Great American Interchange

The Great American Biotic Interchange (commonly abbreviated as GABI), also known as the Great American Interchange and the Great American Faunal Interchange, was an important late Cenozoic paleozoogeographic biotic interchange event in which land and freshwater fauna migrated from North America to South America via Central America and vice versa, as the volcanic Isthmus of Panama rose up from the sea floor and bridged the formerly separated continents. Although earlier dispersals had occurred, probably over water, the migration accelerated dramatically about 2.7 million years (Ma) ago during the Piacenzian age.[1] It resulted in the joining of the Neotropic (roughly South American) and Nearctic (roughly North American) biogeographic realms definitively to form the Americas. The interchange is visible from observation of both biostratigraphy and nature (neontology). Its most dramatic effect is on the zoogeography of mammals, but it also gave an opportunity for reptiles, amphibians, arthropods, weak-flying or flightless birds, and even freshwater fish to migrate. Coastal and marine biota, however, were affected in the opposite manner; the formation of the Central American Isthmus caused what has been termed the Great American Schism, with significant diversification and extinction occurring as a result of the isolation of the Caribbean from the Pacific.[2]

Examples of migrant species in both Americas. Olive green silhouettes denote North American species with South American ancestors; blue silhouettes denote South American species with North American ancestors.

The occurrence of the interchange was first discussed in 1876 by the "father of biogeography", Alfred Russel Wallace.[3][4] Wallace had spent five years exploring and collecting specimens in the Amazon basin. Others who made significant contributions to understanding the event in the century that followed include Florentino Ameghino, W. D. Matthew, W. B. Scott, Bryan Patterson, George Gaylord Simpson and S. David Webb.[5] The Pliocene timing of the formation of the connection between North and South America was discussed in 1910 by Henry Fairfield Osborn.[6]

Analogous interchanges occurred earlier in the Cenozoic, when the formerly isolated land masses of India and Africa made contact with Eurasia about 56 and 30 Ma ago, respectively.[7][8][9][10][11][12][13][14][15][16][17][excessive citations]

Before the interchange edit

Isolation of South America edit

 
Thylacosmilus, a sparassodont

After the late Mesozoic breakup of Gondwana, South America spent most of the Cenozoic era as an island continent whose "splendid isolation" allowed its fauna to evolve into many forms found nowhere else on Earth, most of which are now extinct.[18] Its endemic mammals initially consisted primarily of metatherians (marsupials and sparassodonts), xenarthrans, and a diverse group of native ungulates known as the Meridiungulata: notoungulates (the "southern ungulates"), litopterns, astrapotheres, pyrotheres and xenungulates.[n 1][n 2] A few non-therian mammals – monotremes, gondwanatheres, dryolestids and possibly cimolodont multituberculates – were also present in the Paleocene; while none of these diversified significantly and most lineages did not survive long, forms like Necrolestes and Patagonia remained as recently as the Miocene.[25]

 
The monito del monte, Dromiciops gliroides, South America's only australidelphian marsupial
 
Astrapotherium magnum

Marsupials appear to have traveled via Gondwanan land connections from South America through Antarctica to Australia in the late Cretaceous or early Tertiary.[26][n 3] One living South American marsupial, the monito del monte, has been shown to be more closely related to Australian marsupials than to other South American marsupials (Ameridelphia); however, it is the most basal australidelphian,[n 4] meaning that this superorder arose in South America and then dispersed to Australia after the monito del monte split off.[26] Monotrematum, a 61-Ma-old platypus-like monotreme fossil from Patagonia, may represent an Australian immigrant.[27][28] Paleognath birds (ratites and South American tinamous) may have made a similar migration around the same time to Australia and New Zealand.[29][30] Other taxa that may have dispersed by the same route (if not by flying or oceanic dispersal) are parrots, chelid turtles, and the extinct meiolaniid turtles.

Marsupials remaining in South America included didelphimorphs (opossums), paucituberculatans (shrew opossums) and microbiotheres (monitos del monte). Larger predatory relatives of these also existed, such as the borhyaenids and the saber-toothed Thylacosmilus; these were sparassodont metatherians, which are no longer considered to be true marsupials.[31] As the large carnivorous metatherians declined, and before the arrival of most types of carnivorans, predatory opossums such as Thylophorops temporarily attained larger size (about 7 kg).

Metatherians and a few xenarthran armadillos, such as Macroeuphractus, were the only South American mammals to specialize as carnivores; their relative inefficiency created openings for nonmammalian predators to play more prominent roles than usual (similar to the situation in Australia). Sparassodonts and giant opossums shared the ecological niches for large predators with fearsome flightless "terror birds" (phorusrhacids), whose closest living relatives are the seriemas.[32][33] North America also had large terrestrial predatory birds during the early Cenozoic (the related bathornithids), but they died out before the GABI in the Early Miocene, about 20 million years ago. Through the skies over late Miocene South America (6 Ma ago) soared one of the largest flying birds known, Argentavis, a teratorn that had a wing span of 6 m or more, and which may have subsisted in part on the leftovers of Thylacosmilus kills.[34] Terrestrial sebecid (metasuchian) crocodyliforms with ziphodont teeth[n 5] were also present at least through the middle Miocene[35][36][37][38] and maybe to the Miocene-Pliocene boundary.[39] Some of South America's aquatic crocodilians, such as Gryposuchus, Mourasuchus and Purussaurus, reached monstrous sizes, with lengths up to 12 m (comparable to the largest Mesozoic crocodyliforms). They shared their habitat with one of the largest turtles of all time, the 3.3 m (11 ft) Stupendemys.

 
The giant anteater, Myrmecophaga tridactyla, the largest living descendant of South America's early Cenozoic mammalian fauna

Xenarthrans are a curious group of mammals that developed morphological adaptations for specialized diets very early in their history.[40] In addition to those extant today (armadillos, anteaters, and tree sloths), a great diversity of larger types was present, including pampatheres, the ankylosaur-like glyptodonts, predatory euphractines, various ground sloths, some of which reached the size of elephants (e.g. Megatherium), and even semiaquatic to aquatic marine sloths.[41][42]

 
The litoptern Macrauchenia

The notoungulates and litopterns had many strange forms, such as Macrauchenia, a camel-like litoptern with a small proboscis. They also produced a number of familiar-looking body types that represent examples of parallel or convergent evolution: one-toed Thoatherium had legs like those of a horse, Pachyrukhos resembled a rabbit, Homalodotherium was a semibipedal, clawed browser like a chalicothere, and horned Trigodon looked like a rhinoceros. Both groups started evolving in the Lower Paleocene, possibly from condylarth stock, diversified, dwindled before the great interchange, and went extinct at the end of the Pleistocene. The pyrotheres and astrapotheres were also strange, but were less diverse and disappeared earlier, well before the interchange.

The North American fauna was a typical boreoeutherian one, supplemented with Afrotherian proboscids.

Pre-interchange oceanic dispersals edit

 
Capybara, Hydrochoerus hydrochaeris
 
Emperor tamarin, Saguinus imperator

The invasions of South America started about 40 Ma ago (middle Eocene), when caviomorph rodents arrived in South America.[43][44][45] Their subsequent vigorous diversification displaced some of South America's small marsupials and gave rise to – among others – capybaras, chinchillas, viscachas, and New World porcupines. The independent development of spines by New and Old World porcupines is another example of parallel evolution. This invasion most likely came from Africa.[46][47] The crossing from West Africa to the northeast corner of Brazil was much shorter then, due to continental drift, and may have been aided by island hopping (e.g. via St. Paul's Rocks, if they were an inhabitable island at the time) and westward oceanic currents.[48] Crossings of the ocean were accomplished when at least one fertilised female (more commonly a group of animals) accidentally floated over on driftwood or mangrove rafts. Hutias (Capromyidae) would subsequently colonize the West Indies as far as the Bahamas,[49][50] reaching the Greater Antilles by the early Oligocene.[51] Over time, some caviomorph rodents evolved into larger forms that competed with some of the native South American ungulates, which may have contributed to the gradual loss of diversity suffered by the latter after the early Oligocene.[18] By the Pliocene, some caviomorphs (e.g., Josephoartigasia monesi) attained sizes on the order of 500 kg (1,100 lb) or larger.[52]

Later (by 36 Ma ago),[53] primates followed, again from Africa in a fashion similar to that of the rodents.[43] Primates capable of migrating had to be small. Like caviomorph rodents, South American monkeys are believed to be a clade (i.e., monophyletic). However, although they would have had little effective competition, all extant New World monkeys appear to derive from a radiation that occurred long afterwards, in the Early Miocene about 18 Ma ago.[43] Subsequent to this, monkeys apparently most closely related to titis island-hopped to Cuba, Hispaniola, and Jamaica. Additionally, a find of seven 21-Ma-old apparent cebid teeth in Panama suggests that South American monkeys had dispersed across the seaway separating Central and South America by that early date. However, all extant Central American monkeys are believed to be descended from much later migrants, and there is as yet no evidence that these early Central American cebids established an extensive or long-lasting population, perhaps due to a shortage of suitable rainforest habitat at the time.[54][55]

Fossil evidence presented in 2020 indicates a second lineage of African monkeys also rafted to and at least briefly colonized South America. Ucayalipithecus remains dating from the Early Oligocene of Amazonian Peru are, by morphological analysis, deeply nested within the family Parapithecidae of the Afro-Arabian radiation of parapithecoid simians, with dental features markedly different from those of platyrrhines. The Old World members of this group are thought to have become extinct by the Late Oligocene. Qatrania wingi of lower Oligocene Fayum deposits is considered the closest known relative of Ucayalipithecus.[56][57]

Remarkably, the descendants of those few bedraggled "waifs" that crawled ashore from their rafts of African flotsam in the Eocene now constitute more than twice as many of South America's species as the descendants of all the flightless mammals previously resident on the continent (372 caviomorph and monkey species versus 136 marsupial and xenarthran species).[n 6]

Many of South America's bats may have arrived from Africa during roughly the same period, possibly with the aid of intervening islands, although by flying rather than floating. Noctilionoid bats ancestral to those in the neotropical families Furipteridae, Mormoopidae, Noctilionidae, Phyllostomidae, and Thyropteridae are thought to have reached South America from Africa in the Eocene,[59] possibly via Antarctica.[60] Similarly, free-tailed bats (Molossidae) may have reached South America from Africa in as many as five dispersals, starting in the Eocene.[59] Emballonurids may have also reached South America from Africa about 30 Ma ago, based on molecular evidence.[59][61] Vespertilionids may have arrived in five dispersals from North America and one from Africa.[59] Natalids are thought to have arrived during the Pliocene from North America via the Caribbean.[59]

 
Red-footed tortoise, Chelonoidis carbonaria

Tortoises also arrived in South America in the Oligocene. They were long thought to have come from North America, but a recent comparative genetic analysis concludes that the South American genus Chelonoidis (formerly part of Geochelone) is actually most closely related to African hingeback tortoises.[n 7][62] Tortoises are aided in oceanic dispersal by their ability to float with their heads up, and to survive up to six months without food or water.[62] South American tortoises then went on to colonize the West Indies[63] and Galápagos Islands (the Galápagos tortoise). A number of clades of American geckos seem to have rafted over from Africa during both the Paleogene and Neogene.[64] Skinks of the related genera Mabuya and Trachylepis apparently dispersed across the Atlantic from Africa to South America and Fernando de Noronha, respectively, during the last 9 Ma.[65] Surprisingly, South America's burrowing amphisbaenians[66] and blind snakes[67] also appear to have rafted from Africa, as does the hoatzin, a weak-flying bird of South American rainforests.[68]

 
Megalonyx wheatleyi

The earliest traditionally recognized mammalian arrival from North America was a procyonid that island-hopped from Central America before the Isthmus of Panama land bridge formed, around 7.3 Ma ago.[69] This was South America's first eutherian carnivore. South American procyonids then diversified into forms now extinct (e.g. the "dog-coati" Cyonasua, which evolved into the bear-like Chapalmalania). However, all extant procyonid genera appear to have originated in North America.[70] The first South American procyonids may have contributed to the extinction of sebecid crocodilians by eating their eggs, but this view has not been universally viewed as plausible.[n 8][38] The procyonids were followed to South America by rafting or island-hopping hog-nosed skunks[71] and sigmodontine rodents.[72][73][74][75] The oryzomyine tribe of sigmodontine rodents went on to colonize the Lesser Antilles to Anguilla.

One group has proposed that a number of large Neartic herbivores actually reached South America as early as 9–10 Ma ago, in the late Miocene, via an early incomplete land bridge. These claims, based on fossils recovered from rivers in southwestern Peru, have been viewed with caution by other investigators, due to the lack of corroborating finds from other sites and the fact that almost all of the specimens in question have been collected as float in rivers without little to no stratigraphic control.[76] These taxa are a gomphothere (Amahuacatherium),[77][78] peccaries (Sylvochoerus and Waldochoerus),[79] tapirs and Surameryx, a palaeomerycid (from a family probably ancestral to cervids).[80] The identification of Amahuacatherium and the dating of its site is controversial; it is regarded by a number of investigators as a misinterpreted fossil of a different gomphothere, Notiomastodon, and biostratigraphy dates the site to the Pleistocene.[81][82][83] The early date proposed for Surameryx has also been met with skepticism.[84]

Megalonychid and mylodontid ground sloths island-hopped to North America by 9 Ma ago.[72] A basal group of sloths[85] had colonized the Antilles previously, by the early Miocene.[86] In contrast, megatheriid and nothrotheriid ground sloths did not migrate north until the formation of the isthmus. Terror birds may have also island-hopped to North America as early as 5 Ma ago.[87]

The Caribbean Islands were populated primarily by species from South America, due to the prevailing direction of oceanic currents, rather than to a competition between North and South American forms.[49][50] Except in the case of Jamaica, oryzomyine rodents of North American origin were able to enter the region only after invading South America.

Effects and aftermath edit

The formation of the Isthmus of Panama led to the last and most conspicuous wave, the Great American Biotic Interchange (GABI), starting around 2.7 Ma ago. This included the immigration into South America of North American ungulates (including camelids, tapirs, deer and horses), proboscids (gomphotheres), carnivorans (including felids such as cougars, jaguars and saber-toothed cats, canids, mustelids, procyonids and bears) and a number of types of rodents.[n 9] The larger members of the reverse migration were ground sloths, terror birds, glyptodonts, pampatheres, capybaras, and the notoungulate Mixotoxodon (the only South American ungulate known to have invaded Central America).

