sized Diprotodon Diprotodon sculpture.jpg
Reconstruction of a hippopotamus-sized Diprotodon
Some of the marsupial lions were the largest mammalian predators in Australia of their time Thylacoleo Australia 2.jpg
Some of the marsupial lions were the largest mammalian predators in Australia of their time
Reconstruction of Zygomaturus Zygomaturus BW.jpg
Reconstruction of Zygomaturus

A scarcity of reliably dated megafaunal bone deposits has made it difficult to construct timelines for megafaunal extinctions in certain areas, leading to a divide among researches about when and how megafaunal species went extinct. [116] [117]

There are at least three hypotheses regarding the extinction of the Australian megafauna:

  1. that they went extinct with the arrival of the Aboriginal Australians on the continent,
  2. that they went extinct due to natural climate change.

This theory is based on evidence of megafauna surviving until 40,000 years ago, a full 30,000 years after homo sapiens first landed in Australia, and thus that the two groups coexisted for a long time. Evidence of these animals existing at that time come from fossil records and ocean sediment. To begin with, sediment core drilled in the Indian Ocean off the SW coast of Australia indicate the existence of a fungus called Sporormiella, which survived off the dung of plant-eating mammals. The abundance of these spores in the sediment prior to 45,000 years ago indicates that many large mammals existed in the southwest Australian landscape until that point. The sediment data also indicates that the megafauna population collapsed within a few thousand years, around the 45,000 years ago, suggesting a rapid extinction event. [118] In addition, fossils found at South Walker Creek, which is the youngest megafauna site in northern Australia, indicate that at least 16 species of megafauna survived there until 40,000 years ago. Furthermore, there is no firm evidence of homo sapiens living at South Walker Creek 40,000 years ago, therefore no human cause can be attributed to the extinction of these megafauna. However, there is evidence of major environmental deterioration of South Water Creek 40,000 years ago, which may have caused the extinct event. These changes include increased fire, reduction in grasslands, and the loss of fresh water. [119] The same environmental deterioration is seen across Australia at the time, further strengthening the climate change argument. Australia's climate at the time could best be described as an overall drying of the landscape due to lower precipitation, resulting in less fresh water availability and more drought conditions. Overall, this led to changes in vegetation, increased fires, overall reduction in grasslands, and a greater competition for already scarce fresh water. [120] These environmental changes proved to be too much for the Australian megafauna to cope with, causing the extinction of 90% of megafauna species.

  1. The third hypothesis shared by some scientists is that human impacts and natural climate changes led to the extinction of Australian megafauna. About 75% of Australia is semi-arid or arid, so it makes sense that megafauna species used the same fresh water resources as humans. This competition could have led to more hunting of megafauna. [121] Furthermore, homo sapiens used fire agriculture[ clarification needed ] to burn impassable[ clarification needed ] land. This further diminished the already disappearing grassland which contained plants that were a key dietary component of herbivorous megafauna. While there is no scientific consensus on this, it is plausible that homo sapiens and natural climate change had a combined impact. Overall, there is a great deal of evidence for humans being the culprit, but by ruling out climate change completely as a cause of the Australian megafauna extinction we are not getting the whole picture. The climate change in Australia 45,000 years ago destabilized the ecosystem, making it particularly vulnerable to hunting and fire agriculture by humans; this is probably what led to the extinction of the Australian megafauna.

Several studies provide evidence that climate change caused megafaunal extinction during the Pleistocene in Australia. One group of researchers analyzed fossilized teeth found at Cuddie Springs in southeastern Australia. By analyzing oxygen isotopes, they measured aridity, and by analyzing carbon isotopes and dental microwear texture analysis, they assessed megafaunal diets and vegetation. During the middle Pleistocene, southeastern Australia was dominated by browsers, including fauna that consumed C4 plants. By the late Pleistocene, the C4 plant dietary component had decreased considerably. This shift may have been caused by increasingly arid conditions, which may have caused dietary restrictions. Other isotopic analyses of eggshells and wombat teeth also point to a decline of C4 vegetation after 45 Ka. This decline in C4 vegetation is coincident with increasing aridity. Increasingly arid conditions in southeastern Australia during the late Pleistocene may have stressed megafauna, and contributed to their decline. [122]

Procoptodon goliath reconstruction Procoptodon goliah NT.png
Procoptodon goliath reconstruction
The American flamingo (Phoenicopterus ruber) was one of four species of flamingo present in Australia in the Quaternary, all of which are now either extinct or extirpated. Australia is now the only inhabited continent in the world without flamingoes. American Flamingo - Flamenco (Phoenicopterus ruber) (11854683776).jpg
The American flamingo (Phoenicopterus ruber) was one of four species of flamingo present in Australia in the Quaternary, all of which are now either extinct or extirpated. Australia is now the only inhabited continent in the world without flamingoes.
Megalania skeleton, Melbourne Museum Varanus priscus Melbourne Museum.jpg
Megalania skeleton, Melbourne Museum

In Sahul (a former continent composed of Australia and New Guinea), the sudden and extensive spate of extinctions occurred earlier than in the rest of the world. [123] [124] [125] [126] [127] Most evidence points to a 20,000 year period after human arrival circa 63,000 BCE, [128] but scientific argument continues as to the exact date range. [129] In the rest of the Pacific (other Australasian islands such as New Caledonia, and Oceania) although in some respects far later, endemic fauna also usually perished quickly upon the arrival of humans in the late Pleistocene and early Holocene.

Relationship to later extinctions

There is no general agreement on where the Quaternary extinction event ends, and the Holocene, or anthropogenic, extinction begins, or if they should be considered separate events at all. [136] [137] Some have suggested that anthropogenic extinctions may have begun as early as when the first modern humans spread out of Africa between 100,000 and 200,000 years ago, which is supported by rapid megafaunal extinction following recent human colonisation in Australia, New Zealand and Madagascar, [138] in a similar way that any large, adaptable predator moving into a new ecosystem would. In many cases, it is suggested even minimal hunting pressure was enough to wipe out large fauna, particularly on geographically isolated islands. [139] [140] Only during the most recent parts of the extinction have plants also suffered large losses. [141]

Overall, the Holocene extinction can be characterised by the human impact on the environment. The Holocene extinction continues into the 21st century, with overfishing, ocean acidification and the amphibian crisis being a few broader examples of an almost universal, cosmopolitan decline of biodiversity.

Causes

Human activity

Hunting

The hunting hypothesis suggests that humans hunted megaherbivores to extinction, which in turn caused the extinction of carnivores and scavengers which had preyed upon those animals. [142] [143] [144] This hypothesis holds Pleistocene humans responsible for the megafaunal extinction. One variant, known as blitzkrieg, portrays this process as relatively quick. Some of the direct evidence for this includes: fossils of some megafauna found in conjunction with human remains, embedded arrows and tool cut marks found in megafaunal bones, and European cave paintings that depict such hunting. Biogeographical evidence is also suggestive: the areas of the world where humans evolved currently have more of their Pleistocene megafaunal diversity (the elephants and rhinos of Asia and Africa) compared to other areas such as Australia, the Americas, Madagascar and New Zealand without the earliest humans. The overkill hypothesis, a variant of the hunting hypothesis, was proposed in 1966 by Paul S. Martin, [145] Professor of Geosciences Emeritus at the Desert Laboratory of the University of Arizona. [146]

Despeciation within the genus Homo. Homo-Stammbaum, Version Stringer-en.svg
Despeciation within the genus Homo .
Known H. sapiens migration routes in the Pleistocene. Homo sapiens dispersal routes.jpg
Known H. sapiens migration routes in the Pleistocene.

Circumstantially, the close correlation in time between the appearance of humans in an area and extinction there provides weight for this scenario. [2] The megafaunal extinctions covered a vast period of time and highly variable climatic situations. The earliest extinctions in Australia were complete approximately 50,000 BP, well before the last glacial maximum and before rises in temperature. The most recent extinction in New Zealand was complete no earlier than 500 BP and during a period of cooling. In between these extremes megafaunal extinctions have occurred progressively in such places as North America, South America and Madagascar with no climatic commonality. The only common factor that can be ascertained is the arrival of humans. [147] [148] This phenomenon appears even within regions. The mammal extinction wave in Australia about 50,000 years ago coincides not with known climatic changes, but with the arrival of humans. In addition, large mammal species like the giant kangaroo Protemnodon appear to have succumbed sooner on the Australian mainland than on Tasmania, which was colonised by humans a few thousand years later. [149] [150] A study published in 2015 supported the hypothesis further by running several thousand scenarios that correlated the time windows in which each species is known to have become extinct with the arrival of humans on different continents or islands. This was compared against climate reconstructions for the last 90,000 years. The researchers found correlations of human spread and species extinction indicating that the human impact was the main cause of the extinction, while climate change exacerbated the frequency of extinctions. The study, however, found an apparently low extinction rate in the fossil record of mainland Asia. [151] [152] A 2020 study published in Science Advances found that human population size and/or specific human activities, not climate change, caused rapidly rising global mammal extinction rates during the past 126,000 years. Around 96% of all mammalian extinctions over this time period are attributable to human impacts. According to Tobias Andermann, lead author of the study, "these extinctions did not happen continuously and at constant pace. Instead, bursts of extinctions are detected across different continents at times when humans first reached them. More recently, the magnitude of human driven extinctions has picked up the pace again, this time on a global scale." [153] [154] On a related note, the population declines of still extant megafauna during the Pleistocene have also been shown to correlate with human expansion rather than climate change. [15] The extinction's extreme bias towards larger animals further supports a relationship with human activity rather than climate change. [155]

Extinction through human hunting has been supported by archaeological finds of mammoths with projectile points embedded in their skeletons, by observations of modern naive animals allowing hunters to approach easily [156] [157] [158] and by computer models by Mosimann and Martin, [159] and Whittington and Dyke, [160] and most recently by Alroy. [161]

The timing of extinctions follows the "March of Man" Extinctions Africa Austrailia NAmerica Madagascar.gif
The timing of extinctions follows the "March of Man"

Major objections have been raised regarding the hunting hypothesis. Notable among them is the sparsity of evidence of human hunting of megafauna. [162] There is no archeological evidence that in North America megafauna other than mammoths, mastodons, gomphotheres and bison were hunted, despite the fact that, for example, camels and horses are very frequently reported in fossil history. [163] Overkill proponents, however, say this is due to the fast extinction process in North America and the low probability of animals with signs of butchery to be preserved. [164] A study by Surovell and Grund concluded "archaeological sites dating to the time of the coexistence of humans and extinct fauna are rare. Those that preserve bone are considerably more rare, and of those, only a very few show unambiguous evidence of human hunting of any type of prey whatsoever." [165] Eugene S. Hunn points out that the birthrate in hunter-gatherer societies is generally too low, that too much effort is involved in the bringing down of a large animal by a hunting party, and that in order for hunter-gatherers to have brought about the extinction of megafauna simply by hunting them to death, an extraordinary amount of meat would have had to have been wasted. [166]

Second-order predation

Combination Hypotheses: Climate Change, Overkill + Climate Change, Second-Order Predation + Climate Change CombinedHypothesisPleistoceneExtinction.gif
Combination Hypotheses: Climate Change, Overkill + Climate Change, Second-Order Predation + Climate Change
Overkill Hypothesis and Second-Order Predation SingleHypothesisPleistoceneExtinctions.gif
Overkill Hypothesis and Second-Order Predation

The Second-Order Predation Hypothesis says that as humans entered the New World they continued their policy of killing predators, which had been successful in the Old World but because they were more efficient and because the fauna, both herbivores and carnivores, were more naive, they killed off enough carnivores to upset the ecological balance of the continent, causing overpopulation, environmental exhaustion, and environmental collapse. The hypothesis accounts for changes in animal, plant, and human populations.

The scenario is as follows:

  • After the arrival of H. sapiens in the New World, existing predators must share the prey populations with this new predator. Because of this competition, populations of original, or first-order, predators cannot find enough food; they are in direct competition with humans.
  • Second-order predation begins as humans begin to kill predators.
  • Prey populations are no longer well controlled by predation. Killing of nonhuman predators by H. sapiens reduces their numbers to a point where these predators no longer regulate the size of the prey populations.
  • Lack of regulation by first-order predators triggers boom-and-bust cycles in prey populations. Prey populations expand and consequently overgraze and over-browse the land. Soon the environment is no longer able to support them. As a result, many herbivores starve. Species that rely on the slowest recruiting food become extinct, followed by species that cannot extract the maximum benefit from every bit of their food.
  • Boom-bust cycles in herbivore populations change the nature of the vegetative environment, with consequent climatic impacts on relative humidity and continentality. Through overgrazing and overbrowsing, mixed parkland becomes grassland, and climatic continentality increases.

The second-order predation hypothesis has been supported by a computer model, the Pleistocene extinction model (PEM), which, using the same assumptions and values for all variables (herbivore population, herbivore recruitment rates, food needed per human, herbivore hunting rates, etc.) other than those for hunting of predators. It compares the overkill hypothesis (predator hunting = 0) with second-order predation (predator hunting varied between 0.01 and 0.05 for different runs). The findings are that second-order predation is more consistent with extinction than is overkill [167] [168] (results graph at left). The Pleistocene extinction model is the only test of multiple hypotheses and is the only model to specifically test combination hypotheses by artificially introducing sufficient climate change to cause extinction. When overkill and climate change are combined they balance each other out. Climate change reduces the number of plants, overkill removes animals, therefore fewer plants are eaten. Second-order predation combined with climate change exacerbates the effect of climate change. [169] (results graph at right). The second-order predation hypothesis is further supported by the observation above that there was a massive increase in bison populations. [170]

However, this hypothesis has been criticised on the grounds that the multispecies model produces a mass extinction through indirect competition between herbivore species: small species with high reproductive rates subsidize predation on large species with low reproductive rates. [161] All prey species are lumped in the Pleistocene extinction model. Also, the control of population sizes by predators is not fully supported by observations of modern ecosystems. [171] The hypothesis further assumes decreases in vegetation due to climate change, but deglaciation doubled the habitable area of North America. Any vegetational changes that did occur failed to cause almost any extinctions of small vertebrates, and they are more narrowly distributed on average, which detractors cite as evidence against the hypothesis.

