The Paleogene Period (IPA: /ˈpli.ən, -li.-, ˈpæli-/ PAY-lee-ə-jeen, -⁠lee-oh-, PAL-ee-; also spelled Palaeogene or Palæogene) is a geologic period and system that spans 43 million years from the end of the Cretaceous Period 66 Ma (million years ago) to the beginning of the Neogene Period 23.03 Ma. It is the first period of the Cenozoic Era and is divided into the Paleocene, Eocene, and Oligocene epochs. The earlier term Tertiary Period was used to define the time now covered by the Paleogene Period and subsequent Neogene Period; despite no longer being recognized as a formal stratigraphic term, "Tertiary" still sometimes remains in informal use.[5] Paleogene is often abbreviated "Pg", although the United States Geological Survey uses the abbreviation "Pe" for the Paleogene on the Survey's geologic maps.[6][7]

Paleogene
66.0 – 23.03 Ma
Map of Earth as it appeared 50 million years ago during the Eocene, Ypresian stage[citation needed]
Chronology
Etymology
Name formalityFormal
Alternate spelling(s)Palaeogene, Palæogene
Usage information
Celestial bodyEarth
Regional usageGlobal (ICS)
Time scale(s) usedICS Time Scale
Definition
Chronological unitPeriod
Stratigraphic unitSystem
Time span formalityFormal
Lower boundary definitionIridium enriched layer associated with a major meteorite impact and subsequent K-Pg extinction event.
Lower boundary GSSPEl Kef Section, El Kef, Tunisia
36°09′13″N 8°38′55″E / 36.1537°N 8.6486°E / 36.1537; 8.6486
Lower GSSP ratified1991[3]
Upper boundary definition
Upper boundary GSSPLemme-Carrosio Section, Carrosio, Italy
44°39′32″N 8°50′11″E / 44.6589°N 8.8364°E / 44.6589; 8.8364
Upper GSSP ratified1996[4]
Atmospheric and climatic data
Mean atmospheric O2 contentc. 26 vol %
(125 % of modern)
Mean atmospheric CO2 contentc. 500 ppm
(1.8 times pre-industrial)
Mean surface temperaturec. 18 °C
(4.5 °C above pre-industrial)

During the Paleogene period, mammals continued to diversify from relatively small, simple forms into a large group of diverse animals in the wake of the Cretaceous–Paleogene extinction event that ended the preceding Cretaceous Period.[8]

The Period is marked by considerable changes in climate from the Paleocene–Eocene Thermal Maximum, through global cooling during the Eocene to the first appearance of permanent ice sheets in the Antarctic at the beginning of the Oligocene.[9]

Geology

edit

Stratigraphy

edit

The Paleogene is divided into three series/epochs: the Paleocene, Eocene, and Oligocene. These stratigraphic units can be defined globally or regionally. For global stratigraphic correlation, the International Commission on Stratigraphy (ICS) ratify global stages based on a Global Boundary Stratotype Section and Point (GSSP) from a single formation (a stratotype) identifying the lower boundary of the stage.[10]

Paleocene

edit

The Paleocene is the first series/epoch of the Paleogene and lasted from 66.0 Ma to 56.0 Ma. It is divided into three stages: the Danian 66.0 - 61.6 Ma; Selandian 61.6 - 59.2 Ma; and, Thanetian 59.2 - 56.0 Ma.[11] The GSSP for the base of the Cenozoic, Paleogene and Paleocene is at Oued Djerfane, west of El Kef, Tunisia. It is marked by an iridium anomaly produced by an asteroid impact, and is associated with the Cretaceous–Paleogene extinction event. The boundary is defined as the rusty colored base of a 50 cm thick clay, which would have been deposited over only a few days. Similar layers are seen in marine and continental deposits worldwide. These layers include the iridium anomaly, microtektites, nickel-rich spinel crystals and shocked quartz, all indicators of a major extraterrestrial impact. The remains of the crater are found at Chicxulub on the Yucatan Peninsula in Mexico. The extinction of the non-avian dinosaurs, ammonites and dramatic changes in marine plankton and many other groups of organisms, are also used for correlation purposes.[11]

Eocene

edit

The Eocene is the second series/epoch of the Paleogene, and lasted from 56.0 Ma to 33.9 Ma. It is divided into four stages: the Ypresian 56.0 Ma to 47.8 Ma; Lutetian 47.8 Ma to 41.2 Ma; Bartonian 41.2 Ma to 37.71 Ma; and, Priabonian 37.71 Ma to 33.9 Ma. The GSSP for the base of the Eocene is at Dababiya, near Luxor, Egypt and is marked by the start of a significant variation in global carbon isotope ratios, produced by a major period of global warming. The change in climate was due to a rapid release of frozen methane clathrates from seafloor sediments at the beginning of the Paleocene-Eocene thermal maximum (PETM).[11]