 
Titanis walleri, the only known North American terror bird

In general, the initial net migration was symmetrical. Later on, however, the Neotropic species proved far less successful than the Nearctic. This difference in fortunes was manifested in several ways. Northwardly migrating animals often were not able to compete for resources as well as the North American species already occupying the same ecological niches; those that did become established were not able to diversify much, and in some cases did not survive for long.[88] Southwardly migrating Nearctic species established themselves in larger numbers and diversified considerably more,[88] and are thought to have caused the extinction of a large proportion of the South American fauna.[71][89][90] (No extinctions in North America are plainly linked to South American immigrants.[n 10]) Native South American ungulates did poorly, with only a handful of genera withstanding the northern onslaught. (Several of the largest forms, macraucheniids and toxodontids, have long been recognized to have survived to the end of the Pleistocene. Recent fossil finds indicate that one species of the horse-like proterotheriid litopterns did, as well.[92] The notoungulate mesotheriids and hegetotheriids also managed to hold on at least part way through the Pleistocene.)[A] South America's small marsupials, though, survived in large numbers, while the primitive-looking xenarthrans proved to be surprisingly competitive and became the most successful invaders of North America. The African immigrants, the caviomorph rodents and platyrrhine monkeys, were less impacted by the interchange than most of South America's 'old-timers', although the caviomorphs suffered a significant loss of diversity,[n 11][n 12] including the elimination of the largest forms (e.g. the dinomyids). With the exception of the North American porcupine and several extinct porcupines and capybaras, however, they did not migrate past Central America.[n 13]

Due in large part to the continued success of the xenarthrans, one area of South American ecospace the Nearctic invaders were unable to dominate was the niches for megaherbivores.[94] Before 12,000 years ago, South America was home to about 25 species of herbivores weighing more than 1,000 kg (2,200 lb), consisting of Neotropic ground sloths, glyptodonts, and toxodontids, as well as gomphotheres and camelids of Nearctic origin.[n 14] Native South American forms made up about 75% of these species. However, none of these megaherbivores has survived.

 
The Virginia opossum, Didelphis virginiana, the only marsupial in temperate North America

Armadillos, opossums and porcupines are present in North America today because of the Great American Interchange. Opossums and porcupines were among the most successful northward migrants, reaching as far as Canada and Alaska, respectively. Most major groups of xenarthrans were present in North America until the end-Pleistocene Quaternary extinction event (as a result of at least eight successful invasions of temperate North America, and at least six more invasions of Central America only). Among the megafauna, ground sloths were notably successful emigrants; four different lineages invaded North America. A megalonychid representative, Megalonyx, spread as far north as the Yukon[96] and Alaska,[97] and might well have invaded Eurasia had a suitable habitat corridor across Beringia been present.

Generally speaking, however, the dispersal and subsequent explosive adaptive radiation of sigmodontine rodents throughout South America (leading to over 80 currently recognized genera) was vastly more successful (both spatially and by number of species) than any northward migration of South American mammals. Other examples of North American mammal groups that diversified conspicuously in South America include canids and cervids, both of which currently have three or four genera in North America, two or three in Central America, and six in South America.[n 15][n 16] Although members of Canis (specifically, coyotes) currently range only as far south as Panama,[n 17] South America still has more extant genera of canids than any other continent.[n 15]

The effect of formation of the isthmus on the marine biota of the area was the inverse of its effect on terrestrial organisms, a development that has been termed the "Great American Schism". The connection between the east Pacific Ocean and the Caribbean (the Central American Seaway) was severed, setting now-separated populations on divergent evolutionary paths.[2] Caribbean species also had to adapt to an environment of lower productivity after the inflow of nutrient-rich water of deep Pacific origin was blocked.[101] The Pacific coast of South America cooled as the input of warm water from the Caribbean was cut off. This trend is thought to have caused the extinction of the marine sloths of the area.[102]

Disappearance of native South American predators edit

 
Arctotherium bonariense, a South American short-faced bear

During the last 7 Ma, South America's terrestrial predator guild has changed from one composed almost entirely of nonplacental mammals (metatherians), birds, and reptiles to one dominated by immigrant placental carnivorans (with a few small marsupial and avian predators like didelphine opossums and seriemas). It was originally thought that the native South American predator guild, including sparassodonts, carnivorous opossums like Thylophorops and Hyperdidelphys, armadillos such as Macroeuphractus, terror birds, and teratorns, as well as early-arriving immigrant Cyonasua-group procyonids, were driven to extinction during the GABI by competitive exclusion from immigrating placental carnivorans, and that this turnover was abrupt.[103][104] However, the turnover of South America's predator guild was more complex, with competition only playing a limited role.

In the case of sparassodonts and carnivorans, which has been the most heavily studied, little evidence shows that sparassodonts even encountered their hypothesized placental competitors.[105][106][107] Many supposed Pliocene records of South American carnivorans have turned out to be misidentified or misdated.[108][105] Sparassodonts appear to have been declining in diversity since the middle Miocene, with many of the niches once occupied by small sparassodonts being increasingly occupied by carnivorous opossums,[109][110][111][112][113] which reached sizes of up to roughly 8 kg (~17 lbs).[110] Whether sparassodonts competed with carnivorous opossums or whether opossums began occupying sparassodont niches through passive replacement is still debated.[113][112][111][110] Borhyaenids last occur in the late Miocene, about 4 Ma before the first appearance of canids or felids in South America.[106] Thylacosmilids last occur about 3 Ma ago and appear to be rarer at pre-GABI Pliocene sites than Miocene ones.[105]

In general, sparassodonts appear to have been mostly or entirely extinct by the time most nonprocyonid carnivorans arrived, with little overlap between the groups. Purported ecological counterparts between pairs of analogous groups (thylacosmilids and saber-toothed cats, borhyaenids and felids, hathliacynids and weasels) neither overlap in time nor abruptly replace one another in the fossil record.[103][106] Procyonids dispersed to South America by at least 7 Ma ago, and had achieved a modest endemic radiation by the time other carnivorans arrived (Cyonasua-group procyonids). However, procyonids do not appear to have competed with sparassodonts, the procyonids being large omnivores and sparassodonts being primarily hypercarnivorous.[114] Other groups of carnivorans did not arrive in South America until much later. Dogs and weasels appear in South America about 2.9 Ma ago, but do not become abundant or diverse until the early Pleistocene.[105] Bears, cats, and skunks do not appear in South America until the early Pleistocene (about 1 Ma ago or slightly earlier).[105] Otters and other groups of procyonids (i.e., coatis, raccoons) have been suggested to have dispersed to South America in the Miocene based on genetic data, but no remains of these animals have been found even at heavily sampled northern South American fossil sites such as La Venta (Colombia), which is only 600 km (370 mi) from the Isthmus of Panama.[115][114][116][117]

Other groups of native South American predators have not been studied in as much depth. Terror birds have often been suggested to have been driven to extinction by placental carnivorans, though this hypothesis has not been investigated in detail.[118][119] Titanis dispersed from South America to North America against the main wave of carnivoran migrations, being the only large native South American carnivore to accomplish this.[119] However, it only managed to colonize a small part of North America for a limited time, failing to diversify and going extinct in the early Pleistocene (1.8 Ma ago); the modest scale of its success has been suggested to be due to competition with placental carnivorans.[120] Terror birds also decline in diversity after about 3 Ma ago.[105] At least one genus of relatively small terror birds, Psilopterus, appears to have survived to as recently as about 96,000 years ago.[121][122]

The native carnivore guild appears to have collapsed completely roughly 3 Ma ago (including the extinction of the last sparassodonts), not correlated with the arrival of carnivorans in South America, with terrestrial carnivore diversity being low thereafter.[105][123] This has been suggested to have opened up ecological niches and allowed carnivorans to establish themselves in South America due to low competition.[114][124][125] A meteor impact 3.3 million years ago in southern South America has been suggested as a possible cause of this turnover, but this is still controversial.[126][123] A similar pattern occurs in the crocodilian fauna, where modern crocodiles (Crocodylus) dispersed to South America during the Pliocene and became the dominant member of crocodilian communities after the late Miocene extinction of the previously dominant large native crocodilians such as the giant caiman Purussaurus and giant gharial Gryposuchus, which is thought to be related to the loss of wetlands habitat across northern South America.[127][128]

Whether this revised scenario with a reduced role for competitive exclusion applies to other groups of South American mammals such as notoungulates and litopterns is unclear, though some authors have pointed out a protracted decline in South American native ungulate diversity since the middle Miocene.[129] Regardless of how this turnover happened, it is clear that carnivorans benefitted from it. Several groups of carnivorans such as dogs and cats underwent an adaptive radiation in South America after dispersing there, and the greatest modern diversity of canids in the world is in South America.[99]

Reasons for success or failure edit

 
A north–south climatic asymmetry in the Americas: Tropical climate zones, which are warm year-round and moist at least part of the year (blue zones Af, Am and Aw), cover much of South America and nearly all of Central America, but very little of the rest of North America.

The eventual triumph of the Nearctic migrants was ultimately based on geography, which played into the hands of the northern invaders in two crucial respects. The first was a matter of climate. Any species that reached Panama from either direction obviously had to be able to tolerate moist tropical conditions. Those migrating southward would then be able to occupy much of South America without encountering climates that were markedly different. However, northward migrants would have encountered drier or cooler conditions by the time they reached the vicinity of the Trans-Mexican Volcanic Belt. The challenge this climatic asymmetry (see map on right) presented was particularly acute for Neotropic species specialized for tropical rainforest environments, which had little prospect of penetrating beyond Central America. As a result, Central America currently has 41 mammal species of Neotropical origin,[n 18] compared to only three for temperate North America. However, species of South American origin (marsupials, xenarthrans, caviomorph rodents, and monkeys) still comprise only 21% of species from nonflying, nonmarine mammal groups in Central America, while North American invaders constitute 49% of species from such groups in South America. Thus, climate alone cannot fully account for the greater success of species of Nearctic origin during the interchange.

 
Land areas over which ancestors of Neotropic (green) and Nearctic (red) species could wander via two-way migrations during the latter part of the Cenozoic prior to the interchange: The smaller area available for Neotropic species to evolve in tended to put them at a competitive disadvantage.

The second and more important advantage geography gave to the northerners is related to the land area in which their ancestors evolved. During the Cenozoic, North America was periodically connected to Eurasia via Beringia, allowing repeated migrations back and forth to unite the faunas of the two continents.[n 19] Eurasia was connected in turn to Africa, which contributed further to the species that made their way to North America.[n 20] South America, though, was connected only to Antarctica and Australia, two much smaller and less hospitable continents, and only in the early Cenozoic. Moreover, this land connection does not seem to have carried much traffic (apparently no mammals other than marsupials and perhaps a few monotremes ever migrated by this route), particularly in the direction of South America. This means that Northern Hemisphere species arose within a land area roughly six times greater than was available to South American species. North American species were thus products of a larger and more competitive arena,[n 21][88][130][131] where evolution would have proceeded more rapidly. They tended to be more efficient and brainier,[n 22][n 23] generally able to outrun and outwit their South American counterparts, who were products of an evolutionary backwater. In the cases of ungulates and their predators, South American forms were replaced wholesale by the invaders, possibly a result of these advantages.

The greater eventual success of South America's African immigrants compared to its native early Cenozoic mammal fauna is another example of this phenomenon, since the former evolved over a greater land area; their ancestors migrated from Eurasia to Africa, two significantly larger continents, before finding their way to South America.[58]

Against this backdrop, the ability of South America's xenarthrans to compete effectively against the northerners represents a special case. The explanation for the xenarthrans' success lies in part in their idiosyncratic approach to defending against predation, based on possession of body armor or formidable claws. The xenarthrans did not need to be fleet-footed or quick-witted to survive. Such a strategy may have been forced on them by their low metabolic rate (the lowest among the therians).[139][140] Their low metabolic rate may in turn have been advantageous in allowing them to subsist on less abundant[141] or less nutritious food sources. Unfortunately, the defensive adaptations of the large xenarthrans would have offered little protection against humans armed with spears and other projectiles.

Late Pleistocene extinctions edit

 
The North American porcupine, Erethizon dorsatum, the largest surviving Neotropic migrant to temperate North America

At the end of the Pleistocene epoch, about 12,000 years ago, three dramatic developments occurred in the Americas at roughly the same time (geologically speaking). Paleoindians invaded and occupied the New World (although humans may have been living in the Americas, including what is now the southern US and Chile, more than 15,000 years ago[142]), the last glacial period came to an end, and a large fraction of the megafauna of both North and South America went extinct. This wave of extinctions swept off the face of the Earth many of the successful participants of the GABI, as well as other species that had not migrated.

All the pampatheres, glyptodonts, ground sloths, equids, proboscideans,[143][144][83] giant short-faced bears, dire wolves, and machairodont species of both continents disappeared. The last of the South and Central American notoungulates and litopterns died out, as well as North America's giant beavers, lions, dholes, cheetahs, and many of its antilocaprid, bovid, cervid, tapirid and tayassuid ungulates. Some groups disappeared over most or all of their original range, but survived in their adopted homes, e.g. South American tapirs, camelids, and tremarctine bears (cougars and jaguars may have been temporarily reduced to South American refugia also). Others, such as capybaras, survived in their original range, but died out in areas to which they had migrated. Notably, this extinction pulse eliminated all Neotropic migrants to North America larger than about 15 kg (the size of a big porcupine), and all native South American mammals larger than about 65 kg (the size of a big capybara or giant anteater). In contrast, the largest surviving native North American mammal, the wood bison, can exceed 900 kg (2,000 lb), and the largest surviving Nearctic migrant to South America, Baird's tapir, can reach 400 kg (880 lb).

 
Paleo-Indians and †Glyptodon
 
Baird's tapir, Tapirus bairdii, the largest surviving Nearctic migrant to South America

The near-simultaneity of the megafaunal extinctions with the glacial retreat and the peopling of the Americas has led to proposals that both climate change and human hunting played a role.[94] Although the subject is contentious,[145][146][147][148][149] a number of considerations suggest that human activities were pivotal.[95][150] The extinctions did not occur selectively in the climatic zones that would have been most affected by the warming trend, and no plausible general climate-based megafauna-killing mechanism could explain the continent-wide extinctions. The climate change took place worldwide, but had little effect on the megafauna in Africa and southern Asia, where megafaunal species had coevolved with humans. Numerous very similar glacial retreats had occurred previously within the ice age of the last several million years without ever producing comparable waves of extinction in the Americas or anywhere else.