Climate change

At the end of the 19th and beginning of the 20th centuries, when scientists first realized that there had been glacial and interglacial ages, and that they were somehow associated with the prevalence or disappearance of certain animals, they surmised that the termination of the Pleistocene ice age might be an explanation for the extinctions.

The most obvious change associated with the termination of an ice age is the increase in temperature. Between 15,000 BP and 10,000 BP, a 6 °C increase in global mean annual temperatures occurred. This was generally thought to be the cause of the extinctions. According to this hypothesis, a temperature increase sufficient to melt the Wisconsin ice sheet could have placed enough thermal stress on cold-adapted mammals to cause them to die. Their heavy fur, which helps conserve body heat in the glacial cold, might have prevented the dumping of excess heat, causing the mammals to die of heat exhaustion. Large mammals, with their reduced surface area-to-volume ratio, would have fared worse than small mammals. A study covering the past 56,000 years indicates that rapid warming events with temperature changes of up to 16 °C (29 °F) had an important impact on the extinction of megafauna. Ancient DNA and radiocarbon data indicates that local genetic populations were replaced by others within the same species or by others within the same genus. Survival of populations was dependent on the existence of refugia and long distance dispersals, which may have been disrupted by human hunters. [172]

Other scientists have proposed that increasingly extreme weather—hotter summers and colder winters—referred to as "continentality", or related changes in rainfall caused the extinctions. It has been shown that vegetation changed from mixed woodland-parkland to separate prairie and woodland. [173] [174] [175] This may have affected the kinds of food available. Shorter growing seasons may have caused the extinction of large herbivores and the dwarfing of many others. In this case, as observed, bison and other large ruminants would have fared better than horses, elephants and other monogastrics, because ruminants are able to extract more nutrition from limited quantities of high-fiber food and better able to deal with anti-herbivory toxins. [176] [177] [178] So, in general, when vegetation becomes more specialized, herbivores with less diet flexibility may be less able to find the mix of vegetation they need to sustain life and reproduce, within a given area. Increased continentality resulted in reduced and less predictable rainfall limiting the availability of plants necessary for energy and nutrition. [179] [180] [181] It has been suggested that this change in rainfall restricted the amount of time favorable for reproduction. [182] [183] This could disproportionately harm large animals, since they have longer, more inflexible mating periods, and so may have produced young at unfavorable seasons (i.e., when sufficient food, water, or shelter was unavailable because of shifts in the growing season). In contrast, small mammals, with their shorter life cycles, shorter reproductive cycles, and shorter gestation periods, could have adjusted to the increased unpredictability of the climate, both as individuals and as species which allowed them to synchronize their reproductive efforts with conditions favorable for offspring survival. If so, smaller mammals would have lost fewer offspring and would have been better able to repeat the reproductive effort when circumstances once more favored offspring survival. [184] A study looking at the environmental conditions across Europe, Siberia and the Americas from 25,000 to 10,000 YBP found that prolonged warming events leading to deglaciation and maximum rainfall occurred just prior to the transformation of the rangelands that supported megaherbivores into widespread wetlands that supported herbivore-resistant plants. The study proposes that moisture-driven environmental change led to the megafaunal extinctions and that Africa's trans-equatorial position allowed rangeland to continue to exist between the deserts and the central forests, therefore fewer megafauna species became extinct there. [172]

Evidence in Southeast Asia, in contrast to Europe, Australia, and the Americas, suggests that climate change and an increasing sea level were significant factors in the extinction of several herbivorous species. Alterations in vegetation growth and new access routes for early humans and mammals to previously isolated, localized ecosystems were detrimental to select groups of fauna. [185]

In Beringia, megafauna may have gone extinct because of particularly intense paludification and because the land connection between Eurasia and North America flooded before the Cordilleran Ice Sheet retreated far enough to reopen the corridor between Beringia and the remainder of North America. [186] Woolly mammoths became extirpated from Beringia because of climatic factors, although human activity also played a synergistic role in their decline. [187] In North America, a Radiocarbon-dated Event-Count (REC) modelling study found that megafaunal declines in North America correlated with climatic changes instead of human population expansion. [188]

Critics object that since there were multiple glacial advances and withdrawals in the evolutionary history of many of the megafauna, it is rather implausible that only after the last glacial maximum would there be such extinctions. One study suggests that the Pleistocene megafaunal composition may have differed markedly from that of earlier interglacials, making the Pleistocene populations particularly vulnerable to changes in their environment. [189]

Studies propose that the annual mean temperature of the current interglacial that we have seen for the last 10,000 years is no higher than that of previous interglacials, yet most of the same large mammals survived similar temperature increases. [190] [191] [192] In addition, numerous species such as mammoths on Wrangel Island and St. Paul Island survived in human-free refugia despite changes in climate. [193] This would not be expected if climate change were responsible (unless their maritime climates offered some protection against climate change not afforded to coastal populations on the mainland). Under normal ecological assumptions island populations should be more vulnerable to extinction due to climate change because of small populations and an inability to migrate to more favorable climes.[ citation needed ]

Critics have also identified a number of problems with the continentality hypotheses. Megaherbivores have prospered at other times of continental climate. For example, megaherbivores thrived in Pleistocene Siberia, which had and has a more continental climate than Pleistocene or modern (post-Pleistocene, interglacial) North America. [194] [195] [196] The animals that became extinct actually should have prospered during the shift from mixed woodland-parkland to prairie, because their primary food source, grass, was increasing rather than decreasing. [197] [196] [198] Although the vegetation did become more spatially specialized, the amount of prairie and grass available increased, which would have been good for horses and for mammoths, and yet they became extinct. This criticism ignores the increased abundance and broad geographic extent of Pleistocene bison at the end of the Pleistocene, which would have increased competition for these resources in a manner not seen in any earlier interglacials. [189] Although horses became extinct in the New World, they were successfully reintroduced by the Spanish in the 16th century—into a modern post-Pleistocene, interglacial climate. Today there are feral horses still living in those same environments. They find a sufficient mix of food to avoid toxins, they extract enough nutrition from forage to reproduce effectively and the timing of their gestation is not an issue. Of course, this criticism ignores the obvious fact that present-day horses are not competing for resources with ground sloths, mammoths, mastodons, camels, llamas, and bison. Similarly, mammoths survived the Pleistocene Holocene transition on isolated, uninhabited islands in the Mediterranean Sea until 4,000 to 7,000 years ago, [199] as well as on Wrangel Island in the Siberian Arctic. [200] Additionally, large mammals should have been able to migrate, permanently or seasonally, if they found the temperature too extreme, the breeding season too short, or the rainfall too sparse or unpredictable. [201] Seasons vary geographically. By migrating away from the equator, herbivores could have found areas with growing seasons more favorable for finding food and breeding successfully. Modern-day African elephants migrate during periods of drought to places where there is apt to be water. [202] Large animals also store more fat in their bodies than do medium-sized animals and this should have allowed them to compensate for extreme seasonal fluctuations in food availability. [203]

Some evidence weighs against climate change as a valid hypothesis as applied to Australia. It has been shown that the prevailing climate at the time of extinction (40,000–50,000 BP) was similar to that of today, and that the extinct animals were strongly adapted to an arid climate. The evidence indicates that all of the extinctions took place in the same short time period, which was the time when humans entered the landscape. The main mechanism for extinction was probably fire (started by humans) in a then much less fire-adapted landscape. Isotopic evidence shows sudden changes in the diet of surviving species, which could correspond to the stress they experienced before extinction. [204] [205] [206]

Some evidence obtained from analysis of the tusks of mastodons from the American Great Lakes region appears inconsistent with the climate change hypothesis. Over a span of several thousand years prior to their extinction in the area, the mastodons show a trend of declining age at maturation. This is the opposite of what one would expect if they were experiencing stresses from deteriorating environmental conditions, but is consistent with a reduction in intraspecific competition that would result from a population being reduced by human hunting. [207]

It may be observed that neither the overkill nor the climate change hypotheses can fully explain events: browsers, mixed feeders and non-ruminant grazer species suffered most, while relatively more ruminant grazers survived. [208] However, a broader variation of the overkill hypothesis may predict this, because changes in vegetation wrought by either Second Order Predation (see below) [169] [209] or anthropogenic fire preferentially selects against browse species.[ citation needed ]

Disease

The hyperdisease hypothesis, as advanced by Ross D. E. MacFee and Preston A. Marx, attributes the extinction of large mammals during the late Pleistocene to indirect effects of the newly arrived aboriginal humans. [210] [211] [212] In more recent times, disease has driven many vulnerable species to extinction; the introduction of avian malaria and avipoxvirus, for example, has greatly decreased the populations of the endemic birds of Hawaii, with some going extinct. [213] The hyperdisease hypothesis proposes that humans or animals traveling with them (e.g., chickens or domestic dogs) introduced one or more highly virulent diseases into vulnerable populations of native mammals, eventually causing extinctions. The extinction was biased toward larger-sized species because smaller species have greater resilience because of their life history traits (e.g., shorter gestation time, greater population sizes, etc.). Humans are thought to be the cause because other earlier immigrations of mammals into North America from Eurasia did not cause extinctions. [210] A similar suggestion is that pathogens were transmitted by the expanding humans via the domesticated dogs they brought with them. [214] A related theory proposes that a highly contagious prion disease similar to chronic wasting disease or scrapie that was capable of infecting a large number of species was the culprit. Animals weakened by this "superprion" would also have easily become reservoirs of viral and bacterial diseases as they succumbed to neurological degeneration from the prion, causing a cascade of different diseases to spread among various mammal species. This theory could potentially explain the prevalence of heterozygosity at codon 129 of the prion protein gene in humans, which has been speculated to be the result of natural selection against homozygous genotypes that were more susceptible to prion disease and thus potentially a tell-tale of a major prion pandemic that affected humans of or younger than reproductive age far in the past and disproportionately killed before they could reproduce those with homozygous genotypes at codon 129. [215]

If a disease was indeed responsible for the end-Pleistocene extinctions, then there are several criteria it must satisfy (see Table 7.3 in MacPhee & Marx 1997). First, the pathogen must have a stable carrier state in a reservoir species. That is, it must be able to sustain itself in the environment when there are no susceptible hosts available to infect. Second, the pathogen must have a high infection rate, such that it is able to infect virtually all individuals of all ages and sexes encountered. Third, it must be extremely lethal, with a mortality rate of c. 50–75%. Finally, it must have the ability to infect multiple host species without posing a serious threat to humans. Humans may be infected, but the disease must not be highly lethal or able to cause an epidemic.[ citation needed ]

As with other hypotheses, a number of counterarguments to the hyperdisease hypothesis have been put forth. Generally speaking, disease has to be very virulent to kill off all the individuals in a genus or species. Even such a virulent disease as West Nile fever is unlikely to have caused extinction. [216] The disease would need to be implausibly selective while being simultaneously implausibly broad. Such a disease needs to be capable of killing off wolves such as Canis dirus or goats such as Oreamnos harringtoni while leaving other very similar species (Canis lupus and Oreamnos americanus, respectively) unaffected. It would need to be capable of killing off flightless birds while leaving closely related flighted species unaffected. Yet while remaining sufficiently selective to afflict only individual species within genera it must be capable of fatally infecting across such clades as birds, marsupials, placentals, testudines, and crocodilians. No disease with such a broad scope of fatal infectivity is known, much less one that remains simultaneously incapable of infecting numerous closely related species within those disparate clades. On the other hand, this objection does not account for the possibility of a variety of different diseases being introduced around the same era.[ citation needed ] Numerous species including wolves, mammoths, camelids, and horses had emigrated continually between Asia and North America over the past 100,000 years. For the disease hypothesis to be applicable there it would require that the population remain immunologically naive despite this constant transmission of genetic and pathogenic material.[ citation needed ] The dog-specific hypothesis in particular cannot account for several major extinction events, notably the Americas (for reasons already covered) and Australia. Dogs did not arrive in Australia until approximately 35,000 years after the first humans arrived there, and approximately 30,000 years after the Australian megafaunal extinction was complete.[ citation needed ]

Effects

The extinction of the megafauna could have caused the disappearance of the mammoth steppe rather than the other way around. Alaska now has low nutrient soil unable to support bison, mammoths, and horses. R. Dale Guthrie has claimed this as a cause of the extinction of the megafauna there; however, he may be interpreting it backwards. The loss of large herbivores to break up the permafrost allows the cold soils that are unable to support large herbivores today. Today, in the arctic, where trucks have broken the permafrost grasses and diverse flora and fauna can be supported. [217] [218] In addition, Chapin (Chapin 1980) showed that simply adding fertilizer to the soil in Alaska could make grasses grow again like they did in the era of the mammoth steppe. Possibly, the extinction of the megafauna and the corresponding loss of dung is what led to low nutrient levels in modern-day soil and therefore is why the landscape can no longer support megafauna.