Oligocene

edit

The Oligocene is the third and youngest series/epoch of the Paleogene, and lasted from 33.9 Ma to 23.03 Ma. It is divided into two stages: the Rupelian 33.9 Ma to 27.82 Ma; and, Chattian 27.82 - 23.03 Ma. The GSSP for the base of the Oligocene is at Massignano, near Ancona, Italy. The extinction the hantkeninid planktonic foraminifera is the key marker for the Eocene-Oligocene boundary, which was a time of climate cooling that led to widespread changes in fauna and flora.[11]

Palaeogeography

edit

The final stages of the breakup of Pangaea occurred during the Paleogene as Atlantic Ocean rifting and seafloor spreading extended northwards, separating the North America and Eurasian plates, and Australia and South America rifted from Antarctica, opening the Southern Ocean. Africa and India collided with Eurasia forming the Alpine-Himalayan mountain chains and the western margin of the Pacific Plate changed from a divergent to convergent plate boundary.[12]

Alpine - Himalayan Orogeny

edit

Alpine Orogeny

edit

The Alpine Orogeny developed in response to the collision between the African and Eurasian plates during the closing of the Neotethys Ocean and the opening of the Central Atlantic Ocean. The result was a series of arcuate mountain ranges, from the Tell-Rif-Betic cordillera in the western Mediterranean through the Alps, Carpathians, Apennines, Dinarides and Hellenides to the Taurides in the east.[13][12]

From the Late Cretaceous into the early Paleocene, Africa began to converge with Eurasia. The irregular outlines of the continental margins, including the Adriatic promontory (Adria) that extended north from the African Plate, led to the development of several short subduction zones, rather than one long system.[13] In the western Mediterranean, the European Plate was subducted southwards beneath the African Plate, whilst in the eastern Mediterranean, Africa was subducted beneath Eurasia along a northward dipping subduction zone.[12][14] Convergence between the Iberian and European plates led to the Pyrenean Orogeny[15] and, as Adria pushed northwards the Alps and Carpathian orogens began to develop.[16][14]

 
Present day tectonic map of southern Europe, North Africa and the Middle East, showing structures of the western Alpine-Himalayan orogenic belt.

The collision of Adria with Eurasia in the early Palaeocene was followed by a  c.10 million year pause in the convergence of Africa and Eurasia, connected with the onset of the opening of the North Atlantic Ocean as Greenland rifted from the Eurasian Plate in the Palaeocene.[16] Convergence rates between Africa and Eurasia increased again in the early Eocene and the remaining oceanic basins between Adria and Europe closed.[13][17]

Between about 40 and 30 Ma, subduction began along the western Mediterranean arc of the Tell, Rif, Betic and Apennine mountain chains. The rate of convergence was less than the subduction rate of the dense lithosphere of the western Mediterranean and roll-back of the subducting slab led to the arcuate structure of these mountain ranges.[13][15]

In the eastern Mediterranean, c. 35 Ma, the Anatolide-Tauride platform (northern part of Adria) began to enter the trench leading to the development of the Dinarides, Hellenides and Tauride mountain chains as the passive margin sediments of Adria were scrapped off onto the Eurasia crust during subduction.[13][18]

Zagros Mountains

edit

The Zagros mountain belt stretches for c. 2000  km from the eastern border of Iraq to the Makran coast in southern Iran. It formed as a result of the convergence and collision of the Arabian and Eurasian plates as the Neotethys Ocean closed and is composed sediments scrapped from the descending Arabian Plate.[19][20]

From the Late Cretaceous, a volcanic arc developed on the Eurasia margin as the Neotethys crust was subducted beneath it. A separate intra-oceanic subduction zone in the Neotethys resulted in the obuction of ocean crust onto the Arabian margin in the Late Cretaceous to Paleocene, with break-off of the subducted oceanic plate close to the Arabian margin occurring during the Eocene.[19][20] Continental collision began during the Eocene c. 35 Ma and continued into the Oligocene to c. 26 Ma.[19][20]

Himalayan Orogeny

edit
 
Map showing the northwards drift of the Indian continent between 71 and 0 Ma. The leading edge of Greater India (not shown on the map) collided with the Eurasian plate c. 55 Ma, whilst India itself still lay to the south. (From: Dèzes, 1999)

The Indian continent rifted from Madagascar at c. 83 Ma and drifted rapidly (c. 18 cm/yr in the Paleocene) northwards towards the southern margin of Eurasia. A rapid decrease in velocity to c. 5 cm/yr in the early Eocene records the collision of the Tethyan (Tibetan) Himalayas, the leading edge of Greater India, with the Lhasa Terrane of Tibet (southern Eurasian margin), along the Indus-Yarling-Zangbo suture zone.[12][21] To the south of this zone, the Himalaya are composed of metasedimentary rocks scraped off the now subducted Indian continental crust and mantle lithosphere as the collision progressed.[12]