Similar megafaunal extinctions have occurred on other recently populated land masses (e.g. Australia,[151][152] Japan,[153] Madagascar,[154] New Zealand,[155] and many smaller islands around the world, such as Cyprus,[156] Crete, Tilos and New Caledonia[157]) at different times that correspond closely to the first arrival of humans at each location. These extinction pulses invariably swept rapidly over the full extent of a contiguous land mass, regardless of whether it was an island or a hemisphere-spanning set of connected continents. This was true despite the fact that all the larger land masses involved (as well as many of the smaller ones) contained multiple climatic zones that would have been affected differently by any climate changes occurring at the time. However, on sizable islands far enough offshore from newly occupied territory to escape immediate human colonization, megafaunal species sometimes survived for many thousands of years after they or related species became extinct on the mainland; examples include giant kangaroos in Tasmania,[158][159] giant Chelonoidis tortoises of the Galápagos Islands (formerly also of South America[94]), giant Dipsochelys tortoises of the Seychelles (formerly also of Madagascar), giant meiolaniid turtles on Lord Howe Island, New Caledonia and Vanuatu (previously also of Australia),[160][n 24] ground sloths on the Antilles,[163][164] Steller's sea cows off the Commander Islands[165] and woolly mammoths on Wrangel Island[166] and Saint Paul Island.[167]

The glacial retreat may have played a primarily indirect role in the extinctions in the Americas by simply facilitating the movement of humans southeastward from Beringia to North America. The reason that a number of groups went extinct in North America but lived on in South America (while no examples of the opposite pattern are known) appears to be that the dense rainforest of the Amazon basin and the high peaks of the Andes provided environments that afforded a degree of protection from human predation.[168][n 25][n 26]

List of North American species of South American origin edit

Distributions beyond Mexico edit

Extant or extinct (†) North American taxa whose ancestors migrated out of South America and reached the modern territory of the contiguous United States:[n 27]

Fish

Amphibians

Birds

Mammals

Distributions restricted to Mexico edit

Extant or extinct (†) North American taxa whose ancestors migrated out of South America, but failed to reach the contiguous United States and were confined to Mexico and Central America:[n 27][n 29]

Invertebrates

Fish

Amphibians

Reptiles

Birds

Mammals

List of South American species of North American origin edit

Extant or extinct (†) South American taxa whose ancestors migrated out of North America:[n 27]

Amphibians

Reptiles

Birds

Mammals

See also edit

Notes edit

  1. ^ During the Eocene, astrapotheres[19] and litopterns[20][21] were also present in Antarctica.
  2. ^ Sequencing of collagen from fossils of one recently extinct species each of notoungulates and litopterns has indicated that these orders comprise a sister group to the perissodactyls.[22][23] Mitochondrial DNA obtained from Macrauchenia corroborates this and gives an estimated divergence date of 66 Ma ago.[24]
  3. ^ Once in Australia, facing less competition, marsupials diversified to fill a much larger array of niches than in South America, where they were largely carnivorous.
  4. ^ It is the sister group to a clade containing all other extant australidelphians (roughly 238 species).
  5. ^ Ziphodont (lateromedially compressed, recurved and serrated) teeth tend to arise in terrestrial crocodilians because, unlike their aquatic cousins, they are unable to dispatch their prey by simply holding them underwater and drowning them; they thus need cutting teeth with which to slice open their victims.
  6. ^ It is also notable that both simians (ancestral to monkeys) and hystricognath rodents (ancestral to caviomorphs) are believed to have arrived in Africa by rafting from Eurasia about 40 Ma ago.[58]
  7. ^ North American gopher tortoises are most closely related to the Asian genus Manouria.
  8. ^ An alternative explanation blames climatic and physiographic changes associated with the uplift of the Andes.[38]
  9. ^ Of the 6 families of North American rodents that did not originate in South America, only beavers and mountain beavers failed to migrate to South America. (However, human-introduced beavers have become serious pests in Tierra del Fuego.)
  10. ^ In this connection, however, chalicotheres, clawed perissodactyl herbivores ecologically similar to ground sloths, died out in North America in the Miocene about 9 Ma ago, while they survived to the early Pleistocene in Asia and Africa.[91]
  11. ^ Simpson, 1950, p. 382[93]
  12. ^ Marshall, 1988, p. 386[5]
  13. ^ Of the 11 extant families of South American caviomorph rodents, five are present in Central America; only two of these, Erethizontidae and Caviidae, ever reached North America. (The nutria/coypu has been introduced to a number of North American locales.)
  14. ^ P. S. Martin (2005), pp. 30–37, 119.[95] The figure of 25 South American megaherbivore species breaks down as follows: four gomphotheres, two camelids, nine ground sloths, five glyptodonts, and five toxodontids. This can be compared to Africa's present and recent total of six megaherbivores: one giraffe, one hippo, two rhinos, and two elephants (considering the African forest elephant as a separate species).
  15. ^ a b The extant canid and cervid genera by continent are as follows: Canid genera by continent Cervid genera by continent
  16. ^ Including extinct genera, South America has hosted nine genera of cervids, eight genera of mustelids, and 10 genera of canids. However, some of this diversity of South American forms apparently arose in North or Central America prior to the interchange.[88] Significant disagreement exists in the literature concerning how much of the diversification of South America's canids occurred prior to the invasions. A number of studies concur that the grouping of endemic South American canids (excluding Urocyon and Canis, although sometimes transferring C. gezi to the South American group[98]) is a clade.[98][99][100] However, different authors conclude that members of this clade reached South America in at least two,[99] three to four,[98] or six[100] invasions from North America.
  17. ^ Canis, e.g. Canis dirus, was present in South America until the end of the Pleistocene.
  18. ^ a b c d The Central American species of South American origin (opossums, xenarthrans, caviomorph rodents and platyrrhine monkeys) are as follows: Central American opossum species Central American xenarthran species Central American caviomorph rodent species Central American platyrrhine monkey species
  19. ^ During the Miocene alone, between about 23 and 5 Ma ago, 11 episodes of invasions of North America from Eurasia have been recognized, bringing a total of 81 new genera into North America.[88]
  20. ^ The combination of Africa, Eurasia and North America was termed the "World Continent" by George Gaylord Simpson.[93]
  21. ^ Simpson, 1950, p. 368[93]
  22. ^ According to data on the EQ (encephalization quotient, a measure of the brain to body size ratio adjusted for the expected effect of differences in body size) of fossil ungulates compiled by H. Jerison,[132] North American ungulates showed a trend towards greater EQs going from the Paleogene to the Neogene periods (average EQs of 0.43 and 0.64, respectively), while the EQs of South American ungulates were static over the same time interval (average EQ unchanged at 0.48).[18] This analysis was later criticized.[133] Jerison subsequently presented data suggesting that native South American ungulates also lagged in the relative size of their neocortices (a measurement not subject to the vagaries of body mass estimation).[134] Interestingly, the late survivor Toxodon had one of the highest EQ values (0.88) among native Neotropic ungulates.[133]
    Jerison also found that Neogene xenarthrans had low EQs, similar to those he obtained for South American ungulates.[132]
  23. ^ The estimated EQ of Thylacosmilus atrox, 0.41 (based on a brain mass of 43.2 g, a body mass of 26.4 kg,[135] and an EQ of 43.2/[0.12*26400^(2/3)][134]), is high for a sparassodont,[136] but is lower than that of modern felids, with a mean value of 0.87.[137] Estimates of 0.38[138] and 0.59[137] have been given for the EQ of much larger Smilodon fatalis (based on body mass estimates of 330 and 175 kg, respectively).
  24. ^ The giant tortoises of Asia and Africa[161] died out much earlier in the Quaternary than those of South America, Madagascar and Australia, while those of North America[162] died out around the same time.
  25. ^ P. S. Martin (2005), p. 175.[95]
  26. ^ A number of recently extinct North American (and in some cases also South American) taxa such as tapirs, equids, camelids, saiga antelope, proboscids, dholes, and lions survived in the Old World, probably mostly for different reasons – tapirs being a likely exception, since their Old World representative survived only in the rainforests of Southeast Asia. (Cheetahs in the broadest sense could be added to this list, although the New and Old World forms are in different genera.) Old World herbivores may in many cases have been able to learn to be vigilant about the presence of humans during a more gradual appearance (by development or migration) of advanced human hunters in their ranges. In the cases of predators, the Old World representatives in at least some locations would thus have suffered less from extinctions of their prey species. In contrast, the musk ox represents a rare example of a megafaunal taxon that recently became extinct in Asia, but survived in remote areas of arctic North America (its more southerly-distributed relatives, such as the woodland muskox and shrub-ox, were less fortunate).
  27. ^ a b c This listing currently has fairly complete coverage of mammals, but only spotty coverage of other groups. Crossings by nonflying mammals and birds occurred during the last 10 Ma. Crossings by fish, arthropods, rafting amphibians and reptiles, and flying bats and birds were made before 10 Ma ago in many cases. Taxa listed as invasive did not necessarily cross the isthmus themselves; they may have evolved in the adopted land mass from ancestral taxa that made the crossing.
  28. ^ Mixotoxodon remains have been collected in Central America and Mexico as far north as Veracruz and Michoacán, with a possible find in Tamaulipas;[176] additionally, one fossil tooth has been identified in eastern Texas, United States.[177]
  29. ^ Central America is usually defined physiographically as ending at the Isthmus of Tehuantepec or less commonly, at the Trans-Mexican Volcanic Belt. Most of the taxa that proceeded farther but failed to reach the present Mexican border are or were confined to tropical or subtropical climates similar to those of Central America. Examples include the giant anteater, the grayish mouse opossum, the lowland paca, and Geoffroy's spider monkey.
  30. ^ Fossils of the giant anteater have been found as far north as northwestern Sonora, Mexico.[186]
  31. ^ It has been proposed that monkeys invaded Central America in at least three and probably four waves, as follows: (1) an initial invasion by A. pigra and S. oerstedii ~ 3 Ma ago; (2) an invasion by A. palliata (giving rise to A. coibensis), A. geoffroyi and C. capucinus ~ 2 Ma ago; an invasion by A. zonalis and S. geoffroyi ~ 1 Ma ago; a most recent invasion by A. fusciceps. The species of the first wave have apparently been out-competed by those of the second, and now have much more restricted distributions.[187]
  32. ^ Salamanders apparently dispersed to South America by the Early Miocene, about 23 Ma ago.[191] Nevertheless, the salamander fauna of South America, which is restricted to the tropical region, consists of only 2 clades, and has fewer species and is far less diverse than that of much smaller Central America. Salamanders are believed to have originated in northern Pangea, perhaps not long before it separated to become Laurasia,[179] and are not present anywhere else in the Southern Hemisphere (see the world salamander distribution map). In contrast, caecilians have a mostly Gondwanan distribution. Apart from a small region of overlap in southern China and northern Southeast Asia, Central America and northern South America are the only places in the world where both salamanders and caecilians are present.
  33. ^ Condors apparently reached South America by the late Miocene or early Pliocene (4.5 – 6.0 Ma ago), several million years before the formation of the isthmus.[202] Condor-like forms in North America date back to the Barstovian stage (middle Miocene, 11.8 – 15.5 Ma ago).[201]
  34. ^ This is based on the definition of Sigmodontinae that excludes Neotominae and Tylomyinae.
  35. ^ Hippidion, a relatively short-legged equid that developed in South America after invading from North America about 2.5 Ma ago, has traditionally been thought to have evolved from pliohippines.[203][204] However, recent studies of the DNA of Hippidion and other New World Pleistocene horses indicate that Hippidion is actually a member of Equus, closely related to the extant horse, E. ferus.[203][204] Another invasion of South America by Equus occurred about one Ma ago, and this lineage, traditionally viewed as the subgenus Equus (Amerhippus), appears indistinguishable from E. ferus.[204] Both these lineages became extinct at the end of the Pleistocene, but E. ferus was reintroduced from Eurasia by Europeans in the 16th century. Note: the authors of the DNA sequence study of Equus (Amerhippus) use "E. caballus" as an alternative specific name for "E. ferus".[204]
  36. ^ Not to be confused with the American mastodon (Mammut americanum), a proboscid from a different family whose remains have been found no further south than Honduras.[205]
  37. ^ Not to be confused with the South American gray fox.
  1. ^ The native South American ungulates dwindled gradually as North American ungulates invaded and diversified. The changes in number and composition of South America's ungulate genera over time are given in the table below. The Quaternary extinction event that delivered the coup de grâce to the native Neotropic ungulates also dealt a heavy blow to South America's ungulate immigrants.
    Change in number of South American ungulate genera over time[90]
    Time interval Source region of genera
    Geologic period Range (Ma ago) South America North America Both
    Huayquerian 9.0–6.8 13 0 13
    Montehermosan 6.8–4.0 12 1 13
    Chapadmalalan 4.0–3.0 12 1 13
    Uquian 3.0–1.5 5 10 15
    Ensenadan 1.5–0.8 3 14 17
    Lujanian 0.8–0.011 3 20 23
    Holocene 0.011–0 0 11 11