Megafauna play a significant role in the lateral transport of mineral nutrients in an ecosystem, tending to translocate them from areas of high to those of lower abundance. They do so by their movement between the time they consume the nutrient and the time they release it through elimination (or, to a much lesser extent, through decomposition after death). [219] In South America's Amazon Basin, it is estimated that such lateral diffusion was reduced over 98% following the megafaunal extinctions that occurred roughly 12,500 years ago. [220] [221] Given that phosphorus availability is thought to limit productivity in much of the region, the decrease in its transport from the western part of the basin and from floodplains (both of which derive their supply from the uplift of the Andes) to other areas is thought to have significantly impacted the region's ecology, and the effects may not yet have reached their limits. [221] The extinction of the mammoths allowed grasslands they had maintained through grazing habits to become birch forests. [222] The new forest and the resulting forest fires may have induced climate change. [222] Such disappearances might be the result of the proliferation of modern humans. [223] [224]

Large populations of megaherbivores have the potential to contribute greatly to the atmospheric concentration of methane, which is an important greenhouse gas. Modern ruminant herbivores produce methane as a byproduct of foregut fermentation in digestion, and release it through belching or flatulence. Today, around 20% of annual methane emissions come from livestock methane release. In the Mesozoic, it has been estimated that sauropods could have emitted 520 million tons of methane to the atmosphere annually, [225] contributing to the warmer climate of the time (up to 10 °C warmer than at present). [225] [226] This large emission follows from the enormous estimated biomass of sauropods, and because methane production of individual herbivores is believed to be almost proportional to their mass. [225]

Recent studies have indicated that the extinction of megafaunal herbivores may have caused a reduction in atmospheric methane. One study examined the methane emissions from the bison that occupied the Great Plains of North America before contact with European settlers. The study estimated that the removal of the bison caused a decrease of as much as 2.2 million tons per year. [227] Another study examined the change in the methane concentration in the atmosphere at the end of the Pleistocene epoch after the extinction of megafauna in the Americas. After early humans migrated to the Americas about 13,000 BP, their hunting and other associated ecological impacts led to the extinction of many megafaunal species there. Calculations suggest that this extinction decreased methane production by about 9.6 million tons per year. This suggests that the absence of megafaunal methane emissions may have contributed to the abrupt climatic cooling at the onset of the Younger Dryas. [228] The decrease in atmospheric methane that occurred at that time, as recorded in ice cores, was 2–4 times more rapid than any other decrease in the last half million years, suggesting that an unusual mechanism was at work. [228]

See also

Related Research Articles

<i>Diprotodon</i> Extinct marsupial genus

Diprotodon is an extinct genus of marsupial from the Pleistocene of Australia containing one species, D. optatum. The earliest finds date to 1.77 million to 780,000 years ago but most specimens are dated to after 110,000 years ago. Its remains were first unearthed in 1830 in Wellington Caves, New South Wales, and contemporaneous paleontologists guessed they belonged to rhinos, elephants, hippos or dugongs. Diprotodon was formally described by English naturalist Richard Owen in 1838, and was the first named Australian fossil mammal, and led Owen to become the foremost authority of his time on other marsupials and Australian megafauna, which were enigmatic to European science.

<span class="mw-page-title-main">Holocene extinction</span> Ongoing extinction event caused by human activity

The Holocene extinction, or Anthropocene extinction, is the ongoing extinction event caused by humans during the Holocene epoch. These extinctions span numerous families of plants and animals, including mammals, birds, reptiles, amphibians, fish, and invertebrates, and affecting not just terrestrial species but also large sectors of marine life. With widespread degradation of biodiversity hotspots, such as coral reefs and rainforests, as well as other areas, the vast majority of these extinctions are thought to be undocumented, as the species are undiscovered at the time of their extinction, which goes unrecorded. The current rate of extinction of species is estimated at 100 to 1,000 times higher than natural background extinction rates and is increasing. During the past 100–200 years, biodiversity loss and species extinction have accelerated, to the point that most conservation biologists now believe that human activity has either produced a period of mass extinction, or is on the cusp of doing so. As such, after the "Big Five" mass extinctions, the Holocene extinction event has also been referred to as the sixth mass extinction or sixth extinction; given the recent recognition of the Capitanian mass extinction, the term seventh mass extinction has also been proposed for the Holocene extinction event.

<span class="mw-page-title-main">Mammoth</span> Extinct genus of mammals

A mammoth is any species of the extinct elephantid genus Mammuthus. The various species of mammoth were commonly equipped with long, curved tusks. They lived from the Pliocene epoch into the Holocene about 4,000 years ago, and various species existed in Africa, Europe, Asia, and North America. Mammoths are more closely related to living Asian elephants than African elephants.

<i>Megatherium</i> Genus of ground sloth (extinct)

Megatherium is an extinct genus of ground sloths endemic to South America that lived from the Early Pliocene through the end of the Pleistocene. It is best known for the elephant-sized type species M. americanum, sometimes known as the giant ground sloth, or the megathere, native to the Pampas through southern Bolivia during the Pleistocene. Various other smaller species belonging to the subgenus Pseudomegatherium are known from the Andes.

<span class="mw-page-title-main">Megafauna</span> Large animals

In zoology, megafauna are large animals. The most common thresholds to be a megafauna are weighing over 46 kilograms (100 lb) or weighing over a tonne, 1,000 kilograms (2,205 lb). The first of these include many species not popularly thought of as overly large, and being the only few large animals left in a given range/area, such as white-tailed deer, Thomson's gazelle, and red kangaroo.

<span class="mw-page-title-main">Mastodon</span> Genus of mammals (fossil)

A mastodon is any proboscidean belonging to the extinct genus Mammut. Mastodons inhabited North and Central America from the late Miocene up to their extinction at the end of the Pleistocene 10,000 to 11,000 years ago. Mastodons are the most recent members of the family Mammutidae, which diverged from the ancestors of elephants at least 25 million years ago. M. americanum, the American mastodon, is the youngest and best-known species of the genus. They lived in herds and were predominantly forest-dwelling animals. M. americanum is inferred to have had a browsing diet with a preference for woody material, distinct from that of the contemporary Columbian mammoth. Mastodons became extinct as part of the Quaternary extinction event that exterminated most Pleistocene megafauna present in the Americas, believed to have been caused by a combination of climate changes at the end of the Pleistocene and hunting by recently arrived Paleo-Indians, as evidenced by a number of kill sites where mastodon remains are associated with human artifacts.

<i>Genyornis</i> Extinct genus of birds

Genyornis newtoni, also known as the Newton's mihirung, Newton's thunder bird or mihirung paringmal, is an extinct species of large, flightless bird that lived in Australia during the Pleistocene Epoch until around 50,000 years ago. Over two metres in height, they were likely herbivorous. Many other species of Australian megafauna became extinct in Australia around that time, coinciding with the arrival of humans. It is the last known member of the extinct flightless bird family Dromornithidae which had been part of the fauna of the Australian continent for over 30 million years. They are not closely related to ratites such as emus, and their closest living relatives are thought to be fowl.

<span class="mw-page-title-main">Clovis culture</span> Prehistoric culture in the Americas c. 11, 500 to 10,800 BCE

Clovis culture is a prehistoric Paleoamerican archaeological culture, named for distinct stone and bone tools found in close association with Pleistocene fauna, particularly two Columbian mammoths, at Blackwater Locality No. 1 near Clovis, New Mexico, in 1936 and 1937, though Paleoindian artifacts had been found at the site since the 1920s. It existed from roughly 11,500 to 10,800 BCE near the end of the Last Glacial Period.

<span class="mw-page-title-main">Steppe bison</span> Extinct species of mammal

The steppe bison or steppe wisent is an extinct species of bison that was once found on the mammoth steppe where its range included British Isles, Europe, Central Asia, Northern to Northeastern Asia including the Japanese archipelago, Beringia, and central North America, from northwest Canada to Mexico during the Quaternary. This wide distribution is sometimes called the Pleistocene bison belt, compared to the Great bison belt. It is ancestral to modern American bison. Three chronological subspecies, Bison priscus priscus, Bison priscus mediator, and Bison priscus gigas, have been suggested.

<span class="mw-page-title-main">Australian megafauna</span> Large animals in Australia, past and present era

The term Australian megafauna refers to the megafauna in Australia during the Pleistocene Epoch. Most of these species became extinct during the latter half of the Pleistocene, and the roles of human and climatic factors in their extinction are contested.

<i>Castoroides</i> Extinct genus of beaver

Castoroides, or giant beaver, is an extinct genus of enormous, bear-sized beavers that lived in North America during the Pleistocene. Two species are currently recognized, C. dilophidus in the Southeastern United States and C. ohioensis in most of North America. C. leiseyorum was previously described from the Irvingtonian age but is now regarded as an invalid name. All specimens previously described as C. leiseyorum are considered to belong to C. dilophidus.

<span class="mw-page-title-main">Columbian mammoth</span> Extinct species of mammoth that inhabited North America

The Columbian mammoth is an extinct species of mammoth that inhabited the Americas as far north as the Northern United States and as far south as Costa Rica during the Pleistocene epoch. The Columbian mammoth descended from Eurasian mammoths that colonised North America during the Early Pleistocene around 1.5-1.3 million years ago, and later experienced hybridisation with the woolly mammoth lineage. The Columbian mammoth was among the last mammoth species, and the pygmy mammoths evolved from them on the Channel Islands of California. The closest extant relative of the Columbian and other mammoths is the Asian elephant.

<span class="mw-page-title-main">Mammoth steppe</span> Prehistoric biome

During the Last Glacial Maximum, the mammoth steppe, also known as steppe-tundra, was once the Earth's most extensive biome. It stretched east-to-west, from the Iberian Peninsula in the west of Europe, across Eurasia to North America, through Beringia and Canada; from north-to-south, the steppe reached from the arctic islands southward to China. The mammoth steppe was cold and dry, and relatively featureless, though topography and geography varied considerably throughout. Some areas featured rivers which, through erosion, naturally created gorges, gulleys, or small glens. The continual glacial recession and advancement over millennia contributed more to the formation of larger valleys and different geographical features. Overall, however, the steppe is known to be flat and expansive grassland. The vegetation was dominated by palatable, high-productivity grasses, herbs and willow shrubs.

<i>Simosthenurus</i> Extinct genus of marsupials

Simosthenurus, also referred to as the short-faced kangaroo, is an extinct genus of megafaunal macropods that existed in Australia, specifically Tasmania, during the Pleistocene. Analysis of Simosthenurus fossils has contributed to the finding that there are three lineages of macropods: Sthenurinae, Macropodinae, and Lagostrophinae. The genus Simosthenurus was among the sthenurines.

<span class="mw-page-title-main">Pleistocene rewilding</span> Ecological practice

Pleistocene rewilding is the advocacy of the reintroduction of extant Pleistocene megafauna, or the close ecological equivalents of extinct megafauna. It is an extension of the conservation practice of rewilding, which aims to restore functioning, self-sustaining ecosystems through practices that may include species reintroductions.

<i>Arctodus</i> Extinct genus of bears

Arctodus is an extinct genus of short-faced bear that inhabited North America during the Pleistocene. There are two recognized species: the lesser short-faced bear and the giant short-faced bear. Both species are relatively rare in the fossil record. A. pristinus was largely restricted to the Early Pleistocene of the eastern United States, whereas A. simus had a broader range, with most finds being from the Late Pleistocene of the United States, Mexico and Canada. A. simus evolved from A. pristinus, but both species likely overlapped in the Middle Pleistocene. Of these species, A. simus was larger, is known from more complete remains, and is considered one of the most charismatic of North America's megafauna.

<span class="mw-page-title-main">Paul Schultz Martin</span> American paleontologist

Paul Schultz Martin was an American geoscientist at the University of Arizona who developed the theory that the Pleistocene extinction of large mammals worldwide was caused by overhunting by humans. Martin's work bridged the fields of ecology, anthropology, geosciences, and paleontology.

<span class="mw-page-title-main">Evolutionary anachronism</span> Attributes of living species that arose due to coevolution with other now-extinct species

Evolutionary anachronism, also known as "ecological anachronism", is a term initially referring to attributes of native plant species that seemed best explained as having been favorably selected in the past due to their coevolution with plant-eating megafauna that are now extinct. Diminished effectiveness and distance of seed dispersal by fruit-eating mammals inhabiting the same ecosystems today suggest maladaptation. Maladaptation of such fruiting plants will intensify as ongoing climate change shifts the physical and ecological conditions within their current geographic range.

<span class="mw-page-title-main">Beringian wolf</span> Extinct type of wolf that lived during the Ice Age in Alaska, Yukon, and northern British Columbia

The Beringian wolf is an extinct population of wolf that lived during the Ice Age. It inhabited what is now modern-day Alaska, Yukon, and northern British Columbia. Some of these wolves survived well into the Holocene. The Beringian wolf is an ecomorph of the gray wolf and has been comprehensively studied using a range of scientific techniques, yielding new information on their prey species and feeding behaviors. It has been determined that these wolves are morphologically distinct from modern North American wolves and genetically basal to most modern and extinct wolves. The Beringian wolf has not been assigned a subspecies classification and its relationship with the extinct European cave wolf is not clear.