References edit

  1. ^ O'Dea, A.; Lessios, H. A.; Coates, A. G.; et al. (2016). "Formation of the Isthmus of Panama". Science Advances. 2 (8): e1600883. Bibcode:2016SciA....2E0883O. doi:10.1126/sciadv.1600883. PMC 4988774. PMID 27540590.
  2. ^ a b Lessios, H.A. (December 2008). "The Great American Schism: Divergence of Marine Organisms After the Rise of the Central American Isthmus". Annual Review of Ecology, Evolution, and Systematics. 39: 63–91. doi:10.1146/annurev.ecolsys.38.091206.095815.
  3. ^ Wallace, Alfred Russel (1876). The Geographical Distribution of Animals. With a Study of the Relations of Living and Extinct Faunas as Elucidating the Past Changes of the Earth's Surface. Vol. 1. New York: Harper and Brothers. OCLC 556393.
  4. ^ Wallace, Alfred Russel (1876). Ibid.. Vol. 2. New York: Harper and Brothers. OCLC 556393.
  5. ^ a b Marshall, L. G. (July–August 1988). "Land Mammals and the Great American Interchange" (PDF). American Scientist. 76 (4): 380–388. Bibcode:1988AmSci..76..380M. Archived (PDF) from the original on 2013-03-02. Retrieved 2014-04-22.
  6. ^ Osborn, H. (1910). The Age Of Mammals In Europe, Asia, And North America. New York, EEUU: The Macmillan Company. pp. 80–81.
  7. ^ Karanth, K. Praveen (2006-03-25). "Out-of-India Gondwanan origin of some tropical Asian biota" (PDF). Current Science. 90 (6): 789–792. Retrieved 2008-12-29.
  8. ^ Clementz, Mark; Bajpai, S.; Ravikant, V.; Thewissen, J. G. M.; Saravanan, N.; Singh, I. B.; Prasad, V. (1 January 2011). "Early Eocene warming events and the timing of terrestrial faunal exchange between India and Asia". Geology. 39 (1): 15–18. Bibcode:2011Geo....39...15C. doi:10.1130/G31585.1. Retrieved 28 August 2022.
  9. ^ Rose, Kenneth D.; Rana, Rajendra S.; Sahni, Ashok; Kumar, Kishor; Singh, Lachham; Smith, Thierry (1 June 2009). "First Tillodont from India: Additional Evidence for an Early Eocene Faunal Connection between Europe and India?". Acta Palaeontologica Polonica. 54 (2): 351–355. doi:10.4202/app.2008.0067. S2CID 129644411.
  10. ^ Das, Debasis P.; Carolin, Nora; Bajpai, Sunil (22 Aug 2021). "A nyctitheriid insectivore (Eulipotyphla, Mammalia) of Asian affinity from the early Eocene of India". Historical Biology. 34 (7): 1157–1165. doi:10.1080/08912963.2021.1966002. S2CID 238735010. Retrieved 28 August 2022.
  11. ^ Rana, Rajendra S.; Kumar, Kishor; Escarguel, Gilles; Sahni, Ashok; Rose, Kenneth D.; Smith, Thierry; Singh, Hukam; Singh, Lachham (1 March 2008). "An Ailuravine Rodent from the Lower Eocene Cambay Formation at Vastan, Western India, and Its Palaeobiogeographic Implications". Acta Palaeontologica Polonica. 53 (1): 1–14. doi:10.4202/app.2008.0101. S2CID 56382943.
  12. ^ Li, Fengyuan; Shao, Lili; Li, Shuqiang (3 February 2020). "Tropical Niche Conservatism Explains the Eocene Migration from India to Southeast Asia in Ochyroceratid Spiders". Systematic Biology. 69 (5): 987–998. doi:10.1093/sysbio/syaa006. PMID 32011715. Retrieved 28 August 2022.
  13. ^ Ali, Jason R.; Aitchison, Jonathan C. (June 2008). "Gondwana to Asia: Plate tectonics, paleogeography and the biological connectivity of the Indian sub-continent from the Middle Jurassic through latest Eocene (166–35 Ma)". Earth-Science Reviews. 88 (3–4): 145–166. Bibcode:2008ESRv...88..145A. doi:10.1016/j.earscirev.2008.01.007. Retrieved 28 August 2022.
  14. ^ Dutta, Suryendu; Tripathi, Suryakant M.; Mallick, Monalisa; Mathews, Runcie P.; Greenwood, Paul F.; Rao, Mulagalapalli R.; Summons, Roger E. (July 2011). "Eocene out-of-India dispersal of Asian dipterocarps". Review of Palaeobotany and Palynology. 166 (1–2): 63–68. Bibcode:2011RPaPa.166...63D. doi:10.1016/j.revpalbo.2011.05.002. Retrieved 28 August 2022.
  15. ^ Kapur, Vivesh V.; Carolin, N.; Bajpai, S. (2022). "Early Paleogene mammal faunas of India: a review of recent advances with implications for the timing of initial India-Asia contact". Himalayan Geology. 47 (1B): 337–356. Retrieved 28 August 2022.
  16. ^ Hedges, S. Blair (2001-01-02). "Afrotheria: Plate tectonics meets genomics". Proceedings of the National Academy of Sciences. 98 (1): 1–2. Bibcode:2001PNAS...98....1H. doi:10.1073/pnas.98.1.1. PMC 33345. PMID 11136239.
  17. ^ Kappelman, John; Rasmussen, D. Tab; Sanders, William J.; Feseha, Mulugeta; Bown, Thomas; Copeland, Peter; Crabaugh, Jeff; Fleagle, John; Glantz, Michelle; Gordon, Adam; Jacobs, Bonnie; Maga, Murat; Muldoon, Kathleen; Pan, Aaron; Pyne, Lydia; Richmond, Brian; Ryan, Timothy; Seiffert, Erik R.; Sen, Sevket; Todd, Lawrence; Wiemann, Michael C.; Winkler, Alisa C. (4 December 2003). "Oligocene mammals from Ethiopia and faunal exchange between Afro-Arabia and Eurasia". Nature. 426 (6966): 549–552. Bibcode:2003Natur.426..549K. doi:10.1038/nature02102. hdl:2027.42/62496. PMID 14654838. S2CID 4336007. Retrieved 28 August 2022.
  18. ^ a b c Simpson, George Gaylord (1980). Splendid Isolation: The Curious History of South American Mammals. New Haven: Yale University Press. ISBN 978-0-300-02434-0. OCLC 5219346.
  19. ^ Bond, M.; Kramarz, A.; MacPhee, R. D. E.; Reguero, M. (2011). "A new astrapothere (Mammalia, Meridiungulata) from La Meseta Formation, Seymour (Marambio) Island, and a reassessment of previous records of Antarctic astrapotheres" (PDF). American Museum Novitates (3718): 1–16. doi:10.1206/3718.2. hdl:11336/98139. S2CID 58908785.
  20. ^ M. Bond; M. A. Reguero; S. F. Vizcaíno; S. A. Marenssi (2006). "A new 'South American ungulate' (Mammalia: Litopterna) from the Eocene of the Antarctic Peninsula". In J. E. Francis; D. Pirrie; J. A. Crame (eds.). Cretaceous-tertiary high-latitude palaeoenvironments: James Ross Basin, Antarctica. Vol. 258. The Geological Society of London. pp. 163–176. Bibcode:2006GSLSP.258..163B. doi:10.1144/GSL.SP.2006.258.01.12. S2CID 140546667. {{cite book}}: |journal= ignored (help)
  21. ^ Gelfo, J. N.; Mörs, T.; Lorente, M.; López, G. M.; Reguero, M.; O'Regan, H. (2014-07-16). "The oldest mammals from Antarctica, early Eocene of the La Meseta Formation, Seymour Island". Palaeontology. 58 (1): 101–110. doi:10.1111/pala.12121. S2CID 140177511.
  22. ^ Welker, F.; Collins, M. J.; Thomas, J. A.; Wadsley, M.; Brace, S.; Cappellini, E.; Turvey, S. T.; Reguero, M.; Gelfo, J. N.; Kramarz, A.; Burger, J.; Thomas-Oates, J.; Ashford, D. A.; Ashton, P. D.; Rowsell, K.; Porter, D. M.; Kessler, B.; Fischer, R.; Baessmann, C.; Kaspar, S.; Olsen, J. V.; Kiley, P.; Elliott, J. A.; Kelstrup, C. D.; Mullin, V.; Hofreiter, M.; Willerslev, E.; Hublin, J.-J.; Orlando, L.; Barnes, I.; MacPhee, R. D. E. (2015-03-18). "Ancient proteins resolve the evolutionary history of Darwin's South American ungulates". Nature. 522 (7554): 81–84. Bibcode:2015Natur.522...81W. doi:10.1038/nature14249. hdl:11336/14769. ISSN 0028-0836. PMID 25799987. S2CID 4467386.
  23. ^ Buckley, M. (2015-04-01). "Ancient collagen reveals evolutionary history of the endemic South American 'ungulates'". Proceedings of the Royal Society B: Biological Sciences. 282 (1806): 20142671. doi:10.1098/rspb.2014.2671. PMC 4426609. PMID 25833851.
  24. ^ Westbury, M.; Baleka, S.; Barlow, A.; Hartmann, S.; Paijmans, J. L. A.; Kramarz, A.; Forasiepi, A. M.; Bond, M.; Gelfo, J. N.; Reguero, M. A.; López-Mendoza, P.; Taglioretti, M.; Scaglia, F.; Rinderknecht, A.; Jones, W.; Mena, F.; Billet, G.; de Muizon, C.; Aguilar, J. L.; MacPhee, R. D. E.; Hofreiter, M. (2017-06-27). "A mitogenomic timetree for Darwin's enigmatic South American mammal Macrauchenia patachonica". Nature Communications. 8: 15951. Bibcode:2017NatCo...815951W. doi:10.1038/ncomms15951. PMC 5490259. PMID 28654082.
  25. ^ Chimento, Nicolás R.; Agnolin, Federico L.; Novas, Fernando E. (2015). "The bizarre 'metatherians' Groeberia and Patagonia, late surviving members of gondwanatherian mammals". Historical Biology: An International Journal of Paleobiology. 27 (5): 603–623. Bibcode:2015HBio...27..603C. doi:10.1080/08912963.2014.903945. hdl:11336/85076. S2CID 216591096.
  26. ^ a b Nilsson, M. A.; Churakov, G.; Sommer, M.; Van Tran, N.; Zemann, A.; Brosius, J.; Schmitz, J. (2010-07-27). "Tracking Marsupial Evolution Using Archaic Genomic Retroposon Insertions". PLOS Biology. 8 (7): e1000436. doi:10.1371/journal.pbio.1000436. PMC 2910653. PMID 20668664.
  27. ^ Pascual, R.; Archer, M.; Jaureguizar, E. O.; Prado, J. L.; Godthelp, H.; Hand, S. J. (1992). "First discovery of monotremes in South America". Nature. 356 (6371): 704–706. Bibcode:1992Natur.356..704P. doi:10.1038/356704a0. S2CID 4350045.
  28. ^ Pascual, R.; Goin, F. J.; Balarino, L.; Sauthier, D. E. U. (2002). "New data on the Paleocene monotreme Monotrematum sudamericanum, and the convergent evolution of triangulate molars" (PDF). Acta Palaeontologica Polonica. 47 (3): 487–492. Retrieved 2020-01-21.
  29. ^ Briggs, J. C. (August 2003). "Fishes and Birds: Gondwana Life Rafts Reconsidered". Syst. Biol. 52 (4): 548–553. doi:10.1080/10635150390218385. JSTOR 3651142. PMID 12857645.
  30. ^ Yonezawa, T.; Segawa, T.; Mori, H.; Campos, P. F.; Hongoh, Y.; Endo, H.; Akiyoshi, A.; Kohno, N.; Nishida, S.; Wu, J.; Jin, H.; Adachi, J.; Kishino, H.; Kurokawa, K.; Nogi, Y.; Tanabe, H.; Mukoyama, H.; Yoshida, K.; Rasoamiaramanana, A.; Yamagishi, S.; Hayashi, Y.; Yoshida, A.; Koike, H.; Akishinonomiya, F.; Willerslev, E.; Hasegawa, M. (2016-12-15). "Phylogenomics and Morphology of Extinct Paleognaths Reveal the Origin and Evolution of the Ratites". Current Biology. 27 (1): 68–77. Bibcode:2017CBio...27...68Y. doi:10.1016/j.cub.2016.10.029. PMID 27989673. S2CID 38890667.
  31. ^ Naish, Darren (29 June 2008). "Invasion of the marsupial weasels, dogs, cats and bears... or is it?". scienceblogs.com. Retrieved 2008-12-07.
  32. ^ Naish, Darren (2006-10-27). "Terror birds". darrennaish.blogspot.com. Retrieved 2008-03-29.
  33. ^ Alvarenga, H. M. F.; Höfling, E. (2003). "Systematic Revision of the Phorusrhacidae (Aves: Ralliformes)". Papéis Avulsos de Zoologia. 43 (4): 55–91. doi:10.1590/S0031-10492003000400001.
  34. ^ Palmqvist, Paul; Vizcaíno, Sergio F. (2003-09-30). "Ecological and reproductive constraints of body size in the gigantic Argentavis magnificens (Aves, Theratornithidae) from the Miocene of Argentina" (PDF). Ameghiniana. 40 (3): 379–385. Retrieved 2008-12-11.
  35. ^ Paolillo, A.; Linares, O. J. (2007-06-05). "Nuevos Cocodrilos Sebecosuchia del Cenozoico Suramericano (Mesosuchia: Crocodylia)" (PDF). Paleobiologia Neotropical. 3: 1–25. Archived from the original (PDF) on 2009-03-03. Retrieved 2008-09-28.
  36. ^ Busbey, Arthur B. III (1986-03-07). "New Material of Sebecus cf. huilensis (Crocodilia: Sebecosuchidae) from the Miocene La Venta Formation of Colombia". Journal of Vertebrate Paleontology. 6 (1): 20–27. Bibcode:1986JVPal...6...20B. doi:10.1080/02724634.1986.10011595. JSTOR 4523070.
  37. ^ Salas-Gismondi, R.; et al. (2007). "Middle Miocene Crocodiles From the Fitzcarrald Arch, Amazonian Peru" (PDF). In Díaz-Martínez, E.; Rábano, I. (eds.). 4th European Meeting on the Palaeontology and Stratigraphy of Latin America. Madrid: Instituto Geológico y Minero de España. pp. 355–360. ISBN 978-84-7840-707-1.
  38. ^ a b c Gasparini, Zulma (September 1984). "New Tertiary Sebecosuchia (Crocodylia: Mesosuchia) from Argentina". Journal of Vertebrate Paleontology. 4 (1): 85–95. Bibcode:1984JVPal...4...85G. doi:10.1080/02724634.1984.10011988. JSTOR 4522967.
  39. ^ Liccardo, A.; Weinschütz, L. C. (September 2010). "Registro inédito de fósseis de vertebrados na Bacia Sedimentar de Curitiba (PR)". Revista Brasileira de Geociências (in Portuguese). 40 (3): 330–338. doi:10.25249/0375-7536.2010403330338 (inactive 2024-02-01). Retrieved 2017-10-23.{{cite journal}}: CS1 maint: DOI inactive as of February 2024 (link)
  40. ^ Möller-Krull, Maren; Delsuc, Frédéric; Churakov, Gennady; Marker, Claudia; Superina, Mariella; Brosius, Jürgen; Douzery, Emmanuel J. P.; Schmitz, Jürgen (2007-09-17). "Retroposed Elements and Their Flanking Regions Resolve the Evolutionary History of Xenarthran Mammals (Armadillos, Anteaters, and Sloths)". Molecular Biology and Evolution. 24 (11): 2573–2582. doi:10.1093/molbev/msm201. PMID 17884827.
  41. ^ Muizon, C. de; McDonald, H. G.; Salas, R.; Urbina, M. (June 2004). "The evolution of feeding adaptations of the aquatic sloth Thalassocnus". Journal of Vertebrate Paleontology. 24 (2): 398–410. Bibcode:2004JVPal..24..398D. doi:10.1671/2429b. JSTOR 4524727. S2CID 83859607.
  42. ^ Amson, E.; Muizon, C. de; Laurin, M.; Argot, C.; de Buffrénil, V. (2014). "Gradual adaptation of bone structure to aquatic lifestyle in extinct sloths from Peru". Proceedings of the Royal Society B: Biological Sciences. 281 (1782): 20140192. doi:10.1098/rspb.2014.0192. PMC 3973278. PMID 24621950.
  43. ^ a b c Poux, C.; Chevret, P.; Huchon, D.; De Jong, W. W.; Douzery, E. J. P. (2006). "Arrival and Diversification of Caviomorph Rodents and Platyrrhine Primates in South America" (PDF). Systematic Biology. 55 (2): 228–244. doi:10.1080/10635150500481390. PMID 16551580. Archived from the original (PDF) on 2016-06-02. Retrieved 2011-10-25.
  44. ^ Mangels, J. (2011-10-15). "Case Western Reserve University expert uses fossil teeth to recast history of rodent". Cleveland Live, Inc. Retrieved 2011-10-25.
  45. ^ Antoine, P.-O.; Marivaux, L.; Croft, D. A.; Billet, G.; Ganerod, M.; Jaramillo, C.; Martin, T.; Orliac, M. J.; et al. (2011). "Middle Eocene rodents from Peruvian Amazonia reveal the pattern and timing of caviomorph origins and biogeography". Proceedings of the Royal Society B: Biological Sciences. 279 (1732): 1319–1326. doi:10.1098/rspb.2011.1732. PMC 3282368. PMID 21993503.
  46. ^ Flynn, J. J.; Wyss, A. R. (1998). "Recent advances in South American mammalian paleontology". Trends in Ecology and Evolution. 13 (11): 449–454. doi:10.1016/S0169-5347(98)01457-8. PMID 21238387.
  47. ^ Flynn, John J.; Wyss, André R.; Charrier, Reynaldo (2007). "South America's Missing Mammals". Scientific American. 296 (May): 68–75. Bibcode:2007SciAm.296e..68F. doi:10.1038/scientificamerican0507-68. PMID 17500416.
  48. ^ Fratantoni, D. M.; Johns, W. E.; Townsend, T. L.; Hurlburt, H. E. (August 2000). "Low-Latitude Circulation and Mass Transport Pathways in a Model of the Tropical Atlantic Ocean". Journal of Physical Oceanography. 30 (8): 1944–1966. Bibcode:2000JPO....30.1944F. doi:10.1175/1520-0485(2000)030<1944:LLCAMT>2.0.CO;2. S2CID 54534830.
  49. ^ a b Hedges, S. Blair (November 1996). "Historical biogeography of West Indian vertebrates". Annual Review of Ecology and Systematics. 27 (1): 163–196. doi:10.1146/annurev.ecolsys.27.1.163.
  50. ^ a b Hedges, S. Blair (2006-08-23). "Paleogrography of the Antilles and Origin of West Indian Terrestrial Vertebrates". Annals of the Missouri Botanical Garden. 93 (2): 231–244. doi:10.3417/0026-6493(2006)93[231:POTAAO]2.0.CO;2. S2CID 198149958.
  51. ^ Vélez-Juarbe, J.; Martin, T.; Macphee, R. D. E.; Ortega-Ariza, D. (January 2014). "The earliest Caribbean rodents: Oligocene caviomorphs from Puerto Rico". Journal of Vertebrate Paleontology. 34 (1): 157–163. Bibcode:2014JVPal..34..157V. doi:10.1080/02724634.2013.789039. S2CID 140178414.
  52. ^ Millien, Virginie (May 2008). "The largest among the smallest: the body mass of the giant rodent Josephoartigasia monesi". Proceedings of the Royal Society B. 275 (1646): 1953–5, discussion 1957–8. doi:10.1098/rspb.2008.0087. PMC 2596365. PMID 18495621.
  53. ^ Bond, M.; Tejedor, M. F.; Campbell, K. E.; Chornogubsky, L.; Novo, N.; Goin, F. (2015-02-04). "Eocene primates of South America and the African origins of New World monkeys". Nature. 520 (7548): 538–541. Bibcode:2015Natur.520..538B. doi:10.1038/nature14120. hdl:11336/79088. PMID 25652825. S2CID 4456556.
  54. ^ Wade, L. (2016-04-20). "Monkey ancestors rafted across the sea to North America". Science. doi:10.1126/science.aaf4154.
  55. ^ Bloch, J. I.; Woodruff, E. D.; Wood, A. R.; Rincon, A. F.; Harrington, A. R.; Morgan, G. S.; Foster, D. A.; Montes, C.; Jaramillo, C. A.; Jud, N. A.; Jones, D. S.; MacFadden, B. J. (2016-04-20). "First North American fossil monkey and early Miocene tropical biotic interchange". Nature. 533 (7602): 243–246. Bibcode:2016Natur.533..243B. doi:10.1038/nature17415. PMID 27096364. S2CID 4445687.
  56. ^ Seiffert, E.R.; Tejedor, M.F.; Fleagle, J.G.; Novo, N.M.; Cornejo, F.M.; Bond, M.; de Vries, D.; Campbell, K.E. (2020). "A parapithecid stem anthropoid of African origin in the Paleogene of South America". Science. 368 (6487): 194–197. Bibcode:2020Sci...368..194S. doi:10.1126/science.aba1135. PMID 32273470. S2CID 215550773.
  57. ^ Godinot, M. (2020). "Rafting on a wide and wild ocean". Science. 368 (6487): 136–137. Bibcode:2020Sci...368..136G. doi:10.1126/science.abb4107. PMID 32273458. S2CID 215551148.
  58. ^ a b Chaimanee, Y.; Chavasseau, O.; Beard, K. C.; Kyaw, A. A.; Soe, A. N.; Sein, C.; Lazzari, V.; Marivaux, L.; Marandat, B.; Swe, M.; Rugbumrung, M.; Lwin, T.; Valentin, X.; Zin-Maung-Maung-Thein; Jaeger, J. -J. (2012). "Late Middle Eocene primate from Myanmar and the initial anthropoid colonization of Africa". Proceedings of the National Academy of Sciences. 109 (26): 10293–10297. Bibcode:2012PNAS..10910293C. doi:10.1073/pnas.1200644109. PMC 3387043. PMID 22665790.
  59. ^ a b c d e f g h Lim, B. K. (July 2009). "Review of the Origins and Biogeography of Bats in South America". Chiroptera Neotropical. 15 (1). Departamento de Zoologia – Universidade de Brasília: 391–410. Retrieved 2018-01-24.
  60. ^ a b c d e f g Gunnell, G. F.; Simmons, N. B.; Seiffert, E. R. (2014-02-04). "New Myzopodidae (Chiroptera) from the Late Paleogene of Egypt: Emended Family Diagnosis and Biogeographic Origins of Noctilionoidea". PLoS ONE. 9 (2): e86712. Bibcode:2014PLoSO...986712G. doi:10.1371/journal.pone.0086712. PMC 3913578. PMID 24504061.
  61. ^ a b Teeling, E. C.; Springer, M.; Madsen, O.; Bates, P.; O'Brien, S.; Murphy, W. (2005-01-28). "A Molecular Phylogeny for Bats Illuminates Biogeography and the Fossil Record". Science. 307 (5709): 580–584. Bibcode:2005Sci...307..580T. doi:10.1126/science.1105113. PMID 15681385. S2CID 25912333.
  62. ^ a b Le, M.; Raxworthy, C. J.; McCord, W. P.; Mertz, L. (2006-05-05). "A molecular phylogeny of tortoises (Testudines: Testudinidae) based on mitochondrial and nuclear genes" (PDF). Molecular Phylogenetics and Evolution. 40 (2): 517–531. doi:10.1016/j.ympev.2006.03.003. PMID 16678445. Archived from the original (PDF) on 2012-03-16. Retrieved 2012-04-12.
  63. ^ Williams, E. E.; Anthony, H. E.; Goodwin, G. G. (1952). "A new fossil tortoise from Mona Island, West Indies, and a tentative arrangement of the tortoises of the world". Bulletin of the American Museum of Natural History. 99 (9). American Museum of Natural History: 541–560. hdl:2246/418.
  64. ^ Gamble, T.; Bauer, A. M.; Colli, G. R.; Greenbaum, E.; Jackman, T. R.; Vitt, L. J.; Simons, A. M. (2010-12-03). "Coming to America: Multiple origins of New World geckos" (PDF). Journal of Evolutionary Biology. 24 (2): 231–244. doi:10.1111/j.1420-9101.2010.02184.x. PMC 3075428. PMID 21126276.
  65. ^ Carranza, S.; Arnold, N. E. (2003-08-05). "Investigating the origin of transoceanic distributions: mtDNA shows Mabuya lizards (Reptilia, Scincidae) crossed the Atlantic twice". Systematics and Biodiversity. 1 (2): 275–282. Bibcode:2003SyBio...1..275C. doi:10.1017/S1477200003001099. S2CID 55799145.
  66. ^ Vidal, N.; Azvolinsky, A.; Cruaud, C.; Hedges, S. B. (2007-12-11). "Origin of tropical American burrowing reptiles by transatlantic rafting". Biology Letters. 4 (1): 115–118. doi:10.1098/rsbl.2007.0531. PMC 2412945. PMID 18077239.
  67. ^ Vidal, N.; Marin, J.; Morini, M.; Donnellan, S.; Branch, W. R. [in French]; Thomas, R. [in French]; Vences, M.; Wynn, A.; Cruaud, C.; Hedges, S. B. (2010-03-31). "Blindsnake evolutionary tree reveals long history on Gondwana". Biology Letters. 6 (4): 558–561. doi:10.1098/rsbl.2010.0220. PMC 2936224. PMID 20356885.
  68. ^ Mayr, G.; Alvarenga, H.; Mourer-Chauviré, C. C. (2011-10-01). "Out of Africa: Fossils shed light on the origin of the hoatzin, an iconic Neotropic bird". Naturwissenschaften. 98 (11): 961–966. Bibcode:2011NW.....98..961M. doi:10.1007/s00114-011-0849-1. PMID 21964974. S2CID 24210185.
  69. ^ Woodburne, M. O. (2010-07-14). "The Great American Biotic Interchange: Dispersals, Tectonics, Climate, Sea Level and Holding Pens". Journal of Mammalian Evolution. 17 (4): 245–264. doi:10.1007/s10914-010-9144-8. PMC 2987556. PMID 21125025.
  70. ^ Koepfli, K.-P.; Gompper, M. E.; Eizirik, E.; Ho, C.-C.; Linden, L.; Maldonado, J.; Wayne, E. R. K. (2007). "Phylogeny of the Procyonidae (Mammalia: Carnivora): Molecules, morphology and the Great American Interchange". Molecular Phylogenetics and Evolution. 43 (3): 1076–1095. CiteSeerX 10.1.1.495.2618. doi:10.1016/j.ympev.2006.10.003. PMID 17174109.
  71. ^ a b Webb, S. D. (1976). "Mammalian Faunal Dynamics of the Great American Interchange". Paleobiology. 2 (3): 220–234. Bibcode:1976Pbio....2..220W. doi:10.1017/S0094837300004802. JSTOR 2400220. S2CID 251050063.
  72. ^ a b Marshall, L. G.; Butler, R. F.; Drake, R. E.; Curtis, G. H.; Tedford, R. H. (1979-04-20). "Calibration of the Great American Interchange". Science. 204 (4390): 272–279. Bibcode:1979Sci...204..272M. doi:10.1126/science.204.4390.272. PMID 17800342. S2CID 8625188.
  73. ^ Engel, S. R.; Hogan, K. M.; Taylor, J. F.; Davis, S. K. (1998). "Molecular Systematics and Paleobiogeography of the South American Sigmodontine Rodents". Molecular Biology and Evolution. 15 (1): 35–49. doi:10.1093/oxfordjournals.molbev.a025845. PMID 9491603.
  74. ^ Smith, M. F.; Patton, J. L. (1999). "Phylogenetic Relationships and the Radiation of Sigmodontine Rodents in South America: Evidence from Cytochrome b". Journal of Mammalian Evolution. 6 (2): 89–128. doi:10.1023/A:1020668004578. S2CID 22355532.
  75. ^ Parada, A.; Pardiñas, U. F. J.; Salazar-Bravo, J.; D'Elía, G.; Palma, R. E. (March 2013). "Dating an impressive Neotropical radiation: Molecular time estimates for the Sigmodontinae (Rodentia) provide insights into its historical biogeography". Molecular Phylogenetics and Evolution. 66 (3): 960–968. doi:10.1016/j.ympev.2012.12.001. hdl:11336/5595. PMID 23257216.
  76. ^ Defler, T. (2019). "The Great American Biotic (Faunal) Interchange". History of Terrestrial Mammals in South America. Topics in Geobiology. Vol. 42. pp. 259–287. doi:10.1007/978-3-319-98449-0_13. ISBN 978-3-319-98448-3. S2CID 134089273.
  77. ^ Campbell, K. E.; Frailey, C. D.; Romero-Pittman, L. (2000). "The Late Miocene Gomphothere Amahuacatherium peruvium (Proboscidea: Gomphotheriidae) from Amazonian Peru: Implications for the Great American Faunal Interchange-[Boletín D 23]". Ingemmet.
  78. ^ Campbell, K. E.; Frailey, C. D.; Romero-Pittman, L. (April 2009). "In defense of Amahuacatherium (Proboscidea: Gomphotheriidae)". Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen. 252 (1): 113–128. doi:10.1127/0077-7749/2009/0252-0113.
  79. ^ a b c Frailey, C. D.; Campbell, K. E. (September 2012). "Two New Genera of Peccaries (Mammalia, Artiodactyla, Tayassuidae) from Upper Miocene Deposits of the Amazon Basin". Journal of Paleontology. 86 (5): 852–877. Bibcode:2012JPal...86..852F. doi:10.1666/12-012.1. S2CID 85961848.
  80. ^ a b Prothero, D. R.; Campbell, K. E.; Beatty, B. L.; Frailey, C. D. (May 2014). "New late Miocene dromomerycine artiodactyl from the Amazon Basin: Implications for interchange dynamics". Journal of Paleontology. 88 (3): 434–443. doi:10.1666/13-022. S2CID 55245294.
  81. ^ Prado, J. L.; Alberdi, M. T.; Azanza, B.; Sánchez, B.; Frassinetti, D. (2005). "The Pleistocene Gomphotheriidae (Proboscidea) from South America". Quaternary International. 126–128: 21–30. Bibcode:2005QuInt.126...21P. doi:10.1016/j.quaint.2004.04.012.
  82. ^ Lucas, S.G. (January 2013). "The palaeobiogeography of South American gomphotheres". Journal of Palaeogeography. 2 (1): 19–40. Bibcode:2013JPalG...2...19L. doi:10.3724/SP.J.1261.2013.00015 (inactive 31 January 2024). Retrieved 2020-01-23.{{cite journal}}: CS1 maint: DOI inactive as of January 2024 (link)
  83. ^ a b c Mothé, D.; dos Santos Avilla, L.; Asevedo, L.; Borges-Silva, L.; Rosas, M.; Labarca-Encina, R.; Souberlich, R.; Soibelzon, E.; Roman-Carrion, J.L.; Ríos, S.D.; Rincon, A.D.; de Oliveira, G.C.; Lopes, R.P. (2017). "Sixty years after 'The mastodonts of Brazil': The state of the art of South American proboscideans (Proboscidea, Gomphotheriidae)". Quaternary International. 443: 52–64. Bibcode:2017QuInt.443...52M. doi:10.1016/j.quaint.2016.08.028.