<span class="mw-page-title-main">Wood-pasture hypothesis</span> Ecological theory

The wood-pasture hypothesis is a scientific hypothesis positing that open and semi-open pastures and wood-pastures formed the predominant type of landscape in post-glacial temperate Europe, rather than the common belief of primeval forests. The hypothesis proposes that such a landscape would be formed and maintained by large wild herbivores. Although others, including landscape ecologist Oliver Rackham, had previously expressed similar ideas, it was the Dutch researcher Frans Vera, who, in his 2000 book Grazing Ecology and Forest History, first developed a comprehensive framework for such ideas and formulated them into a theorem. Vera's proposals, although highly controversial, came at a time when the role grazers played in woodlands was increasingly being reconsidered, and are credited for ushering in a period of increased reassesment and interdisciplinary research in European conservation theory and practice. Although Vera largely focused his research on the European situation, his findings could also be applied to other temperate ecological regions worldwide, especially the broadleaved ones.

References

  1. 1 2 3 Stuart, Anthony John (1999), MacPhee, Ross D. E. (ed.), "Late Pleistocene Megafaunal Extinctions: A European Perspective", Extinctions in Near Time, Boston, MA: Springer US, pp. 257–269, doi:10.1007/978-1-4757-5202-1_11, ISBN   978-1-4419-3315-7 , retrieved 2023-05-28
  2. 1 2 3 4 Sandom, Christopher; Faurby, Søren; Sandel, Brody; Svenning, Jens-Christian (2014-07-22). "Global late Quaternary megafauna extinctions linked to humans, not climate change". Proceedings of the Royal Society B: Biological Sciences. 281 (1787): 20133254. doi:10.1098/rspb.2013.3254. ISSN   0962-8452. PMC   4071532 . PMID   24898370.
  3. 1 2 3 Faith, J. Tyler; Surovell, Todd A. (2009-12-08). "Synchronous extinction of North America's Pleistocene mammals". Proceedings of the National Academy of Sciences. 106 (49): 20641–20645. Bibcode:2009PNAS..10620641F. doi: 10.1073/pnas.0908153106 . ISSN   0027-8424. PMC   2791611 . PMID   19934040.
  4. Bird, Michael I.; Hutley, Lindsay B.; Lawes, Michael J.; Lloyd, Jon; Luly, Jon G.; Ridd, Peter V.; Roberts, Richard G.; Ulm, Sean; Wurster, Christopher M. (July 2013). "Humans, megafauna and environmental change in tropical Australia". Journal of Quaternary Science. 28 (5): 439–452. doi:10.1002/jqs.2639. S2CID   129936890.
  5. Dembitzer, Jacob; Barkai, Ran; Ben-Dor, Miki; Meiri, Shai (2022-01-15). "Levantine overkill: 1.5 million years of hunting down the body size distribution". Quaternary Science Reviews . 276: 107316. Bibcode:2022QSRv..27607316D. doi:10.1016/j.quascirev.2021.107316. S2CID   245236379.
  6. Martin, Paul S.; Klein, Richard G. (1989). Quaternary Extinctions: A Prehistoric Revolution. University of Arizona Press. ISBN   978-0-8165-1100-6.
  7. Smith, Felisa A.; et al. (April 20, 2018). "Body size downgrading of mammals over the late Quaternary". Science. 360 (6386): 310–313. Bibcode:2018Sci...360..310S. doi: 10.1126/science.aao5987 . PMID   29674591.
  8. Koch, Paul L.; Barnosky, Anthony D. (2006-01-01). "Late Quaternary Extinctions: State of the Debate". Annual Review of Ecology, Evolution, and Systematics. 37 (1): 215–250. doi:10.1146/annurev.ecolsys.34.011802.132415. S2CID   16590668.
  9. Prado, José L.; Martinez-Maza, Cayetana; Alberdi, María T. (May 2015). "Megafauna extinction in South America: A new chronology for the Argentine Pampas". Palaeogeography, Palaeoclimatology, Palaeoecology. 425: 41–49. Bibcode:2015PPP...425...41P. doi:10.1016/j.palaeo.2015.02.026.
  10. Frédérik Saltré; Marta Rodríguez-Rey; Barry W. Brook; Christopher N. Johnson (2016). "Climate change not to blame for late Quaternary megafauna extinctions in Australia". Nature Communications. 7: 10511. Bibcode:2016NatCo...710511S. doi: 10.1038/ncomms10511 . PMC   4740174 . PMID   26821754.
  11. 1 2 Anthony D. Barnosky; Paul L. Koch; Robert S. Feranec; Scott L. Wing; Alan B. Shabel (2004). "Assessing the Causes of Late Pleistocene Extinctions on the Continents". Science. 306 (5693): 70–75. Bibcode:2004Sci...306...70B. CiteSeerX   10.1.1.574.332 . doi:10.1126/science.1101476. PMID   15459379. S2CID   36156087.
  12. Putshkov, P. V. (1997). "Were the Mammoths killed by the warming? (Testing of the climatic versions of the Wurm extinctions)". Vestnik Zoologii (Supplement No.4).
  13. 1 2 Gillespie, Richard (2008). "Updating Martin's global extinction model". Quaternary Science Reviews. 27 (27–28): 2522–2529. Bibcode:2008QSRv...27.2522G. doi:10.1016/j.quascirev.2008.09.007.
  14. Anderson, Paul K. (July 1995). "Competition, Predation, and the Evolution and Extinction of Steller's Sea Cow, Hydrodamalis Gigas". Marine Mammal Science. 11 (3): 391–4. Bibcode:1995MMamS..11..391A. doi:10.1111/j.1748-7692.1995.tb00294.x.
  15. 1 2 Bergman, Juraj; Pedersen, Rasmus Ø; Lundgren, Erick J.; Lemoine, Rhys T.; Monsarrat, Sophie; Pearce, Elena A.; Schierup, Mikkel H.; Svenning, Jens-Christian (24 November 2023). "Worldwide Late Pleistocene and Early Holocene population declines in extant megafauna are associated with Homo sapiens expansion rather than climate change". Nature Communications . 14 (1): 7679. doi: 10.1038/s41467-023-43426-5 . ISSN   2041-1723.
  16. Holliday VT, Bartlein PJ, Scott AC, Marlon JR (5 December 2019). "Extraordinary Biomass-Burning Episode and Impact Winter Triggered by the Younger Dryas Cosmic Impact ∼12,800 Years Ago, Parts 1 and 2: A Discussion". The Journal of Geology . 128 (1): 69–94. Bibcode:2020JG....128...69H. doi:10.1086/706264. ISSN   0022-1376. Wikidata   Q91978737.
  17. Lan, Tianying; Lindqvist, Charlotte (2018). "Paleogenomics: Genome-Scale Analysis of Ancient DNA and Population and Evolutionary Genomic Inferences". In Lindqvist, C.; Rajora, O. (eds.). Population Genomics. pp. 323–360. doi:10.1007/13836_2017_7. ISBN   978-3-030-04587-6.
  18. Faith, J. Tyler (January 2014). "Late Pleistocene and Holocene mammal extinctions on continental Africa". Earth-Science Reviews. 128: 105–121. Bibcode:2014ESRv..128..105F. doi:10.1016/j.earscirev.2013.10.009.
  19. Heinrich, Earl (31 October 2013). "Ancient Nubia" (PDF). Cambridge Online Histories.
  20. Turvey, Samuel T.; Sathe, Vijay; Crees, Jennifer J.; Jukar, Advait M.; Chakraborty, Prateek; Lister, Adrian M. (January 2021). "Late Quaternary megafaunal extinctions in India: How much do we know?". Quaternary Science Reviews. 252: 106740. Bibcode:2021QSRv..25206740T. doi:10.1016/j.quascirev.2020.106740. S2CID   234265221.
  21. Turvey, Samuel T.; Tong, Haowen; Stuart, Anthony J.; Lister, Adrian M. (September 2013). "Holocene survival of Late Pleistocene megafauna in China: a critical review of the evidence". Quaternary Science Reviews. 76: 156–166. Bibcode:2013QSRv...76..156T. doi:10.1016/j.quascirev.2013.06.030.
  22. Louys, Julien; Roberts, Patrick (2020-10-15). "Environmental drivers of megafauna and hominin extinction in Southeast Asia". Nature. 586 (7829): 402–406. Bibcode:2020Natur.586..402L. doi:10.1038/s41586-020-2810-y. hdl: 10072/402368 . ISSN   0028-0836. PMID   33029012. S2CID   222217295.
  23. Hasegawa, Y.; Okumura, Y.; Tatsukawa, H. (2009). "First record of Late Pleistocene Bison from the fissure deposits of the Kuzuu Limestone, Yamasuge,Sano-shi,Tochigi Prefecture, Japan" (PDF). Gunma Museum of Natural History and Kuzuu Fossil Museum. 13: 47–52.
  24. Kurosawa, Y. "モノが語る牛と人間の文化 – ② 岩手の牛たち" (PDF). LIAJ News No.109- Oshu City Cattle Museum p. 29-31.
  25. Bison hanaizumiensis - TrekGEO
  26. Rozzi, Roberto (2017-02-01). "A new extinct dwarfed buffalo from Sulawesi and the evolution of the subgenus Anoa: An interdisciplinary perspective". Quaternary Science Reviews. 157: 188–205. Bibcode:2017QSRv..157..188R. doi: 10.1016/j.quascirev.2016.12.011 .
  27. Geist, Valerius (1998). Deer of the World: Their Evolution, Behaviour, and Ecology. Stackpole Books. ISBN   9780811704960.
  28. Jukar, Advait M.; Patnaik, Rajeev; Chauhan, Parth R.; Li, Hong-Chun; Lin, Jih-Pai (2019-09-10). "The youngest occurrence of Hexaprotodon Falconer and Cautley, 1836 (Hippopotamidae, Mammalia) from South Asia with a discussion on its extinction". Quaternary International. AMS 14C Applications II. 528: 130–137. Bibcode:2019QuInt.528..130J. doi:10.1016/j.quaint.2019.01.005. ISSN   1040-6182. S2CID   133765385.
  29. "Rock paintings show species that roamed India". www.newindianexpress.com. Retrieved 2016-09-14.
  30. Feldhamer, George A.; Drickamer, Lee C.; Vessey, Stephen H.; Merritt, Joseph F.; Krajewski, Carey (2015). Mammalogy: Adaptation, Diversity, Ecology. JHU Press. ISBN   9781421415888.
  31. Watanabe, Junya; Matsuoka, Hiroshige (2015-11-02). "Flightless diving duck (Aves, Anatidae) from the Pleistocene of Shiriya, northeast Japan". Journal of Vertebrate Paleontology. 35 (6): e994745. Bibcode:2015JVPal..35E4745W. doi:10.1080/02724634.2014.994745. S2CID   83689129.
  32. Hoffecker, John F.; Elias, Scott A. (2012-05-29). Human Ecology of Beringia. Columbia University Press. ISBN   9780231503884.
  33. Rivals, Florent (2006). "Découverte de Capra caucasica et d'Hemitragus cedrensis (Mammalia, Bovidae) dans les niveaux du Pléistocène supérieur de la Caune de l'Arago (Tautavel, France): Implication biochronologique dans le contexte du Bassin Méditerranéen". Geobios. 39 (1): 85–102. Bibcode:2006Geobi..39...85R. doi:10.1016/j.geobios.2004.08.004.
  34. Creégut-Bonnoure, Evelyne (12 March 2009). "Biochronologie et grands mammifères au Pléistocène moyen et supérieur en Europe occidentale: l'Apport des genres hemitragus et capra". Quaternaire (in French). 20: 481–508. doi: 10.4000/quaternaire.5345 .
  35. Kurtén, Björn (2008). Pleistocene mammals of Europe. Aldine Transaction. ISBN   9780202309538. OCLC   751413776.
  36. Baryshnikov, G.; Tikhonov, A. (1994-10-01). "Notes on skulls of Pleistocene Saiga of Northern Eurasia". Historical Biology. 8 (1–4): 209–234. doi:10.1080/10292389409380478.
  37. 1 2 3 4 Hoffecker, John F.; Elias, Scott A. (2012-05-29). Human Ecology of Beringia. Columbia University Press. ISBN   9780231503884.
  38. 1 2 3 4 Vereshchagin, N. K.; Baryshnikov, G. F. (1991-01-01). "The ecological structure of the "Mammoth Fauna" in Eurasia". Annales Zoologici Fennici. 28 (3/4): 253–259. JSTOR   23735450.
  39. Sanz, Montserrat; Daura, Joan; Brugal, Jean-Philip (2014-01-01). "First occurrence of the extinct deer Haploidoceros in the Iberian Peninsula in the Upper Pleistocene of the Cova del Rinoceront (Castelldefels, Barcelona)". Comptes Rendus Palevol. 13 (1): 27–40. Bibcode:2014CRPal..13...27S. doi:10.1016/j.crpv.2013.06.005.
  40. Rivals, Florent; Sanz, Montserrat; Daura, Joan (2016-05-01). "First reconstruction of the dietary traits of the Mediterranean deer (Haploidoceros mediterraneus) from the Cova del Rinoceront (NE Iberian Peninsula)". Palaeogeography, Palaeoclimatology, Palaeoecology. 449: 101–107. Bibcode:2016PPP...449..101R. doi:10.1016/j.palaeo.2016.02.014.
  41. 1 2 Elias, Scott; Mock, Cary (2013-03-25). Encyclopedia of Quaternary Science. Newnes. ISBN   9780444536426.