  84. ^ Antoine, P.-O.; Salas-Gismondi, R.; Pujos, F.; Ganerød, M.; Marivaux, L. (2016). "Western Amazonia as a Hotspot of Mammalian Biodiversity Throughout the Cenozoic". Journal of Mammalian Evolution. 24 (1): 5–17. doi:10.1007/s10914-016-9333-1. S2CID 43163341.
  85. ^ Presslee, S.; Slater, G. J.; Pujos, F.; Forasiepi, A. M.; Fischer, R.; Molloy, K.; Mackie, M.; Olsen, J. V.; Kramarz, A.; Taglioretti, M.; Scaglia, F.; Lezcano, M.; Lanata, J. L.; Southon, J.; Feranec, R.; Bloch, J.; Hajduk, A.; Martin, F. M.; Gismondi, R. S.; Reguero, M.; de Muizon, C.; Greenwood, A.; Chait, B. T.; Penkman, K.; Collins, M.; MacPhee, R.D.E. (2019). "Palaeoproteomics resolves sloth relationships" (PDF). Nature Ecology & Evolution. 3 (7): 1121–1130. Bibcode:2019NatEE...3.1121P. doi:10.1038/s41559-019-0909-z. PMID 31171860. S2CID 174813630.
  86. ^ Morgan, Gary S. (2002), "Late Rancholabrean Mammals from Southernmost Florida, and the Neotropical Influence in Florida Pleistocene Faunas", in Emry, Robert J. (ed.), Cenozoic Mammals of Land and Sea: Tributes to the Career of Clayton E. Ray, Smithsonian Contributions to Paleobiology, vol. 93, Washington, D.C., pp. 15–38{{citation}}: CS1 maint: location missing publisher (link)
  87. ^ McFadden, B.; Labs-Hochstein, J.; Hulbert, R.C. Jr.; Baskin, J.A. (2007). "Revised age of the late Neogene terror bird (Titanis) in North America during the Great American Interchange". Geology. 35 (2): 123–126. Bibcode:2007Geo....35..123M. doi:10.1130/G23186A.1.
  88. ^ a b c d e Webb, S. David (23 August 2006). "The Great American Biotic Interchange: Patterns and Processes". Annals of the Missouri Botanical Garden. 93 (2): 245–257. doi:10.3417/0026-6493(2006)93[245:TGABIP]2.0.CO;2. S2CID 198152030.
  89. ^ Marshall, L. G.; Cifelli, R. L. (1990). "Analysis of changing diversity patterns in Cenozoic land mammal age faunas, South America". Palaeovertebrata. 19: 169–210. Retrieved 2018-10-08.
  90. ^ a b Webb, S. D. (1991). "Ecogeography and the Great American Interchange". Paleobiology. 17 (3): 266–280. Bibcode:1991Pbio...17..266W. doi:10.1017/S0094837300010605. JSTOR 2400869. S2CID 88305955.
  91. ^ "Chalicotheriidae". www.amnh.org. AMNH. Retrieved 2018-10-08.
  92. ^ Gaudioso, P. J.; Gasparini, G. M.; Herbst, R.; Barquez, R. M. (2017). "First record of the Neolicaphrium recens Frenguelli, 1921 (Mammalia, Litopterna) in the Pleistocene of Santiago del Estero Province, Argentina". Papéis Avulsos de Zoologia. 57 (31): 23–29. doi:10.11606/0031-1049.2017.57.03. hdl:11336/56638.
  93. ^ a b c Simpson, George Gaylord (July 1950). "History of the Fauna of Latin America". American Scientist. 38 (3): 361–389. JSTOR 27826322. Retrieved 2013-02-14.
  94. ^ a b c Cione, A. L.; Tonni, E. P.; Soibelzon, L. (2003). "The Broken Zig-Zag: Late Cenozoic large mammal and tortoise extinction in South America" (PDF). Rev. Mus. Argentino Cienc. Nat. N.S. 5 (1): 1–19. doi:10.22179/REVMACN.5.26. Archived from the original (PDF) on 2011-07-06. Retrieved 2011-02-06.
  95. ^ a b c 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. OCLC 58055404.
  96. ^ McDonald, H. G.; Harington, C. R.; De Iuliis, G. (September 2000). "The Ground Sloth Megalonyx from Pleistocene Deposits of the Old Crow Basin, Yukon, Canada" (PDF). Arctic. 53 (3): 213–220. doi:10.14430/arctic852. Retrieved 2008-08-16.
  97. ^ Stock, C. (1942-05-29). "A ground sloth in Alaska". Science. 95 (2474): 552–553. Bibcode:1942Sci....95..552S. doi:10.1126/science.95.2474.552. PMID 17790868.
  98. ^ a b c Prevosti, F. J. (2010-09-07). "Phylogeny of the large extinct South American Canids (Mammalia, Carnivora, Canidae) using a "total evidence" approach". Cladistics. 26 (5): 456–481. doi:10.1111/j.1096-0031.2009.00298.x. PMID 34875763. S2CID 86650539.
  99. ^ a b c Perini, F. A.; Russo, C. A. M.; Schrago, C. G. (2009-11-26). "The evolution of South American endemic canids: a history of rapid diversification and morphological parallelism". Journal of Evolutionary Biology. 23 (2): 311–322. doi:10.1111/j.1420-9101.2009.01901.x. PMID 20002250. S2CID 20763999.
  100. ^ a b Slater, G. J.; Thalmann, O.; Leonard, J. A.; Schweizer, R. M.; Koepfli, K.-P.; Pollinger, J. P.; Rawlence, N. J.; Austin, J. J.; et al. (2009-11-03). "Evolutionary history of the Falklands wolf". Current Biology. 19 (20): R937–R938. Bibcode:2009CBio...19.R937S. doi:10.1016/j.cub.2009.09.018. hdl:10261/58562. PMID 19889366. S2CID 36185744.
  101. ^ Jain, S.; Collins, L. S. (2007-04-30). "Trends in Caribbean Paleoproductivity related to the Neogene closure of the Central American Seaway". Marine Micropaleontology. 63 (1–2): 57–74. Bibcode:2007MarMP..63...57J. doi:10.1016/j.marmicro.2006.11.003.
  102. ^ Amson, E.; Argot, C.; McDonald, H. G.; de Muizon, C. (2015). "Osteology and functional morphology of the axial postcranium of the marine sloth Thalassocnus (Mammalia, Tardigrada) with paleobiological implications". Journal of Mammalian Evolution. 22 (4): 473–518. doi:10.1007/s10914-014-9280-7. S2CID 16700349.
  103. ^ a b Marshall, L.G. (1976). Evolution of the Thylacosmilidae, extinct saber-tooth marsupials of South America. Museum of Paleontology, University of California. OCLC 8409501.
  104. ^ Marshall, L.G. (1977). "Evolution of the Carnivorous Adaptive Zone in South America". Major Patterns in Vertebrate Evolution. Springer US. pp. 709–721. doi:10.1007/978-1-4684-8851-7_24. ISBN 978-1-4684-8853-1. Retrieved 2020-01-24.
  105. ^ a b c d e f g Prevosti, F.J.; Forasiepi, A.M. (2018). Evolution of South American Mammalian Predators During the Cenozoic: Paleobiogeographic and Paleoenvironmental Contingencies. Springer Geology. Cham: Springer International Publishing. Bibcode:2018esam.book.....P. doi:10.1007/978-3-319-03701-1. ISBN 978-3-319-03700-4. S2CID 134939823.
  106. ^ a b c Prevosti, F.J.; Forasiepi, A.; Zimicz, N. (March 2013). "The Evolution of the Cenozoic Terrestrial Mammalian Predator Guild in South America: Competition or Replacement?". Journal of Mammalian Evolution. 20 (1): 3–21. doi:10.1007/s10914-011-9175-9. hdl:11336/2663. S2CID 15751319.
  107. ^ López-Aguirre, C.; Archer, M.; Hand, S.J.; Laffan, S.W. (January 2017). Smith, A. (ed.). "Extinction of South American sparassodontans (Metatheria): environmental fluctuations or complex ecological processes?". Palaeontology. 60 (1): 91–115. Bibcode:2017Palgy..60...91L. doi:10.1111/pala.12272. S2CID 132584097.
  108. ^ Prevosti, F.J.; Pardiñas, U.F.J. (September 2009). "Comment on "The oldest South American Cricetidae (Rodentia) and Mustelidae (Carnivora): Late Miocene faunal turnover in central Argentina and the Great American Biotic Interchange" by D.H. Verzi and C.I. Montalvo [Palaeogeography, Palaeoclimatology, Palaeoecology 267 (2008) 284–291]". Palaeogeography, Palaeoclimatology, Palaeoecology. 280 (3–4): 543–547. Bibcode:2009PPP...280..543P. doi:10.1016/j.palaeo.2009.05.021. hdl:11336/94480.
  109. ^ Croft, D.A.; Engelman, R.K.; Dolgushina, T.; Wesley, G. (2018-01-10). "Diversity and disparity of sparassodonts (Metatheria) reveal non-analogue nature of ancient South American mammalian carnivore guilds". Proceedings of the Royal Society B: Biological Sciences. 285 (1870): 20172012. doi:10.1098/rspb.2017.2012. PMC 5784193. PMID 29298933.
  110. ^ a b c Zimicz, N. (December 2014). "Avoiding Competition: the Ecological History of Late Cenozoic Metatherian Carnivores in South America". Journal of Mammalian Evolution. 21 (4): 383–393. doi:10.1007/s10914-014-9255-8. ISSN 1064-7554. S2CID 10161199.
  111. ^ a b Engelman, R.K.; Anaya, F.; Croft, D.A. (2020). "Australogale leptognathus, gen. et sp. nov., a Second Species of Small Sparassodont (Mammalia: Metatheria) from the Middle Miocene Locality of Quebrada Honda, Bolivia". Journal of Mammalian Evolution. 27: 37–54. doi:10.1007/s10914-018-9443-z. ISSN 1064-7554. S2CID 49473591.
  112. ^ a b Beck, R.M.D.; Taglioretti, M.L. (2019). "A Nearly Complete Juvenile Skull of the Marsupial Sparassocynus derivatus from the Pliocene of Argentina, the Affinities of "Sparassocynids", and the Diversification of Opossums (Marsupialia; Didelphimorphia; Didelphidae)". Journal of Mammalian Evolution. 27 (3): 385–417. doi:10.1007/s10914-019-09471-y. S2CID 198190603.
  113. ^ a b Engelman, R.K.; Croft, D.A. (2014-04-16). "A new species of small-bodied sparassodont (Mammalia, Metatheria) from the middle Miocene locality of Quebrada Honda, Bolivia". Journal of Vertebrate Paleontology. 34 (3): 672–688. Bibcode:2014JVPal..34..672E. doi:10.1080/02724634.2013.827118. ISSN 0272-4634. S2CID 84680371.
  114. ^ a b c Engelman, R.K.; Croft, D.A. (September 2019). "Strangers in a strange land: Ecological dissimilarity to metatherian carnivores may partly explain early colonization of South America by Cyonasua-group procyonids". Paleobiology. 45 (4): 598–611. Bibcode:2019Pbio...45..598E. doi:10.1017/pab.2019.29. ISSN 0094-8373. S2CID 209308325.
  115. ^ Nigenda-Morales, S.F.; Gompper, M.E.; Valenzuela-Galván, D.; Lay, A.R.; Kapheim, K.M.; Hass, C.; Booth-Binczik, S.D.; Binczik, G.A.; Hirsch, B.T.; McColgin, M.; Koprowski, J.L. (February 2019). "Phylogeographic and diversification patterns of the white-nosed coati (Nasua narica): Evidence for south-to-north colonization of North America". Molecular Phylogenetics and Evolution. 131: 149–163. doi:10.1016/j.ympev.2018.11.011. PMID 30468940. S2CID 53712751.
  116. ^ Koepfli, K.-P.; Gompper, M.E.; Eizirik, E.; Ho, C.-C.; Linden, L.; Maldonado, J.E.; Wayne, R.K. (June 2007). "Phylogeny of the Procyonidae (Mammalia: Carnivora): Molecules, morphology and the Great American Interchange". Molecular Phylogenetics and Evolution. 43 (3): 1076–1095. doi:10.1016/j.ympev.2006.10.003. PMID 17174109.
  117. ^ Eizirik, E. (2012). "A Molecular View on the Evolutionary History and Biogeography of Neotropical Carnivores (Mammalia, Carnivora)". Bones, Clones, and Biomes. University of Chicago Press. pp. 123–142. doi:10.7208/chicago/9780226649214.003.0007. ISBN 978-0-226-64919-1.
  118. ^ Marshall, L.G. (February 1994). "The Terror Birds of South America". Scientific American. 270 (2): 90–95. Bibcode:1994SciAm.270b..90M. doi:10.1038/scientificamerican0294-90. ISSN 0036-8733.
  119. ^ a b Angst, D.; Buffetaut, É. (2017). Palaeobiology of giant flightless birds. London: ISTE Press. ISBN 978-0-08-101143-0. OCLC 1012400051.
  120. ^ MacFadden, B.J.; Labs-Hochstein, J.; Hulbert, R.C.; Baskin, J.A. (2007). "Revised age of the late Neogene terror bird (Titanis) in North America during the Great American Interchange". Geology. 35 (2): 123. Bibcode:2007Geo....35..123M. doi:10.1130/G23186A.1. ISSN 0091-7613.
  121. ^ Alvarenga, H.; Jones, W.; Rinderknecht, A. (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.
  122. ^ Jones, W.; Rinderknecht, A.; Alvarenga, H.; Montenegro, F.; Ubilla, M. (June 2018). "The last terror birds (Aves, Phorusrhacidae): new evidence from the late Pleistocene of Uruguay". PalZ. 92 (2): 365–372. Bibcode:2018PalZ...92..365J. doi:10.1007/s12542-017-0388-y. S2CID 134344096.
  123. ^ a b Vizcaíno, S.F.; Fariña, R.A.; Zárate, M.A.; Bargo, M.S.; Schultz, P. (October 2004). "Palaeoecological implications of the mid-Pliocene faunal turnover in the Pampean Region (Argentina)". Palaeogeography, Palaeoclimatology, Palaeoecology. 213 (1–2): 101–113. Bibcode:2004PPP...213..101V. doi:10.1016/j.palaeo.2004.07.004. hdl:11336/81756.
  124. ^ Mitchell, K.J.; Bray, S.C.; Bover, P.; Soibelzon, L.; Schubert, B.W.; Prevosti, F.; Prieto, A.; Martin, F.; A., J.J.; Cooper, A. (2016-04-30). "Ancient mitochondrial DNA reveals convergent evolution of giant short-faced bears (Tremarctinae) in North and South America". Biology Letters. 12 (4): 20160062. doi:10.1098/rsbl.2016.0062. PMC 4881349. PMID 27095265.
  125. ^ Van Valkenburgh, B. (1991). "Iterative evolution of hypercarnivory in canids (Mammalia: Carnivora): evolutionary interactions among sympatric predators". Paleobiology. 17 (4): 340–362. Bibcode:1991Pbio...17..340V. doi:10.1017/s0094837300010691. ISSN 0094-8373. S2CID 251052044.
  126. ^ Schultz, P.H. (1998-12-11). "A 3.3-Ma Impact in Argentina and Possible Consequences". Science. 282 (5396): 2061–2063. Bibcode:1998Sci...282.2061S. doi:10.1126/science.282.5396.2061. PMID 9851923.