  42. "Habitat conditions for Camelus knoblochi and factors in its extinction by Vadim V. Titov" (PDF).
  43. 1 2 3 Foronova, I. (2006). "Late quaternary equids (genus Equus) of South-western and South-central Siberia". In M. Mashkour (ed.). Equids in time and space. Papers in honour of Véra Eisenmann. Proceedings of the 9th conference of the International Council of Archaeozoology, Durham, August 2002. Oxbow Books. pp. 20–30.
  44. Chase, Philip G. (2009). The Cave of Fontéchevade: Recent Excavations and Their Paleoanthropological Implications. Cambridge University Press. ISBN   9780521898447.
  45. Yanko-Hombach, Valentina; Gilbert, Allan S.; Panin, Nicolae; Dolukhanov, Pavel M. (2006-11-15). The Black Sea Flood Question: Changes in Coastline, Climate and Human Settlement. Springer Science & Business Media. ISBN   9781402053023.
  46. Hopkins, David M.; Matthews, John V.; Schweger, Charles E. (2013). Paleoecology of Beringia. Elsevier Science. p. 349. ISBN   9781483273402.
  47. Lu, Dan; Yang, Yangheshan; Li, Qiang; Ni, Xijun (2021-07-30). "A late Pleistocene fossil from Northeastern China is the first record of the dire wolf (Carnivora: Canis dirus) in Eurasia". Quaternary International. Cave Deposits from Luotuo Hill, Northeast China: A Geochronologically Calibrated Mammalian Biostratigraphic Standard for the Quaternary of Eastern Asia. 591: 87–92. Bibcode:2021QuInt.591...87L. doi:10.1016/j.quaint.2020.09.054. ISSN   1040-6182. S2CID   224877090.
  48. Marciszak, Adrian; Schouwenburg, Charles; Lipecki, Grzegorz; Talamo, Sahra; Shpansky, Andrey; Malikov, Dmitriy; Gornig, Wiktoria (2019-12-01). "Steppe brown bear Ursus arctos "priscus" from the Late Pleistocene of Europe". Quaternary International. Quaternary Stratigraphy and Hominids Around Europe:: SEQS 2017 Meeting. 534: 158–170. Bibcode:2019QuInt.534..158M. doi:10.1016/j.quaint.2019.02.042. ISSN   1040-6182. S2CID   133750923.
  49. Münzel, Susanne C.; Rivals, Florent; Pacher, Martina; Döppes, Doris; Rabeder, Gernot; Conard, Nicholas J.; Bocherens, Hervé (2014-08-07). "Behavioural ecology of Late Pleistocene bears (Ursus spelaeus, Ursus ingressus): Insight from stable isotopes (C, N, O) and tooth microwear". Quaternary International. Fossil remains in karst and their role in reconstructing Quaternary paleoclimate and paleoenvironments. 339–340: 148–163. Bibcode:2014QuInt.339..148M. doi:10.1016/j.quaint.2013.10.020.
  50. Ghezzo, Elena; Boscaini, Alberto; Madurell-Malapeira, Joan; Rook, Lorenzo (2014-12-16). "Lynx remains from the Pleistocene of Valdemino cave (Savona, Northwestern Italy), and the oldest occurrence of Lynx spelaeus (Carnivora, Felidae)". Rendiconti Lincei. 26 (2): 87–95. doi:10.1007/s12210-014-0363-4. hdl: 11336/59435 . S2CID   85194755.
  51. "Search for images at Natural History Museum Picture Library". piclib.nhm.ac.uk. Archived from the original on 2016-04-19. Retrieved 2016-04-20.
  52. Stuart, Anthony John (May 2015). "Late Quaternary megafaunal extinctions on the continents: a short review: LATE QUATERNARY MEGAFAUNAL EXTINCTIONS". Geological Journal. 50 (3): 338–363. doi:10.1002/gj.2633. S2CID   128868400.
  53. Vershinina, Alisa O.; Heintzman, Peter D.; Froese, Duane G.; Zazula, Grant; Cassatt‐Johnstone, Molly; Dalén, Love; Der Sarkissian, Clio; Dunn, Shelby G.; Ermini, Luca; Gamba, Cristina; Groves, Pamela; Kapp, Joshua D.; Mann, Daniel H.; Seguin‐Orlando, Andaine; Southon, John (December 2021). "Ancient horse genomes reveal the timing and extent of dispersals across the Bering Land Bridge". Molecular Ecology. 30 (23): 6144–6161. Bibcode:2021MolEc..30.6144V. doi:10.1111/mec.15977. hdl: 10995/118212 . ISSN   0962-1083. PMID   33971056. S2CID   234360028.
  54. Cirilli, Omar; Machado, Helena; Arroyo-Cabrales, Joaquin; Barrón-Ortiz, Christina I.; Davis, Edward; Jass, Christopher N.; Jukar, Advait M.; Landry, Zoe; Marín-Leyva, Alejandro H.; Pandolfi, Luca; Pushkina, Diana; Rook, Lorenzo; Saarinen, Juha; Scott, Eric; Semprebon, Gina (2022-08-24). "Evolution of the Family Equidae, Subfamily Equinae, in North, Central and South America, Eurasia and Africa during the Plio-Pleistocene". Biology. 11 (9): 1258. doi: 10.3390/biology11091258 . ISSN   2079-7737. PMC   9495906 . PMID   36138737.
  55. Mitchell, Kieren J.; Bover, Pere; Salis, Alexander T.; Mudge, Caitlin; Heiniger, Holly; Thompson, Mary; Hockett, Bryan; Weyrich, Laura S.; Cooper, Alan; Meachen, Julie A. (2023-02-20). "Evidence for Pleistocene gene flow through the ice-free corridor from extinct horses and camels from Natural Trap Cave, Wyoming". Quaternary International. 647–648: 71–80. doi:10.1016/j.quaint.2021.11.017.
  56. Lundelius, Ernest L; Bryant, Vaughn M; Mandel, Rolfe; Thies, Kenneth J; Thoms, Alston (2013). "The First Occurrence of a Toxodont (Mammalia, Notoungulata) in the United States". Journal of Vertebrate Paleontology. 33 (1): 229–232. Bibcode:2013JVPal..33..229L. doi:10.1080/02724634.2012.711405. hdl: 1808/13587 . JSTOR   23361085. S2CID   53601518 . Retrieved 2016-01-23.
  57. Rodriguez-de la Rosa, Ruben A.; Guzman Gutierrez, Jose Ruben; Mendoza, Carlos (December 2001). "A New Occurrence of Toxodonts in the Pleistocene of México". Current Research in the Pleistocene. 28: 29–30. Retrieved 2016-01-23.
  58. "Fossilworks: Lynx lynx". fossilworks.org. Retrieved 17 December 2021.
  59. 1 2 Schubert, Blaine W.; Chatters, James C.; Arroyo-Cabrales, Joaquin; Samuels, Joshua X.; Soibelzon, Leopoldo H.; Prevosti, Francisco J.; Widga, Christopher; Nava, Alberto; Rissolo, Dominique; Erreguerena, Pilar Luna (2019-05-31). "Yucatán carnivorans shed light on the Great American Biotic Interchange". Biology Letters. 15 (5): 20190148. doi:10.1098/rsbl.2019.0148. PMC   6548739 . PMID   31039726.
  60. Youngman, Phillip M. (1986-03-01). "The extinct short-faced skunk Brachyprotoma obtusata (Mammalia, Carnivora): first records for Canada and Beringia". Canadian Journal of Earth Sciences. 23 (3): 419–424. Bibcode:1986CaJES..23..419Y. doi:10.1139/e86-043.
  61. Anderson, E. (1973). "Ferrets from the pleistocene of central Alaska". J. Mammal. 54 (3): 778–779. doi:10.2307/1378982. JSTOR   1378982.
  62. 1 2 3 4 5 6 7 8 9 10 11 Sanz, Nuria (31 December 2015). Human origin sites and the World Heritage Convention in Americas. Vol. 1. Paris: UNESCO. ISBN   9789231001406. OCLC   1002234186.
  63. Ceballos, Gerardo; Arroyo-Cabrales, Joaquín; Ponce, Eduardo (2010). "Effects of Pleistocene environmental changes on the distribution and community structure of the mammalian fauna of Mexico". Quaternary Research. 73 (3): 464–473. Bibcode:2010QuRes..73..464C. doi:10.1016/j.yqres.2010.02.006. S2CID   73620371.
  64. Lucas, Spencer G.; Morgan, Gary S.; Spielmann, Justin A.; Prothero, Donald R. (2008). Neogene Mammals: Bulletin 44. New Mexico Museum of Natural History and Science.
  65. Sanchez, Guadalupe; Holliday, Vance T.; Gaines, Edmund P.; Arroyo-Cabrales, Joaquín; Martínez-Tagüeña, Natalia; Kowler, Andrew; Lange, Todd; Hodgins, Gregory W. L.; Mentzer, Susan M. (2014-07-29). "Human (Clovis)–gomphothere (Cuvieronius sp.) association ~13,390 calibrated yBP in Sonora, Mexico". Proceedings of the National Academy of Sciences. 111 (30): 10972–10977. Bibcode:2014PNAS..11110972S. doi: 10.1073/pnas.1404546111 . PMC   4121807 . PMID   25024193.
  66. "Fossilworks: Erethizon kleini". fossilworks.org. Retrieved 17 December 2021.
  67. McDonald, H. Gregory; Chatters, James C.; Gaudin, Timothy J. (2017-05-04). "A new genus of megalonychid ground sloth (Mammalia, Xenarthra) from the late Pleistocene of Quintana Roo, Mexico". Journal of Vertebrate Paleontology. 37 (3): e1307206. Bibcode:2017JVPal..37E7206M. doi:10.1080/02724634.2017.1307206. ISSN   0272-4634. S2CID   90414512.
  68. "Ice Age Predators Found Alongside Oldest Human in Americas". 2017-08-25. Archived from the original on August 25, 2017. Retrieved 2017-10-13.
  69. Stinnesbeck, Sarah R.; Frey, Eberhard; Olguín, Jerónimo Avíles; Stinnesbeck, Wolfgang; Zell, Patrick; Mallison, Heinrich; González, Arturo González; Núñez, Eugenio Aceves; Morlet, Adriana Velázquez (2017-06-01). "Xibalbaonyx oviceps, a new megalonychid ground sloth (Folivora, Xenarthra) from the Late Pleistocene of the Yucatán Peninsula, Mexico, and its paleobiogeographic significance". PalZ. 91 (2): 245–271. doi:10.1007/s12542-017-0349-5. ISSN   0031-0220. S2CID   134188352.
  70. "Ancient Giant Sloth Fossil Found in Underwater Cave". 2017-08-18. Archived from the original on August 19, 2017. Retrieved 2017-10-13.
  71. Debus, Allen (June 2002). Dinosaur Memories. iUniverse. ISBN   9780595229888.
  72. McDonough, Colleen M.; Loughry, W. J. (2013-03-18). The Nine-Banded Armadillo: A Natural History. University of Oklahoma Press. ISBN   9780806189215.
  73. Zicha, Ondrej. "BioLib: Biological library". www.biolib.cz. Retrieved 2016-04-12.
  74. "Fossilworks: Phoenicopterus copei". fossilworks.org. Retrieved 17 December 2021.
  75. Feduccia, Alan (1999). The Origin and Evolution of Birds. Yale University Press. ISBN   978-0300078619.
  76. Gillespie, Rosemary G.; Clague, D. A. (2009). Encyclopedia of Islands. University of California Press. ISBN   9780520256491.
  77. Turvey, Sam (2009-05-28). Holocene Extinctions. Oxford: Oxford University Press. ISBN   9780199535095.
  78. MacPhee, R.D.E. (1999). Extinctions in Near Time: Causes, Contexts, and Consequences. Kluwer Academic Publishers. ISBN   978-0-306-46092-0.[ page needed ]
  79. Bell, C.J.; et al. (2004). "The Blancan, Irvingtonian, and Rancholabrean mammal ages". In Woodburne, M.O. (ed.). Late Cretaceous and Cenozoic Mammals of North America: Biostratigraphy and Geochronology. New York: Columbia Univ. Press. pp. 232–314. ISBN   978-0-231-13040-0.
  80. Scott, E.; Cox, S.M. (2008). "Late Pleistocene distribution of Bison (Mammalia; Artiodactyla) in the Mojave Desert of Southern California and Nevada". In Wang, X.; Barnes, L.G. (eds.). Geology and Vertebrate Paleontology of Western and Southern North America. Los Angeles: Natural History Museum of Los Angeles County. pp. 359–382.
  81. Sanders, A.E.; R.E. Weems; L.B. Albright III (2009). "Formalization of the mid-Pleistocene "Ten Mile Hill beds" in South Carolina with evidence for placement of the Irvingtonian–Rancholabrean boundary". In Albright III, L.B. (ed.). Papers on Geology, Vertebrate Paleontology, and Biostratigraphy in Honor of Michael O. Woodburne. Flagstaff: Museum of Northern Arizona. pp. 369–375.
  82. Shapiro, B.; et al. (2004). "Rise and Fall of the Beringian Steppe Bison". Science. 306 (5701): 1561–1565. Bibcode:2004Sci...306.1561S. doi:10.1126/science.1101074. PMID   15567864. S2CID   27134675.
  83. Wilson, M.C.; L.V. Hills; B. Shapiro (2008). "Late Pleistocene northward-dispersing Bos antiquus from the Bighill Creek Formation, Gallelli Gravel Pit, Alberta, Canada, and the fate of Bison occidentalis". Canadian Journal of Earth Sciences. 45 (7): 827–859. Bibcode:2008CaJES..45..827W. doi:10.1139/E08-027. S2CID   129131047.
  84. Prates, Luciano; Perez, S. Ivan (2021-04-12). "Late Pleistocene South American megafaunal extinctions associated with rise of Fishtail points and human population". Nature Communications. 12 (1): 2175. Bibcode:2021NatCo..12.2175P. doi:10.1038/s41467-021-22506-4. ISSN   2041-1723. PMC   8041891 . PMID   33846353.
  85. "Fossilworks: Agalmaceros". fossilworks.org. Retrieved 17 December 2021.
  86. "Fossilworks: Agalmaceros blicki". fossilworks.org. Retrieved 17 December 2021.
  87. "Fossilworks: Odocoileus salinae". fossilworks.org. Retrieved 17 December 2021.
  88. 1 2 Turvey, Samuel T. (2009-05-28). Holocene Extinctions. Oxford University Press. ISBN   9780191579981.
  89. 1 2 MacFadden, Bruce J. (2013-03-20). "Dispersal of Pleistocene Equus (Family Equidae) into South America and Calibration of GABI 3 Based on Evidence from Tarija, Bolivia". PLOS ONE. 8 (3): e59277. Bibcode:2013PLoSO...859277M. doi: 10.1371/journal.pone.0059277 . ISSN   1932-6203. PMC   3603859 . PMID   23527150.
  90. Pereira, Jamil Corrêa; Lopes, Renato Pereira; Kerber, Leonardo (2012). "New remains of Late Pleistocene mammals from the Chuí Creek, Southern Brazil". Revista Brasileira de Paleontologia. 15 (2): 228–239. doi: 10.4072/rbp.2012.2.10 .
  91. Ransom, Jason I.; Kaczensky, Petra (2016-05-15). Wild Equids: Ecology, Management, and Conservation. JHU Press. ISBN   9781421419107.
  92. "Late Pleistocene vertebrates from Touro Passo Creek (Touro Passo Formation), southern Brazil: A review". Revista Mexicana de Ciencias Geológicas. 31 (2): 248–259. Retrieved 2017-09-02.
  93. "Fossilworks: Macraucheniopsis". fossilworks.org. Retrieved 17 December 2021.
  94. Gaudioso, Pablo Javier; Gasparini, Germán M.; Herbst, Rafael; Barquez, Rubén Mario (2017-03-16). "First record of the Neolicaphrium recens Frenguelli, 1921 (Mammalia, Litopterna) in the Pleistocene of Santiago del Estero Province, Argentina". Papéis Avulsos de Zoologia. 57 (3): 23–29. doi: 10.11606/0031-1049.2017.57.03 . ISSN   1807-0205.
  95. Kurtén, Björn; Werdelin, Lars (1990). "Relationships between North and South American Smilodon". Journal of Vertebrate Paleontology. 10 (2): 158–169. Bibcode:1990JVPal..10..158K. doi:10.1080/02724634.1990.10011804. JSTOR   4523312.
  96. Chimento, Nicolás R.; Agnolin, Federico L. (November 2017). "The fossil American lion ( Panthera atrox ) in South America: Palaeobiogeographical implications". Comptes Rendus Palevol. 16 (8): 850–864. Bibcode:2017CRPal..16..850C. doi: 10.1016/j.crpv.2017.06.009 .
  97. Prevosti, F. J.; Tonni, E. P.; Bidegain, J. C. (2009-12-01). "Stratigraphic range of the large canids (Carnivora, Canidae) in South America, and its relevance to quaternary biostratigraphy". Quaternary International. The Ensenadan Stage/Age in southern South America. 210 (1): 76–81. Bibcode:2009QuInt.210...76P. doi:10.1016/j.quaint.2009.06.034. hdl: 11336/103725 . ISSN   1040-6182.
  98. "Fossilworks: Valgipes". fossilworks.org. Retrieved 17 December 2021.
  99. Haynes, Gary (2008-12-23). American Megafaunal Extinctions at the End of the Pleistocene. Springer Science & Business Media. ISBN   9781402087936.
  100. Prothero, Donald R. (2016-11-15). The Princeton Field Guide to Prehistoric Mammals. Princeton University Press. ISBN   9781400884452.
  101. Eisenberg, John F.; Redford, Kent H. (2000-05-15). Mammals of the Neotropics, Volume 3: Ecuador, Bolivia, Brazil. University of Chicago Press. ISBN   9780226195421.
  102. Fariña, Richard A.; Vizcaíno, Sergio F.; Iuliis, Gerry De (2013-05-22). Megafauna: Giant Beasts of Pleistocene South America. Indiana University Press. ISBN   978-0253007193.
  103. 1 2 3 Oliveira, Édison V.; Porpino, Kleberson O.; Barreto, Alcina F. (2010). "On the presence of Glyptotherium in the Late Pleistocene of Northeastern Brazil, and the status of "Glyptodon" and "Chlamydotherium". Paleobiogeographic implications". Neues Jahrbuch für Geologie und Paläontologie - Abhandlungen. 258 (3): 353–363. doi:10.1127/0077-7749/2010/0116.
  104. Martin, Paul Schultz (2005). Twilight of the Mammoths: Ice Age Extinctions and the Rewilding of America . University of California Press. ISBN   9780520231412.
  105. "Fossilworks: Neuryurus". fossilworks.org. Retrieved 17 December 2021.
  106. "Fossilworks: Parapanochthus". fossilworks.org. Retrieved 17 December 2021.
  107. Martin, Robert A.; Martin, Robert Allen; Barnosky, Anthony D. (2005-10-06). Morphological Change in Quaternary Mammals of North America. Cambridge University Press. ISBN   9780521020817.
  108. Turvey, Sam (2009-05-28). Holocene Extinctions. Oxford: Oxford University Press. ISBN   9780199535095.
  109. 1 2 Góis, Flávio; Ruiz, Laureano Raúl González; Scillato-Yané, Gustavo Juan; Soibelzon, Esteban (2015-06-17). "A Peculiar New Pampatheriidae (Mammalia: Xenarthra: Cingulata) from the Pleistocene of Argentina and Comments on Pampatheriidae Diversity". PLOS ONE. 10 (6): e0128296. Bibcode:2015PLoSO..1028296G. doi: 10.1371/journal.pone.0128296 . PMC   4470999 . PMID   26083486.
  110. Jones, Washington; Rinderknecht, Andrés; Alvarenga, Herculano; Montenegro, Felipe; Ubilla, Martín (2017-12-30). "The last terror birds (Aves, Phorusrhacidae): new evidence from the late Pleistocene of Uruguay". PalZ. 92 (2): 365–372. doi:10.1007/s12542-017-0388-y. ISSN   0031-0220. S2CID   134344096.
  111. Alvarenga, Herculano; Jones, Washington; Rinderknecht, Andrés (2010-05-01). "The youngest record of phorusrhacid birds (Aves, Phorusrhacidae) from the late Pleistocene of Uruguay". Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen. 256 (2): 229–234. doi:10.1127/0077-7749/2010/0052.
  112. Agnolin, Federico (2013). "La posición sistemática de Hermosiornis (Aves, Phororhacoidea) y sus implicancias filogenéticas". Revista del Museo Argentino de Ciencias Naturales. Nueva Series. 15 (1): 39–60. doi: 10.22179/revmacn.15.167 . ISSN   1853-0400.
  113. Jones, Washington; Rinderknecht, Andrés; Migotto, Rafael; Blanco, R. Ernesto (2013). "Body Mass Estimations and Paleobiological Inferences on a New Species of Large Caracara (Aves, Falconidae) from the Late Pleistocene of Uruguay". Journal of Paleontology. 87 (1): 151–158. Bibcode:2013JPal...87..151J. doi:10.1666/12-026R.1. JSTOR   23353814. S2CID   83648963.
  114. Suárez, William; Olson, Storrs L. (2014-09-01). "A new fossil species of small crested caracara (Aves: Falconidae: Caracara) from the Pacific lowlands of western South America". Proceedings of the Biological Society of Washington. 127 (2): 299–310. doi:10.2988/0006-324X-127.2.299. ISSN   0006-324X. S2CID   130085421.
  115. "Fossilworks: Milvago brodkorbi". fossilworks.org. Retrieved 17 December 2021.
  116. David, Bruno; Arnold, Lee J.; Delannoy, Jean-Jacques; Fresløv, Joanna; Urwin, Chris; Petchey, Fiona; McDowell, Matthew C.; Mullett, Russell; Mialanes, Jerome; Wood, Rachel; Crouch, Joe; Berthet, Johan; Wong, Vanessa N.L.; Green, Helen; Hellstrom, John; GunaiKurnai Land; Waters Aboriginal Corporation (2021-02-01). "Late survival of megafauna refuted for Cloggs Cave, SE Australia: Implications for the Australian Late Pleistocene megafauna extinction debate". Quaternary Science Reviews. 253: 106781. Bibcode:2021QSRv..25306781D. doi:10.1016/j.quascirev.2020.106781. ISSN   0277-3791. S2CID   234010059.
  117. Price, Gilbert J.; Zhao, Jian-xin; Feng, Yue-Xing; Hocknull, Scott A. (2009-02-01). "New U/Th ages for Pleistocene megafauna deposits of southeastern Queensland, Australia". Journal of Asian Earth Sciences. 34 (2): 190–197. Bibcode:2009JAESc..34..190P. doi:10.1016/j.jseaes.2008.04.008. ISSN   1367-9120.
  118. "Humans, not climate change, wiped out Australian megafauna". CU Boulder Today. 2017-01-20. Retrieved 2021-05-05.
  119. Wilson, Lito Vilisoni (2020-05-18). "Mysteries of megafauna extinction unlocked". Newsroom. Retrieved 2021-05-05.
  120. Hocknull, Scott A.; Lewis, Richard; Arnold, Lee J.; Pietsch, Tim; Joannes-Boyau, Renaud; Price, Gilbert J.; Moss, Patrick; Wood, Rachel; Dosseto, Anthony; Louys, Julien; Olley, Jon (2020-05-18). "Extinction of eastern Sahul megafauna coincides with sustained environmental deterioration". Nature Communications. 11 (1): 2250. Bibcode:2020NatCo..11.2250H. doi: 10.1038/s41467-020-15785-w . ISSN   2041-1723. PMC   7231803 . PMID   32418985.
  121. Saltré, Frédérik; Chadoeuf, Joël; Peters, Katharina J.; McDowell, Matthew C.; Friedrich, Tobias; Timmermann, Axel; Ulm, Sean; Bradshaw, Corey J. A. (2019-11-22). "Climate-human interaction associated with southeast Australian megafauna extinction patterns". Nature Communications. 10 (1): 5311. Bibcode:2019NatCo..10.5311S. doi: 10.1038/s41467-019-13277-0 . ISSN   2041-1723. PMC   6876570 . PMID   31757942.
  122. DeSantis, Larisa R. G.; Field, Judith H.; Wroe, Stephen; Dodson, John R. (2017-01-26). "Dietary responses of Sahul (Pleistocene Australia–New Guinea) megafauna to climate and environmental change". Paleobiology. 43 (2): 181–195. Bibcode:2017Pbio...43..181D. doi: 10.1017/pab.2016.50 . ISSN   0094-8373.
  123. Martin, Paul S.; Klein, Richard G. (1989). Quaternary Extinctions: A Prehistoric Revolution. University of Arizona Press. ISBN   9780816511006.
  124. 1 2 Lourandos, Harry (1997-02-28). Continent of Hunter-Gatherers: New Perspectives in Australian Prehistory. Cambridge University Press. ISBN   9780521359467.
  125. 1 2 MacPhee, R.D.E. (1999). Extinctions in Near Time: Causes, Contexts, and Consequences. Kluwer Academic Publishers. p. 251. ISBN   978-0-306-46092-0.
  126. Long, John A.; Archer, Michael (2002). Prehistoric Mammals of Australia and New Guinea: One Hundred Million Years of Evolution. UNSW Press. ISBN   9780868404356.
  127. Tyndale-Biscoe, Hugh (2005-04-22). Life of Marsupials. Csiro Publishing. ISBN   9780643099210.
  128. Marwick, Ben. "Buried tools and pigments tell a new history of humans in Australia for 65,000 years". The Conversation. Retrieved 2017-09-07.
  129. Webb, Steve (2013-02-27). Corridors to Extinction and the Australian Megafauna. Newnes. ISBN   9780124078406.
  130. "Megafauna". austhrutime.com. Retrieved 2017-04-20.
  131. "Anaspides.net". www.anaspides.net. Archived from the original on 2017-07-20. Retrieved 2017-04-20.
  132. Webb, Steve (2013-02-27). Corridors to Extinction and the Australian Megafauna. Newnes. ISBN   9780124078406.
  133. Long, John A.; Archer, Michael (2002). Prehistoric Mammals of Australia and New Guinea: One Hundred Million Years of Evolution. UNSW Press. ISBN   9780868404356.
  134. White, Lauren C.; Saltré, Frédérik; Bradshaw, Corey J. A.; Austin, Jeremy J. (January 2018). "High-quality fossil dates support a synchronous, Late Holocene extinction of devils and thylacines in mainland Australia". Biology Letters. 14 (1): 20170642. doi:10.1098/rsbl.2017.0642. ISSN   1744-9561. PMC   5803592 . PMID   29343562.
  135. Bayly, I. A. E. (1993). "The fauna of athalassic saline waters in Australia and the Altiplano of South America: Comparisons and historical perspectives". In Hurlbert, Stuart H. (ed.). Saline Lakes V. Developments in Hydrobiology. Vol. 87. Springer Netherlands. pp. 225–231. doi:10.1007/978-94-011-2076-0_18. ISBN   9789401049214.