  127. ^ Moreno-Bernal, J.W.; Head, J.; Jaramillo, C.A. (2016-05-03). "Fossil Crocodilians from the High Guajira Peninsula of Colombia: Neogene faunal change in northernmost South America". Journal of Vertebrate Paleontology. 36 (3): e1110586. Bibcode:2016JVPal..36E0586M. doi:10.1080/02724634.2016.1110586. ISSN 0272-4634. S2CID 130332367.
  128. ^ Cidade, G. M.; Fortier, D.; Hsiou, A.S. (March 2019). "The crocodylomorph fauna of the Cenozoic of South America and its evolutionary history: a review". Journal of South American Earth Sciences. 90: 392–411. Bibcode:2019JSAES..90..392C. doi:10.1016/j.jsames.2018.12.026. S2CID 134902094.
  129. ^ Croft, D.A.; Gelfo, J.N.; López, G.M. (2020-05-30). "Splendid Innovation: The Extinct South American Native Ungulates". Annual Review of Earth and Planetary Sciences. 48 (1): annurev–earth–072619-060126. Bibcode:2020AREPS..48..259C. doi:10.1146/annurev-earth-072619-060126. ISSN 0084-6597. S2CID 213737574.
  130. ^ Gould, Stephen Jay (1980). The Panda's Thumb. W. W. Norton & Company. pp. 352 (see pp. 294–5). ISBN 978-0-393-01380-1. OCLC 6331415.
  131. ^ Wilson, Edward O. (1999). The Diversity of Life. W. W. Norton & Company. pp. 432 (see p. 130). ISBN 978-0-393-31940-8. OCLC 25508994.
  132. ^ a b Jerison, Harry J. (1973). Evolution of the Brain and Intelligence. New York and London: Academic Press. pp. (see pp. 320–339). ISBN 978-0-12-385250-2. OCLC 700636.
  133. ^ a b Radinsky, L. (1981). "Brain Evolution in Extinct South American Ungulates". Brain, Behavior and Evolution. 18 (4): 169–187. doi:10.1159/000121785. PMID 7284752.
  134. ^ a b Jerison, Harry J. (2007). "What Fossils Tell Us about the Evolution of the Neocortex" (PDF). In Kaas, J. H. (ed.). Evolution of Nervous Systems, Vol. 3. Vol. 3. New York and Oxford: Elsevier. pp. 1–12. doi:10.1016/B0-12-370878-8/00065-3. ISBN 978-0-12-392560-2.
  135. ^ Wroe, S.; Myers, T.; Seebacher, F; Kear, B.; Gillespie, A.; Crowther, M.; Salisbury, S. (2003). "An alternative method for predicting body mass: the case of the Pleistocene marsupial lion" (PDF). Paleobiology. 29 (3): 403–411. doi:10.1666/0094-8373(2003)029<0403:AAMFPB>2.0.CO;2. S2CID 85776991.
  136. ^ Quiroga, J. C.; Dozo, M. T. (1988). "The brain of Thylacosmilus atrox. Extinct South American saber-tooth carnivore marsupial". J. Hirnforschung. 29 (5): 573–586. PMID 3216103.
  137. ^ a b Radinsky, L. (1975). "Evolution of the Felid Brain". Brain, Behavior and Evolution. 11 (3–4): 214–254 (see p. 242). doi:10.1159/000123635. PMID 1181005.
  138. ^ Jerison, Harry J. (1973). Evolution of the Brain and Intelligence. New York and London: Academic Press. pp. (see p. 359). ISBN 978-0-12-385250-2. OCLC 700636.
  139. ^ Elgar, M. A.; Harvey, P. H. (1987). "Basal Metabolic Rates in Mammals: Allometry, Phylogeny and Ecology". Functional Ecology. 1 (1): 25–36. Bibcode:1987FuEco...1...25E. doi:10.2307/2389354. JSTOR 2389354.
  140. ^ Lovegrove, B. G. (August 2000). "The Zoogeography of Mammalian Basal Metabolic Rate". The American Naturalist. 156 (2): 201–219, see 214–215. doi:10.1086/303383. JSTOR 3079219. PMID 10856202. S2CID 4436119.
  141. ^ McNab, Brian K. (November 1980). "Energetics and the limits to the temperate distribution in armadillos". Journal of Mammalogy. 61 (4): 606–627, see p. 618. doi:10.2307/1380307. JSTOR 1380307.
  142. ^ Magazine, Smithsonian; Gerszak, Jennie Rothenberg Gritz,Fen Montaigne,Rafal. "The Story of How Humans Came to the Americas Is Constantly Evolving". Smithsonian Magazine. Retrieved 2022-03-12.{{cite web}}: CS1 maint: multiple names: authors list (link)
  143. ^ Agenbroad, L. D. (2004-06-26). "North American Proboscideans: Mammoths: The state of Knowledge, 2003". Quaternary International. 126–128: 73–92. Bibcode:2005QuInt.126...73A. doi:10.1016/j.quaint.2004.04.016.
  144. ^ Graham, R. W. (2001). "Late Quaternary Biogeography and Extinction of Proboscideans in North America" (PDF). In Cavarretta, G.; Gioia, P.; Mussi, M.; et al. (eds.). The World of Elephants (La Terra degli Elefanti) – Proceedings of the 1st International Congress (Atti del 1 Congrsso Internazionale), Rome October 16–20, 2001. Rome: Consiglio Nazionale delle Ricerche. pp. 707–709. ISBN 978-88-8080-025-5.
  145. ^ Martin, Paul S. (1973). "The Discovery of America: The first Americans may have swept the Western Hemisphere and decimated its fauna within 1000 years" (PDF). Science. 179 (4077): 969–974. Bibcode:1973Sci...179..969M. doi:10.1126/science.179.4077.969. PMID 17842155. S2CID 10395314.
  146. ^ Grayson, D. K.; Meltzer, D. J. (2003). "A requiem for North American overkill" (PDF). Journal of Archaeological Science. 30 (5): 585–593. Bibcode:2003JArSc..30..585G. doi:10.1016/S0305-4403(02)00205-4.
  147. ^ Fiedel, S.; Haynes, G. (January 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.
  148. ^ Grayson, D. K.; Meltzer, D. J. (2004). "North American overkill continued?" (PDF). Journal of Archaeological Science. 31 (1): 133–136. Bibcode:2004JArSc..31..133G. doi:10.1016/j.jas.2003.09.001. Retrieved 2009-06-12.
  149. ^ Haynes, G. (July 2007). "A review of some attacks on the overkill hypothesis, with special attention to misrepresentations and doubletalk". Quaternary International. 169–170: 84–94. Bibcode:2007QuInt.169...84H. doi:10.1016/j.quaint.2006.07.002.
  150. ^ 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 2014-05-25. Retrieved 2014-11-11.
  151. ^ Roberts, R. G.; Flannery, T. F.; Ayliffe, L. K.; Yoshida, H.; Olley, J. M.; Prideaux, G. J.; Laslett, G. M.; Baynes, A.; Smith, M. A.; Jones, R.; Smith, B. L. (2001-06-08). "New Ages for the Last Australian Megafauna: Continent-Wide Extinction About 46,000 Years Ago" (PDF). Science. 292 (5523): 1888–1892. Bibcode:2001Sci...292.1888R. doi:10.1126/science.1060264. PMID 11397939. S2CID 45643228. Retrieved 2011-08-26.
  152. ^ Johnson, C. N.; Alroy, J.; Beeton, N. J.; Bird, M. I.; Brook, B. W.; Cooper, A.; Gillespie, R.; Herrando-Pérez, S.; Jacobs, Z.; Miller, G. H.; Prideaux, G. J.; Roberts, R. G.; Rodríguez-Rey, M.; Saltré, F.; Turney, C. S. M.; Bradshaw, C. J. A. (10 February 2016). "What caused extinction of the Pleistocene megafauna of Sahul?". Proceedings of the Royal Society B: Biological Sciences. 283 (1824): 20152399. doi:10.1098/rspb.2015.2399. PMC 4760161. PMID 26865301.
  153. ^ Norton, C. J.; Kondo, Y.; Ono, A.; Zhang, Y.; Diab, M. C. (2009-05-23). "The nature of megafaunal extinctions during the MIS 3–2 transition in Japan". Quaternary International. 211 (1–2): 113–122. Bibcode:2010QuInt.211..113N. doi:10.1016/j.quaint.2009.05.002.
  154. ^ Burney, D. A.; Burney, L. P.; Godfrey, L. R.; Jungers, W. L.; Goodman, S. M.; Wright, H. T.; Jull, A. J. T. (July 2004). "A chronology for late prehistoric Madagascar". Journal of Human Evolution. 47 (1–2): 25–63. doi:10.1016/j.jhevol.2004.05.005. PMID 15288523.
  155. ^ Holdaway, R. N.; Jacomb, C. (2000-03-24). "Rapid Extinction of the Moas (Aves: Dinornithiformes): Model, Test, and Implications". Science. 287 (5461): 2250–2254. Bibcode:2000Sci...287.2250H. doi:10.1126/science.287.5461.2250. PMID 10731144.
  156. ^ Simmons, A. H. (1999). Faunal extinction in an island society: pygmy hippopotamus hunters of Cyprus. Interdisciplinary Contributions to Archaeology. Kluwer Academic/Plenum Publishers. pp. 229–230. doi:10.1007/b109876. ISBN 978-0-306-46088-3. OCLC 41712246.
  157. ^ Anderson, A.; Sand, C.; Petchey, F.; Worthy, T. H. (2010). "Faunal extinction and human habitation in New Caledonia: Initial results and implications of new research at the Pindai Caves". Journal of Pacific Archaeology. 1 (1): 89–109. hdl:10289/5404.
  158. ^ Diamond, Jared (2008-08-13). "Palaeontology: The last giant kangaroo". Nature. 454 (7206): 835–836. Bibcode:2008Natur.454..835D. doi:10.1038/454835a. PMID 18704074. S2CID 36583693.
  159. ^ Turney, C. S. M.; Flannery, T. F.; Roberts, R. G. (2008-08-21). "Late-surviving megafauna in Tasmania, Australia, implicate human involvement in their extinction". Proceedings of the National Academy of Sciences. 105 (34): 12150–12153. Bibcode:2008PNAS..10512150T. doi:10.1073/pnas.0801360105. PMC 2527880. PMID 18719103.
  160. ^ White, A. W.; Worthy, T. H.; Hawkins, S.; Bedford, S.; Spriggs, M. (2010-08-16). "Megafaunal meiolaniid horned turtles survived until early human settlement in Vanuatu, Southwest Pacific". Proc. Natl. Acad. Sci. USA. 107 (35): 15512–15516. Bibcode:2010PNAS..10715512W. doi:10.1073/pnas.1005780107. PMC 2932593. PMID 20713711.
  161. ^ Harrison, T. (2011). "Tortoises (Chelonii, Testudinidae)". Paleontology and Geology of Laetoli: Human Evolution in Context. Vertebrate Paleobiology and Paleoanthropology. Springer Science+Business Media. pp. 479–503. doi:10.1007/978-90-481-9962-4_17. ISBN 978-90-481-9961-7.
  162. ^ Hansen, D. M.; Donlan, C. J.; Griffiths, C. J.; Campbell, K. J. (April 2010). "Ecological history and latent conservation potential: large and giant tortoises as a model for taxon substitutions". Ecography. 33 (2): 272–284. Bibcode:2010Ecogr..33..272H. doi:10.1111/j.1600-0587.2010.06305.x.
  163. ^ Steadman, D. W.; Martin, P. S.; MacPhee, R. D. E.; Jull, A. J. T.; McDonald, H. G.; Woods, C. A.; Iturralde-Vinent, M.; Hodgins, G. W. L. (2005-08-16). "Asynchronous extinction of late Quaternary sloths on continents and islands". Proceedings of the National Academy of Sciences. 102 (33): 11763–11768. Bibcode:2005PNAS..10211763S. doi:10.1073/pnas.0502777102. PMC 1187974. PMID 16085711.
  164. ^ Cooke, S. B.; Dávalos, L. M.; Mychajliw, A. M.; Turvey, S. T.; Upham, N. S. (2017). "Anthropogenic Extinction Dominates Holocene Declines of West Indian Mammals". Annual Review of Ecology, Evolution, and Systematics. 48 (1): 301–327. doi:10.1146/annurev-ecolsys-110316-022754. S2CID 90558542.
  165. ^ Anderson, Paul K. (1995). "Competition, Predation, and the Evolution and Extinction of Steller's Sea Cow, Hydrodamalis gigas". Marine Mammal Science. 11 (3): 391–394. Bibcode:1995MMamS..11..391A. doi:10.1111/j.1748-7692.1995.tb00294.x.
  166. ^ Vartanyan, S. L.; Garutt, V. E.; Sher, A. V. (1993). "Holocene dwarf mammoths from Wrangel Island in the Siberian Arctic". Nature. 362 (6418): 337–349. Bibcode:1993Natur.362..337V. doi:10.1038/362337a0. PMID 29633990. S2CID 4249191.
  167. ^ Guthrie, R. Dale (2004). "Radiocarbon evidence of mid-Holocene mammoths stranded on an Alaskan Bering Sea island". Nature. 429 (6993): 746–749. Bibcode:2004Natur.429..746D. doi:10.1038/nature02612. PMID 15201907. S2CID 186242235.
  168. ^ Martin, P. S.; Steadman, D. W. (1999-06-30). "Prehistoric extinctions on islands and continents". In MacPhee, R. D. E (ed.). Extinctions in near time: causes, contexts and consequences. Advances in Vertebrate Paleobiology. Vol. 2. New York: Kluwer/Plenum. pp. 17–56, see p. 38. ISBN 978-0-306-46092-0. OCLC 41368299. Retrieved 2015-11-07.
  169. ^ a b c Vanzolini, P. E.; Heyer, W. R. (1985). "The American Herpetofauna and the Interchange" (PDF). In Stehli, F. G.; Webb, S. D. (eds.). The Great American Biotic Interchange. Topics in Geobiology, vol. 4. Plenum Press. pp. 475–487. ISBN 978-0-306-42021-4.
  170. ^ Pauly, G. B; Hillis, D. M.; Cannatella, D. C. (November 2004). "The history of a Nearctic colonization: Molecular phylogenetics and biogeography of the Nearctic toads (Bufo)". Evolution. 58 (11): 2517–2535. doi:10.1111/j.0014-3820.2004.tb00881.x. PMID 15612295. S2CID 10281132.
  171. ^ Faivovich, J.; Haddad, C. F. B.; Garcia, P. C. A.; Frost, D. R.; Campbell, J. A.; Wheeler, W. C. (2005-06-24). "Systematic review of the frog family Hylidae, with special reference to Hylinae: phylogenetic analysis and taxonomic revision" (PDF). Bulletin of the American Museum of Natural History. 294: 240 pages, see pp. 125–128. CiteSeerX 10.1.1.470.2967. doi:10.1206/0003-0090(2005)294[0001:srotff]2.0.co;2. S2CID 83925199. Archived from the original (PDF) on 2011-07-26. Retrieved 2010-03-07.
  172. ^ Heinicke, M. P.; Duellman, W. E.; Hedges, S. B. (2007-06-04). "Major Caribbean and Central American frog faunas originated by ancient oceanic dispersal". Proc. Natl. Acad. Sci. 104 (24): 10092–10097. Bibcode:2007PNAS..10410092H. doi:10.1073/pnas.0611051104. PMC 1891260. PMID 17548823.