  136. Doughty, C. E.; Wolf, A.; Field, C. B. (2010). "Biophysical feedbacks between the Pleistocene megafauna extinction and climate: The first human‐induced global warming?". Geophys. Res. Lett. 37 (15): L15703. Bibcode:2010GeoRL..3715703D. doi: 10.1029/2010GL043985 . S2CID   54849882.
  137. Grayson, Donald K.; Meltzer, David J. (December 2012). "Clovis Hunting and Large Mammal Extinction: A Critical Review of the Evidence". Journal of World Prehistory. 16 (4): 313–359. doi:10.1023/A:1022912030020. S2CID   162794300.
  138. Kolbert, Elizabeth (2014). The Sixth Extinction: An Unnatural History. Bloomsbury Publishing. ISBN   9781408851210.
  139. Perry, George L. W.; Wheeler, Andrew B.; Wood, Jamie R.; Wilmshurst, Janet M. (2014-12-01). "A high-precision chronology for the rapid extinction of New Zealand moa (Aves, Dinornithiformes)". Quaternary Science Reviews. 105: 126–135. Bibcode:2014QSRv..105..126P. doi:10.1016/j.quascirev.2014.09.025.
  140. Crowley, Brooke E. (2010-09-01). "A refined chronology of prehistoric Madagascar and the demise of the megafauna". Quaternary Science Reviews. Special Theme: Case Studies of Neodymium Isotopes in Paleoceanography. 29 (19–20): 2591–2603. Bibcode:2010QSRv...29.2591C. doi:10.1016/j.quascirev.2010.06.030.
  141. Li, Sophia (2012-09-20). "Has Plant Life Reached Its Limits?". Green Blog. Retrieved 2016-01-22.
  142. Martin, P. S. (1963). The last 10,000 years: A fossil pollen record of the American Southwest. Tucson, AZ: Univ. Ariz. Press. ISBN   978-0-8165-1759-6.
  143. Martin, P. S. (1967). "Prehistoric overkill". In Martin, P. S.; Wright, H. E. (eds.). Pleistocene extinctions: The search for a cause. New Haven: Yale Univ. Press. ISBN   978-0-300-00755-8.
  144. Martin, P. S. (1989). "Prehistoric overkill: A global model". In Martin, P.S.; Klein, R.G. (eds.). Quaternary extinctions: A prehistoric revolution. Tucson, AZ: Univ. Arizona Press. pp. 354–404. ISBN   978-0-8165-1100-6.
  145. Weston, Phoebe (25 November 2022). "Humans v nature: our long and destructive journey to the age of extinction". The Guardian. Retrieved 28 November 2022.
  146. Martin, Paul S. (1966). "Africa and Pleistocene overkill" (PDF). Nature. 212 (5060): 339–342. Bibcode:1966Natur.212..339M. doi:10.1038/212339a0. S2CID   27013299. Archived from the original (PDF) on 2020-11-29. Retrieved 2019-07-31.
  147. Martin, P. S. (2005). Twilight of the Mammoths: Ice Age Extinctions and the Rewilding of America . University of California Press. ISBN   978-0-520-23141-2.
  148. Burney, D. A.; Flannery, T. F. (July 2005). "Fifty millennia of catastrophic extinctions after human contact" (PDF). Trends in Ecology & Evolution. 20 (7): 395–401. doi:10.1016/j.tree.2005.04.022. PMID   16701402. Archived from the original (PDF) on 2010-06-10.
  149. Diamond J (2008). "Palaeontology: The last giant kangaroo". Nature. 454 (7206): 835–6. Bibcode:2008Natur.454..835D. doi:10.1038/454835a. PMID   18704074. S2CID   36583693.
  150. Turneya CS, Flannery TF, Roberts RG, Reid C, Fifield LK, Higham TF, Jacobs Z, Kemp N, Colhoun EA, Kalin RM, Ogle N (26 August 2008). "Late-surviving megafauna in Tasmania, Australia, implicate human involvement in their extinction". Proc. Natl. Acad. Sci. U.S.A. 105 (34): 12150–3. Bibcode:2008PNAS..10512150T. doi: 10.1073/pnas.0801360105 . PMC   2527880 . PMID   18719103.
  151. "Humans responsible for demise of gigantic ancient mammals". University of Exeter. 13 August 2015. Retrieved 14 August 2015.
  152. Lewis J. Bartlett; David R. Williams; Graham W. Prescott; Andrew Balmford; Rhys E. Green; Anders Eriksson; Paul J. Valdes; Joy S. Singarayer; Andrea Manica (2016). "Robustness despite uncertainty: regional climate data reveal the dominant role of humans in explaining global extinctions of Late Quaternary megafauna". Ecography . 39 (2): 152–161. Bibcode:2016Ecogr..39..152B. doi: 10.1111/ecog.01566 .
  153. "Humans, not climate, have driven rapidly rising mammal extinction rate". phys.org. Retrieved 9 October 2020.
  154. Andermann, Tobias; Faurby, Søren; Turvey, Samuel T.; Antonelli, Alexandre; Silvestro, Daniele (September 2020). "The past and future human impact on mammalian diversity". Science Advances. 6 (36). eabb2313. Bibcode:2020SciA....6.2313A. doi:10.1126/sciadv.abb2313. ISSN   2375-2548. PMC   7473673 . PMID   32917612. CC-BY icon.svg Text and images are available under a Creative Commons Attribution 4.0 International License.
  155. Lemoine, Rhys Taylor; Buitenwerf, Robert; Svenning, Jens-Christian (1 December 2023). "Megafauna extinctions in the late-Quaternary are linked to human range expansion, not climate change". Anthropocene. 44: 100403. doi: 10.1016/j.ancene.2023.100403 . ISSN   2213-3054.
  156. Flannery, Tim (2002-10-16). The future eaters: an ecological history of the Australasian lands and people. New York: Grove/Atlantic, Inc. ISBN   978-0-8021-3943-6. OCLC   32745413.
  157. Diamond, J. (1984). "Historic extinctions: a Rosetta stone for understanding prehistoric extinctions". In Martin, P.S.; Klein, R.G. (eds.). Quaternary extinctions: A prehistoric revolution. Tucson, AZ: Univ. Arizona Press. pp. 824–62. ISBN   978-0-8165-1100-6.
  158. Diamond, J. (1997). Guns, germs, and steel; the fates of human societies. New York: Norton. ISBN   978-0-393-31755-8.
  159. Mossiman, J. E. & Martin, P. S. (1975). "Simulating Overkill by Paleoindians". American Scientist. 63 (3): 304–13. Bibcode:1975AmSci..63..304M.
  160. Whittington, S. L. & Dyke, B. (1984). "Simulating overkill: experiment with the Mossiman and Martin model". In Martin, P.S. & Klein, R.G. (eds.). Quaternary extinctions: A prehistoric revolution. Tucson, AZ: Univ. Arizona Press. pp. 451–66. ISBN   978-0-8165-1100-6.
  161. 1 2 Alroy, J. (2001). "A multispecies overkill simulation of the end-Pleistocene megafaunal mass extinction" (PDF). Science. 292 (5523): 1893–6. Bibcode:2001Sci...292.1893A. doi:10.1126/science.1059342. PMID   11397940. S2CID   35769639. Archived from the original (PDF) on 2008-05-11. Retrieved 2008-08-09.
  162. Meltzer, David J. (21 October 2015). "Pleistocene Overkill and North American Mammalian Extinctions". Annual Review of Anthropology . 44 (1): 33–53. doi: 10.1146/annurev-anthro-102214-013854 . ISSN   0084-6570.
  163. Grayson, D.K.; Meltzer, D.J. (2003). "A requiem for North American overkill". Journal of Archaeological Science. 30 (5): 585–593. Bibcode:2003JArSc..30..585G. CiteSeerX   10.1.1.399.1153 . doi:10.1016/s0305-4403(02)00205-4.
  164. Fiedel, S.; Haynes, G. (2004). "A premature burial: comments on Grayson and Meltzer's Requiem for overkill". Journal of Archaeological Science. 31 (1): 121–131. Bibcode:2004JArSc..31..121F. doi:10.1016/j.jas.2003.06.004.
  165. Surovell, Todd A; Brigid S Grund (2012). "The associational critique of Quaternary overkill". American Antiquity. 77 (4): 673–688. doi:10.7183/0002-7316.77.4.672. S2CID   163726293.
  166. Nadasdy, Paul (2006). "Transcending the Debate over the Ecologically Noble Indian: Indigenous Peoples and Environmentalism". Ethnohistory. 52 (2): 291–331. doi:10.1215/00141801-52-2-291. S2CID   52037849.
  167. Whitney-Smith, E. (2004). "Late Pleistocene extinctions through second-order predation". In Barton, C.M.; Clark, G.A.; Yesner, D.R. (eds.). Settlement of the American Continents: A Multidisciplinary Approach to Human Biogeography. Tucson, AZ: University of Arizona Press. ISBN   978-0-8165-2323-8.
  168. Whitney-Smith, E. (2009). The Second-Order Predation Hypothesis of Pleistocene Extinctions: A System Dynamics Model. Saarbruken, Germany: VDM Verlag. ISBN   978-3-639-11579-6.
  169. 1 2 Whitney-Smith, E. (2006). Clovis and Extinctions – Overkill, Second Order Predation, Environmental Degradation in a Non-equilibrium Ecosystem "Clovis Age Continent". University of New Mexico Press.
  170. Scott, E. (2010). "Extinctions, scenarios, and assumptions: Changes in latest Pleistocene large herbivore abundance and distribution in western North America". Quat. Int.
  171. Fløjgaard, Camilla; Pedersen, Pil Birkefeldt Møller; Sandom, Christopher J.; Svenning, Jens‐Christian; Ejrnæs, Rasmus (January 2022). "Exploring a natural baseline for large‐herbivore biomass in ecological restoration". Journal of Applied Ecology. 59 (1): 18–24. doi: 10.1111/1365-2664.14047 . ISSN   0021-8901. S2CID   243489626.
  172. 1 2 Rabanus-Wallace, M. Timothy; Wooller, Matthew J.; Zazula, Grant D.; Shute, Elen; Jahren, A. Hope; Kosintsev, Pavel; Burns, James A.; Breen, James; Llamas, Bastien; Cooper, Alan (2017). "Megafaunal isotopes reveal role of increased moisture on rangeland during late Pleistocene extinctions". Nature Ecology & Evolution. 1 (5): 0125. doi:10.1038/s41559-017-0125. PMID   28812683. S2CID   4473573.
  173. Birks, H.H. (1973). "Modern macrofossil assemblages in lake sediments in Minnesota". In Birks, H.J.B.; West, R.G. (eds.). Quaternary plant ecology: the 14th symposium of the British Ecological Society, University of Cambridge, 28–30 March 1972. Oxford: Blackwell Scientific. ISBN   0-632-09120-7.
  174. Birks, H.J.B.; Birks, H.H. (1980). Quaternary paleoecology. Baltimore: Univ. Park Press. ISBN   978-1-930665-56-9.
  175. Davis, M. B. (1976). "Pleistocene biogeography of temperate deciduous forests". Geoscience and man: Ecology of the Pleistocene. Vol. 13. Baton Rouge: School of Geoscience, Louisiana State Univ.
  176. Guthrie, R. D. (1988). Frozen Fauna of the Mammoth Steppe: The Story of Blue Babe. University Of Chicago Press. ISBN   978-0-226-31122-7.
  177. Guthrie, R. D. (1989). "Mosaics, allochemics, and nutrients: an ecological theory of Late Pleistocene megafaunal extinctions". In Martin, P.S.; Klein, R.G. (eds.). Quaternary extinctions: A prehistoric revolution. Tucson, AZ: Univ. Arizona Press. pp. 259–99. ISBN   978-0-8165-1100-6.
  178. Hoppe, P.P. (1978). "Rumen fermentation in African ruminants". Proceedings of the 13th Annual Congress of Game Biologists. Atlanta.
  179. Bryson, R.A.; Baerreis, D.A.; Wendland, W.M. (1970). "The character of late-glacial and post-glacial climatic changes" . In Dort Jr., W.; Jones, Jr. J.K. (eds.). Pleistocene and recent environments of the central Great Plains. Lawrence: Univ. Press Kan. ISBN   978-0-7006-0063-2. Univ. Kan. Spec. Pub. 3.
  180. Graham, R.W.; Lundelius, E.L. (1989). "Coevolutionary disequilibrium and Pleistocene extinctions". In Martin, P.S.; Klein, R.G. (eds.). Quaternary extinctions: A prehistoric revolution. Tucson, AZ: Univ. Arizona Press. pp. 354–404. ISBN   978-0-8165-1100-6.
  181. King, J.E.; Saunders, J.J. (1989). "Environmental insularity and the extinction of the American mastodont". In Martin, P.S.; Klein R.G. (eds.). Quaternary extinctions: A prehistoric revolution. Tucson, AZ: Univ. Arizona Press. pp. 354–404. ISBN   978-0-8165-1100-6.
  182. Axelrod, D. I. (1967). "Quaternary extinctions of large mammals". University of California Publications in Geological Sciences. 74: 1–42. ASIN B0006BX8LG.
  183. Slaughter, B. H. (1967). "Animal ranges as a clue to late-Pleistocene extinction". In Martin, P.S.; Wright H.E. (eds.). Pleistocene extinctions: The search for a cause. New Haven: Yale Univ. Press. ISBN   978-0-300-00755-8.