  173. ^ a b c d e f g h i j Weir, J. T.; Bermingham, E.; Schluter, D. (2009-12-22). "The Great American Biotic Interchange in birds". Proceedings of the National Academy of Sciences. 106 (51): 21737–21742. Bibcode:2009PNAS..10621737W. doi:10.1073/pnas.0903811106. ISSN 0027-8424. PMC 2799814. PMID 19996168.
  174. ^ Ricklefs, R. E. (2002). "Splendid isolation: historical ecology of the South American passerine fauna". Journal of Avian Biology. 33 (3): 207–211. doi:10.1034/j.1600-048X.2002.330301.x.
  175. ^ a b Castaneda, O. C.; Miller, W. (July 2004). "Late Tertiary Terrestrial Mammals from Central Mexico and Their Relationship to South American Immigrants" (PDF). Revista Brasileira de Paleontologia. 7 (2): 249–261. doi:10.4072/rbp.2004.2.19. Retrieved 2009-01-29.
  176. ^ Arroyo-Cabrales, Joaquín; Polaco, Oscar J.; Johnson, Eileen; Ferrusquía-Villafranca, Ismael (2010-02-01). "A perspective on mammal biodiversity and zoogeography in the Late Pleistocene of México". Quaternary International. 212 (2): 187–197. Bibcode:2010QuInt.212..187A. doi:10.1016/j.quaint.2009.05.012.
  177. ^ Lundelius, E. L.; Bryant, V. M.; Mandel, R.; Thies, K. J.; Thoms, A. (January 2013). "The first occurrence of a toxodont (Mammalia, Notoungulata) in the United States" (PDF). Journal of Vertebrate Paleontology. 33 (1): 229–232. Bibcode:2013JVPal..33..229L. doi:10.1080/02724634.2012.711405. hdl:1808/13587. ISSN 0272-4634. S2CID 53601518.
  178. ^ Culver, M.; Johnson, W. E.; Pecon-Slattery, J.; O'Brien, S. J. (2000). "Genomic Ancestry of the American Puma (Puma concolor)" (PDF). Journal of Heredity. 91 (3): 186–197. doi:10.1093/jhered/91.3.186. PMID 10833043.[permanent dead link]
  179. ^ a b c Feller, A. E.; Hedges, S. B. (June 1998). "Molecular Evidence for the Early History of Living Amphibians". Molecular Phylogenetics and Evolution. 9 (3): 509–516. doi:10.1006/mpev.1998.0500. PMID 9667999.
  180. ^ Santos, J. C.; Coloma, L. A.; Summers, K.; Caldwell, J. P.; Ree, R.; Cannatella, D. C. (March 2009). "Amazonian Amphibian Diversity is Primarily Derived from Late Miocene Andean Lineages". PLOS Biology. 7 (3): 448–461 (see pp. 452, 454–455). doi:10.1371/journal.pbio.1000056. PMC 2653552. PMID 19278298.
  181. ^ Taplin, L. E.; Grigg, G. C. (1989). "Historical Zoogeography of the Eusuchian Crocodilians: A Physiological Perspective". American Zoologist. 29 (3): 885–901. doi:10.1093/icb/29.3.885.
  182. ^ Hastings, Alexander K.; Bloch, Jonathan I.; Jaramillo, Carlos A.; Rincon, Aldo F.; Macfadden, Bruce J. (5 March 2013). "Systematics and biogeography of crocodylians from the Miocene of Panama". Journal of Vertebrate Paleontology. 33 (2): 239–263. Bibcode:2013JVPal..33..239H. doi:10.1080/02724634.2012.713814. S2CID 83972694.
  183. ^ Pereira, S. L.; Baker, A. J. (2004). "Vicariant Speciation of Curassows (Aves, Cracidae): A Hypothesis Based On Mitochondrial DNA Phylogeny". The Auk. 121 (3): 682–694. doi:10.1642/0004-8038(2004)121[0682:VSOCAC]2.0.CO;2. S2CID 86320083.
  184. ^ Irestedt, Martin; Fjeldså, Jon; Johansson, Ulf S.; Ericson, Per G.P. (2002). "Systematic relationships and biogeography of the tracheophone suboscines (Aves: Passeriformes)". Molecular Phylogenetics and Evolution. 23 (3): 499–512. doi:10.1016/S1055-7903(02)00034-9. PMID 12099801.
  185. ^ "Scelidotherium in the Paleobiology Database". Fossilworks. Retrieved 17 December 2021.
  186. ^ Shaw, C. A.; McDonald, H. G. (1987-04-10). "First Record of Giant Anteater (Xenarthra, Myrmecophagidae) in North America". Science. 236 (4798): 186–188. Bibcode:1987Sci...236..186S. doi:10.1126/science.236.4798.186. JSTOR 1698387. PMID 17789783. S2CID 27112941.
  187. ^ Ford, S. (2006). "The Biogeographic History of Mesoamerican Primates". In Estrada, A.; Garber, P.A.; Pavelka, M.S.M.; Luecke, L. (eds.). New Perspectives in the Study of Mesoamerican Primates. Developments in Primatology: Progress and Prospects. New York: Springer. pp. 81–114. doi:10.1007/0-387-25872-8_4. ISBN 978-0-387-25854-6.
  188. ^ Helgen, K. M.; Pinto, M.; Kays, R.; Helgen, L.; Tsuchiya, M.; Quinn, A.; Wilson, D.; Maldonado, J. (2013-08-15). "Taxonomic revision of the olingos (Bassaricyon), with description of a new species, the Olinguito". ZooKeys (324): 1–83. Bibcode:2013ZooK..324....1H. doi:10.3897/zookeys.324.5827. PMC 3760134. PMID 24003317.
  189. ^ a b Schubert, B. W.; Chatters, J. C.; Arroyo-Cabrales, J.; Samuels, J. X.; Soibelzon, L. H.; Prevosti, F. J.; Widga, C.; Nava, A.; Rissolo, D.; Erreguerena, P. L. (2019). "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. 20190148.
  190. ^ Savage, J. M.; Wake, M. H.; Price, A. (2001). "Reevaluation of the Status of Taxa of Central American Caecilians (Amphibia: Gymnophiona), with Comments on Their Origin and Evolution". Copeia. 2001 (1): 52–64. doi:10.1643/0045-8511(2001)001[0052:ROTSOT]2.0.CO;2. S2CID 86086706.
  191. ^ a b Elmer, K. R.; Bonett, R. M.; Wake, D. B.; Lougheed, S. C. (2013-03-04). "Early Miocene origin and cryptic diversification of South American salamanders". BMC Evolutionary Biology. 13 (1): 59. Bibcode:2013BMCEE..13...59E. doi:10.1186/1471-2148-13-59. PMC 3602097. PMID 23497060.
  192. ^ Parra-Olea, G.; Garcia-Paris, M.; Wake, D. B. (2004). "Molecular diversification of salamanders of the tropical American genus Bolitoglossa (Caudata: Plethodontidae) and its evolutionary and biogeographical implications". Biological Journal of the Linnean Society. 81 (3): 325–346. doi:10.1111/j.1095-8312.2003.00303.x. Retrieved 2008-01-11.
  193. ^ Le, Minh; McCord, William P. (2008-07-25). "Phylogenetic relationships and biogeographical history of the genus Rhinoclemmys Fitzinger, 1835 and the monophyly of the turtle family Geoemydidae (Testudines: Testudinoidea)". Zoological Journal of the Linnean Society. 153 (4): 751–767. doi:10.1111/j.1096-3642.2008.00413.x.
  194. ^ Slowinski, J. B.; Keogh, J. S. (April 2000). "Phylogenetic Relationships of Elapid Snakes Based on Cytochrome b mtDNA Sequences". Molecular Phylogenetics and Evolution. 15 (1): 157–164. doi:10.1006/mpev.1999.0725. PMID 10764543.
  195. ^ Slowinski, J. B.; Boundy, J.; Lawson, R. (June 2001). "The Phylogenetic Relationships of Asian Coral Snakes (Elapidae: Calliophis and Maticora) Based on Morphological and Molecular Characters". Herpetologica. 57 (2): 233–245. JSTOR 3893186.
  196. ^ Place, A. J.; Abramson, C. I. (2004). "A Quantitative Analysis of the Ancestral Area of Rattlesnakes". Journal of Herpetology. 38 (1): 152–156. doi:10.1670/103-03N. S2CID 86252575.
  197. ^ Parkinson, C. L. (1999). "Molecular Systematics and Biogeographical History of Pit Vipers as Determined by Mitochondrial Ribosomal DNA Sequences". Copeia. 1999 (3): 576–586. doi:10.2307/1447591. JSTOR 1447591.
  198. ^ Ericson, P. G. P.; Christidis, L.; Cooper, A.; Irestedt, M.; Jackson, J.; Johansson, U. S.; Norman, J. A. (2002-02-07). "A Gondwanan origin of passerine birds supported by DNA sequences of the endemic New Zealand wrens". Proceedings of the Royal Society B: Biological Sciences. 269 (1488): 235–241. doi:10.1098/rspb.2001.1877. ISSN 0962-8452. PMC 1690883. PMID 11839192.
  199. ^ Dacosta, J. M.; Klicka, J. (2008-02-21). "The Great American Interchange in birds: a phylogenetic perspective with the genus Trogon". Molecular Ecology. 17 (5): 1328–1343. Bibcode:2008MolEc..17.1328D. doi:10.1111/j.1365-294X.2007.03647.x. PMID 18302692. S2CID 25090736.
  200. ^ Emslie, Steven D. (22 June 1988). "The Fossil History and Phylogenetic Relationships of Condors (Ciconiiformes: Vulturidae) in the New World". Journal of Vertebrate Paleontology. 8 (2): 212–228. Bibcode:1988JVPal...8..212E. doi:10.1080/02724634.1988.10011699. JSTOR 4523192.
  201. ^ a b Emslie, Steven D. (July 1988). "An early condor-like vulture from North America" (PDF). The Auk. 105 (3): 529–535. doi:10.1093/auk/105.3.529.
  202. ^ a b Stucchi, Marcelo; Emslie, Steven D. (February 2005). "A New Condor (Ciconiiformes, Vulturidae) From The Late Miocene/Early Pliocene Pisco Formation, Peru". The Condor. 107 (1): 107–113. doi:10.1650/7475. S2CID 85805971.
  203. ^ a b Weinstock, J.; et al. (2005). "Evolution, systematics, and phylogeography of Pleistocene horses in the New World: a molecular perspective". PLOS Biology. 3 (8): e241. doi:10.1371/journal.pbio.0030241. PMC 1159165. PMID 15974804.
  204. ^ a b c d Orlando, L.; Male, D.; Alberdi, M. T.; Prado, J. L.; Prieto, A.; Cooper, A.; Hänni, C. (2008). "Ancient DNA Clarifies the Evolutionary History of American Late Pleistocene Equids". Journal of Molecular Evolution. 66 (5): 533–538. Bibcode:2008JMolE..66..533O. doi:10.1007/s00239-008-9100-x. PMID 18398561. S2CID 19069554.
  205. ^ Polaco, O. J.; Arroyo-Cabrales, J.; Corona-M., E.; López-Oliva, J. G. (2001). "The American Mastodon Mammut americanum in Mexico" (PDF). In Cavarretta, G.; Gioia, P.; Mussi, M.; et al. (eds.). The World of Elephants – Proceedings of the 1st International Congress, Rome October 16–20, 2001. Rome: Consiglio Nazionale delle Ricerche. pp. 237–242. ISBN 978-88-8080-025-5.
  206. ^ Soibelzon, L. H.; Tonni, E. P.; Bond, M. (October 2005). "The fossil record of South American short-faced bears (Ursidae, Tremarctinae)". Journal of South American Earth Sciences. 20 (1–2): 105–113. Bibcode:2005JSAES..20..105S. doi:10.1016/j.jsames.2005.07.005.
  207. ^ "Canis dirus (dire wolf)". Paleobiology Database. Retrieved 2013-01-27.
  208. ^ Hodnett, John-Paul M.; Mead, Jim I.; Baez, A. (2009). "Dire Wolf, Canis dirus (Mammalia; Carnivora; Canidae), from the Late Pleistocene (Rancholabrean) of East-Central Sonora, Mexico". The Southwestern Naturalist. 54 (1): 74–81. doi:10.1894/CLG-12.1. S2CID 84760786.
  209. ^ Berta, A. (November 1988). Quaternary Evolution and Biogeography of the Large South American Canidae (Mammalia: Carnivora). University of California Publications in Geological Sciences. Vol. 132. University of California Press. ISBN 978-0-520-09960-9. Retrieved 2013-01-28.
  210. ^ "New Clues To Extinct Falklands Wolf Mystery". EurekAlert. Science Daily. 2009-11-03. Retrieved 2011-09-03.
  211. ^ Cooper, A.; Mena, F.; Austin, J. J.; Soubrier, J.; Prevosti, F.; Prates, L.; Trejo, V. (2013). "The origins of the enigmatic Falkland Islands wolf". Nature Communications. 4: 1552. Bibcode:2013NatCo...4.1552A. doi:10.1038/ncomms2570. hdl:2440/74885. PMID 23462995. S2CID 205315969.
  212. ^ Mones, A.; Rinderknecht, A. (2004). "The First South American Homotheriini (Mammalia: Carnivora: Felidae)" (PDF). Comunicaciones Paleontologicas Museo Nacional de Historia Natural y Anthropologia. 2 (35): 201–212. Retrieved 2013-02-08.
  213. ^ Sanchez, Fabiola (2008-08-21). "Saber-toothed cat fossils discovered". Associated Press. Retrieved 2017-05-07.
  214. ^ Orozco, José (2008-08-22). "Sabertooth Cousin Found in Venezuela Tar Pit – A First". National Geographic News. National Geographic Society. Archived from the original on January 4, 2013. Retrieved 2008-08-30.
  215. ^ a b Rincón, Ascanio D.; Prevosti, Francisco J.; Parra, Gilberto E. (2011). "New saber-toothed cat records (Felidae: Machairodontinae) for the Pleistocene of Venezuela, and the Great American Biotic Interchange". Journal of Vertebrate Paleontology. 31 (2): 468–478. Bibcode:2011JVPal..31..468R. doi:10.1080/02724634.2011.550366. hdl:11336/69016. S2CID 129693331.
  216. ^ Manzuetti, A.; Perea, D.; Ubilla, M.; Rinderknecht, A. (2018). "First record of Smilodon fatalis Leidy, 1868 (Felidae, Machairodontinae) in the extra-Andean region of South America (late Pleistocene, Sopas Formation), Uruguay: Taxonomic and paleobiogeographic implications". Quaternary Science Reviews. 180: 57–62. Bibcode:2018QSRv..180...57M. doi:10.1016/j.quascirev.2017.11.024.
  217. ^ Kurtén, B.; Anderson, E. (1980). Pleistocene Mammals of North America. Columbia University Press. ISBN 978-0231037334. OCLC 759120597.
  218. ^ Chimento, N. R.; Agnolin, F. L. (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.
  219. ^ Seymour, K. 2015. Perusing Talara: Overview of the Late Pleistocene fossils from the tar seeps of Peru Archived 2018-10-01 at the Wayback Machine. Natural History Museum of Los Angeles County Science Series, 42: 97-109

Further reading edit