  184. Kilti, R. A. (1988). "Seasonality, gestation time, and large mammal extinctions". In Martin, P.S.; Klein R.G. (eds.). Quaternary extinctions: A prehistoric revolution. Tucson, AZ: Univ. Arizona Press. pp. 354–404. ISBN   978-0-8165-1100-6.
  185. Louys, Julien; Curnoe, D.; Tong, H. (2007). "Characteristics of Pleistocene megafauna extinctions in Southeast Asia" (PDF). Palaeogeography, Palaeoclimatology, Palaeoecology. 243 (1–2): 152–173. Bibcode:2007PPP...243..152L. doi:10.1016/j.palaeo.2006.07.011.
  186. Mann, Daniel H.; Groves, Pamela; Reanier, Richard E.; Gaglioti, Benjamin V.; Kunz, Michael L.; Shapiro, Beth (17 November 2015). "Life and extinction of megafauna in the ice-age Arctic". Proceedings of the National Academy of Sciences of the United States of America . 112 (46): 14301–14306. doi: 10.1073/pnas.1516573112 . ISSN   0027-8424. PMC   4655518 . PMID   26578776.
  187. MacDonald, G. M.; Beilman, D. W.; Kuzmin, Y. V.; Orlova, L. A.; Kremenetski, K. V.; Shapiro, B.; Wayne, R. K.; Van Valkenburgh, B. (12 June 2012). "Pattern of extinction of the woolly mammoth in Beringia". Nature Communications . 3 (1): 893. doi: 10.1038/ncomms1881 . ISSN   2041-1723.
  188. Stewart, Mathew; Carleton, W. Christopher; Groucutt, Huw S. (16 February 2021). "Climate change, not human population growth, correlates with Late Quaternary megafauna declines in North America". Nature Communications . 12 (1): 965. doi: 10.1038/s41467-021-21201-8 . ISSN   2041-1723.
  189. 1 2 Scott, E. (2010). "Extinctions, scenarios, and assumptions: Changes in latest Pleistocene large herbivore abundance and distribution in western North America". Quaternary International. 217 (1–2): 225–239. Bibcode:2010QuInt.217..225S. doi:10.1016/j.quaint.2009.11.003.
  190. Andersen, S. T. (1973). "The differential pollen productivity of trees and its significance for the interpretation of a pollen diagram from a forested region". In Birks, H.J.B.; West, R.G. (eds.). Quaternary plant ecology: the 14th symposium of the British Ecological society, University of Cambridge, 28–30 March 1972. Oxford: Blackwell Scientific. ISBN   0-632-09120-7.
  191. Ashworth, C.A. (1980). "Environmental implications of a beetle assemblage from the Gervais formation (Early Wisconsinian?), Minnesota". Quat. Res. 13 (2): 200–12. Bibcode:1980QuRes..13..200A. doi:10.1016/0033-5894(80)90029-0. S2CID   130045540.
  192. Bradley, R. S. (1985). Quaternary Paleoclimatology: Methods of Paleoclimatic Reconstruction . Winchester, MA: Allen & Unwin. ISBN   978-0-04-551068-9.
  193. Vartanyan SL, Arslanov KA, Tertychnaya TV, Chernov SB (1995). "Radiocarbon dating evidence for mammoths on Wrangel Island, Arctic Ocean, until 2000 BC". Radiocarbon. 37 (1): 1–6. Bibcode:1995Radcb..37....1V. doi: 10.1017/S0033822200014703 .
  194. Flereov, C.C. (1967). "On the origin of the mammalian fauna of Canada" . In Hopkins, D.M. (ed.). The Bering Land Bridge. Palo Alto: Stanford Univ. Press. pp.  271–80. ISBN   978-0-8047-0272-0.
  195. Frenzel, B. (1968). "The Pleistocene vegetation of northern Eurasia". Science. 161 (3842): 637–49. Bibcode:1968Sci...161..637F. doi:10.1126/science.161.3842.637. PMID   17801456.
  196. 1 2 McDonald, J. (1989). "The reordered North American selection regime and late Quaternary megafaunal extinctions". In Martin, P.S.; Klein, R.G. (eds.). Quaternary extinctions: A prehistoric revolution. Tucson, AZ: Univ. Arizona Press. pp. 354–404. ISBN   978-0-8165-1100-6.
  197. Birks, H.J.B.; West, R.G. (1973). "A Symposium of the British Ecological Society". Quaternary plant ecology: the 14th symposium of the British Ecological society, University of Cambridge, 28–30 March 1972. Oxford: Blackwell Scientific. ISBN   0-632-09120-7.
  198. McDonald, J. (1981). North American Bison: Their classification and evolution. Berkeley: Univ. Calif. Press. ISBN   978-0-520-04002-1.
  199. Burney, D. A. (1993). "Recent animal extinctions: recipes for disaster". American Scientist. 81 (6): 530–41. Bibcode:1993AmSci..81..530B.
  200. Vartanyan, S.L.; Garutt, V.E.; Sher, A.V. (1993). "Holocene dwarf mammoths from Wangel Island in the Siberian Arctic". Nature. 362 (6418): 337–40. Bibcode:1993Natur.362..337V. doi:10.1038/362337a0. PMID   29633990. S2CID   4249191.
  201. Pennycuick, C.J. (1979). "Energy costs of locomotion and the concept of "Foraging radius"". In Sinclair A.R.E.; Norton-Griffiths M. (eds.). Serengetti: Dynamics of an Ecosystem. Chicago: Univ. Chicago Press. pp. 164–85. ISBN   978-0-226-76029-2.
  202. Wing, L.D.; Buss, I.O. (1970). "Elephants and Forests". Wildl. Mong. (19).
  203. Owen-Smith, R.N. (1992). Megaherbivores: The influence of very large body size on ecology. Cambridge studies in ecology. Cambridge: Cambridge Univ. Press. ISBN   978-0-521-42637-4.
  204. Presenter: Willis, Paul; Producer: Bryce, Clay; Producer: Searle, Mike (17 August 2006). "Thylacoleo "The Beast of the Nullarbor"". Catalyst. Australian Broadcasting Commission. Archived from the original on 2020-09-30.
  205. Prideaux GJ, Long JA, Ayliffe LK, et al. (January 2007). "An arid-adapted middle Pleistocene vertebrate fauna from south-central Australia". Nature. 445 (7126): 422–425. Bibcode:2007Natur.445..422P. doi:10.1038/nature05471. PMID   17251978. S2CID   4429899.
  206. Roberts RG, Flannery TF, Ayliffe LK, et al. (June 2001). "New ages for the last Australian megafauna: continent-wide extinction about 46,000 years ago". Science. 292 (5523): 1888–92. Bibcode:2001Sci...292.1888R. doi:10.1126/science.1060264. PMID   11397939. S2CID   45643228.
  207. Fisher, Daniel C. (2009). "Paleobiology and Extinction of Proboscideans in the Great Lakes Region of North America". In Haynes, Gary (ed.). American Megafaunal Extinctions at the End of the Pleistocene. Vertebrate Paleobiology and Paleoanthropology. Springer. pp. 55–75. doi:10.1007/978-1-4020-8793-6_4. ISBN   978-1-4020-8792-9. S2CID   21544285.
  208. Whitney-Smith, Elin (2008). "The Evolution of an Ecosystem: Pleistocene Extinctions". In Minai, Ali A.; Bar-Yam, Yaneer (eds.). Unifying themes in complex systems IV proceedings of the Fourth International Conference on Complex Systems. Springer. ISBN   978-3-540-73849-7.
  209. Whitney-Smith, Elin (2012). "Creating the tiniest bison: A system dynamics model of ecological evolution". In Mendes, Ricardo Evandro (ed.). Ruminants: Anatomy, Behavior, and Diseases. Nova Biomedical. ISBN   9781620810644.
  210. 1 2 MacFee, Ross D. E.; Marx, Preston A. (1997). "Humans, hyperdisease and first-contact extinctions". In Goodman, S.; Patterson, B. D. (eds.). Natural Change and Human Impact in Madagascar. Washington DC: Smithsonian Press. pp. 169–217. ISBN   978-1-56098-683-6.
  211. MacFee, Ross D. E.; Marx, Preston A. (1998). "Lightning Strikes Twice: Blitzkrieg, Hyperdisease, and Global Explanations of the Late Quaternary Catastrophic Extinctions". American Museum of Natural History. Archived from the original on 2011-09-03.
  212. MacFee, Ross D. E.; Marx, Preston A. (1997). The 40,000-year Plague: Humans, Hyperdisease, and First-contact Extinctions. Natural Change and Human Impact in Madagascar. Washington DC: Smithsonian Institution Press. pp. 169–217.
  213. Lapointe, D. A.; Atkinson, C. T.; Samuel, M. D. (2012). "Ecology and conservation biology of avian malaria". Annals of the New York Academy of Sciences. 1249 (1): 211–26. Bibcode:2012NYASA1249..211L. doi:10.1111/j.1749-6632.2011.06431.x. PMID   22320256. S2CID   1885904.
  214. Fiedel, S. (2005). "Man's best friend: mammoth's worst enemy?". World Archaeology. 37: 11–35. doi:10.1080/0043824042000329540. S2CID   161477134.
  215. Nyström, Sofie; Hammarström, Per (7 January 2014). "Is the prevalent human prion protein 129M/V mutation a living fossil from a Paleolithic panzootic superprion pandemic?". Prion. 8 (1): 2–10. doi:10.4161/pri.27601. PMC   7030913 . PMID   24398570.
  216. Lyons, K.; Smith, F. A.; Wagner, P. J.; White, E. P.; Brown, J. H. (2004). "Was a 'hyperdisease' responsible for the late Pleistocene megafaunal extinction?" (PDF). Ecology Letters. 7 (9): 859–68. Bibcode:2004EcolL...7..859K. doi:10.1111/j.1461-0248.2004.00643.x.
  217. Kershaw, G.P. (1984). "Tundra plant communities of the Mackenzie mountains, Northwest Territories; floristic characteristics of long term surface disturbances". In Olson, R.; Hastings, R.; Geddes, F. (eds.). Northern Ecology and Resource Management: Memorial Essays honoring Don Gill. Edmonton, Canada: Univ. Alberta Press. pp.  239–311. ISBN   978-0-88864-047-5.
  218. Webber, P.J.; Miller, P.C.; Chapin, F.S. III; MacCown, B.H. (1980). "The vegetation: pattern and succession". In Brown, J.; Miller, P.C.; Tieszen, L.L.; Bunnell, F.L. (eds.). An Arctic ecosystem: the coastal tundra at Barrow, Alaska. US/IBP Synthesis. Stroudsburg, PA: Dowden Hutchinson & Ross. pp. 186–219. 12.
  219. Wolf, A.; Doughty, C. E.; Malhi, Y. (2013). "Lateral Diffusion of Nutrients by Mammalian Herbivores in Terrestrial Ecosystems". PLoS ONE . 8 (8): e71352. Bibcode:2013PLoSO...871352W. doi: 10.1371/journal.pone.0071352 . PMC   3739793 . PMID   23951141.
  220. Marshall, M. (2013-08-11). "Ecosystems still feel the pain of ancient extinctions". New Scientist . Retrieved 12 August 2013.
  221. 1 2 Doughty, C. E.; Wolf, A.; Malhi, Y. (2013). "The legacy of the Pleistocene megafauna extinctions on nutrient availability in Amazonia". Nature Geoscience . 6 (9): 761–764. Bibcode:2013NatGe...6..761D. doi:10.1038/ngeo1895.[ permanent dead link ]
  222. 1 2 Doughty, C. E.; Wolf, A.; Field, C. B. (2010). "Biophysical feedbacks between the Pleistocene megafauna extinction and climate: The first human‐induced global warming?". Geophysical Research Letters. 37 (15): n/a. Bibcode:2010GeoRL..3715703D. doi: 10.1029/2010GL043985 . S2CID   54849882.
  223. Vignieri, S. (25 July 2014). "Vanishing fauna (Special issue)". Science. 345 (6195): 392–412. Bibcode:2014Sci...345..392V. doi: 10.1126/science.345.6195.392 . PMID   25061199.
  224. Sandom, Christopher; Faurby, Søren; Sandel, Brody; Svenning, Jens-Christian (4 June 2014). "Global late Quaternary megafauna extinctions linked to humans, not climate change". Proceedings of the Royal Society B . 281 (1787): 20133254. doi:10.1098/rspb.2013.3254. PMC   4071532 . PMID   24898370.
  225. 1 2 3 Wilkinson, D. M.; Nisbet, E. G.; Ruxton, G. D. (2012). "Could methane produced by sauropod dinosaurs have helped drive Mesozoic climate warmth?". Current Biology . 22 (9): R292–R293. doi: 10.1016/j.cub.2012.03.042 . PMID   22575462.
  226. "Dinosaur gases 'warmed the Earth'". BBC Nature News. 7 May 2012. Archived from the original on 1 December 2015. Retrieved 8 May 2012.
  227. Kelliher, F. M.; Clark, H. (15 March 2010). "Methane emissions from bison—An historic herd estimate for the North American Great Plains". Agricultural and Forest Meteorology . 150 (3): 473–577. Bibcode:2010AgFM..150..473K. doi:10.1016/j.agrformet.2009.11.019.
  228. 1 2 Smith, F. A.; Elliot, S. M.; Lyons, S. K. (23 May 2010). "Methane emissions from extinct megafauna". Nature Geoscience . 3 (6): 374–375. Bibcode:2010NatGe...3..374S. doi:10.1038/ngeo877. S2CID   128832000.

Hyperdisease hypothesis

Second-order predation