Palaeomagnetic data place the present day Indian continent further south at the time of collision and decrease in plate velocity, indicating the presence of a large region to the north of India that has now been subducted beneath the Eurasian Plate or incorporated into the mountain belt. This region, known as Greater India, formed by extension along the northern margin of India during the opening of the Neotethys. The Tethyan Himalaya block lay along its northern edge, with the Neotethys Ocean lying between it and southern Eurasia.[12][22]

Debate about the amount of deformation seen in the geological record in the India–Eurasia collision zone versus the size of Greater India, the timing and nature of the collision relative to the decrease in plate velocity, and explanations for the unusually high velocity of the Indian plate have led to several models for Greater India: 1) A Late Cretaceous to early Paleocene subduction zone may have lain between India and Eurasia in the Neotethys, dividing the region into two plates, subduction was followed by collision of India with Eurasia in the middle Eocene. In this model Greater India would have been less than 900 km wide;[22] 2) Greater India may have formed a single plate, several thousand kilometres wide, with the Tethyan Himalaya microcontinent separated from the Indian continent by an oceanic basin. The microcontinent collided with southern Eurasia c. 58 Ma (late Paleocene), whilst the velocity of the plate did not decrease until c. 50 Ma when subduction rates dropped as young, oceanic crust entered the subduction zone;[23] 3) This model assigns older dates to parts of Greater India, which changes its paleogeographic position relative to Eurasia and creates a Greater India formed of extended continental crust 2000 - 3000 km wide.[24]

South East Asia

edit

The Alpine-Himalayan Orogenic Belt in Southeast Asia extends from the Himalayas in India through Myanmar (West Burma block) Sumatra, Java to West Sulawesi.[25]

During the Late Cretaceous to Paleogene, the northward movement of the Indian Plate led to the highly oblique subduction of the Neotethys along the edge of the West Burma block and the development of a major north-south transform fault along the margin of Southeast Asia to the south.[26][25] Between c. 60 and 50 Ma, the leading northeastern edge of Greater India collided with the West Burma block resulting in deformation and metamorphism.[26] During the middle Eocene, north-dipping subduction resumed along the southern edge of Southeast Asia, from west Sumatra to West Sulawesi, as the Australian Plate drifted slowly northwards.[25]

Collision between India and the West Burma block was complete by the late Oligocene. As the India-Eurasia collision continued, movement of material away from the collision zone was accommodated along, and extended, the already existing major strike slip systems of the region.[26]

Atlantic Ocean

edit

During the Paleocene, seafloor spreading along the Mid-Atlantic Ridge propagated from the Central Atlantic northwards between North America and Greenland in the Labrador Sea (c. 62 Ma) and Baffin Bay (c. 57 Ma), and, by the early Eocene (c. 54 Ma), into the northeastern Atlantic between Greenland and Eurasia.[12][27] Extension between North America and Eurasia, also in the early Eocene, led to the opening of the Eurasian Basin across the Arctic, which was linked to the Baffin Bay Ridge and Mid-Atlantic Ridge to the south via major strike slip faults.[12][28]

From the Eocene and into the early Oligocene, Greenland acted as an independent plate moving northwards and rotating anticlockwise. This led to compression across the Canadian Arctic Archipelago, Svalbard and northern Greenland resulting in the Eureka Orogeny.[12][28] From c. 47 Ma, the eastern margin of Greenland was cut by the Reykjanes Ridge (the northeastern branch of the Mid-Atlantic Ridge) propagating northwards and splitting off the Jan Mayen microcontinent.[12]

After c. 33 Ma seafloor spreading in Labrador Sea and Baffin Bay gradually ceased and seafloor spreading focused along the northeast Atlantic. By the late Oligocene, the plate boundary between North America and Eurasia was established along the Mid-Atlantic Ridge, with Greenland attached to the North American plate again, and the Jan Mayen microcontinent part of the Eurasian Plate, where its remains now lie to the east and possibly beneath the southeast of Iceland.[12][28]

North Atlantic Large Igneous Province

edit
 
Three Paleogene-aged basaltic lava flows on the Isle of Staffa, Scotland. As they have cooled the flows have formed rock with different characteristics. The middle layer shows spectacular columnar jointing.

The North Atlantic Igneous Province stretches across the Greenland and northwest European margins and is associated with the proto-Icelandic mantle plume, which rose beneath the Greenland lithosphere at c. 65 Ma.[28] There were two main phases of volcanic activity with peaks at c. 60 Ma and c. 55 Ma. Magmatism in the British and Northwest Atlantic volcanic provinces occurred mainly in the early Palaeocene, the latter associated with an increased spreading rate in the Labrador Sea, whilst northeast Atlantic magmatism occurred mainly during the early Eocene and is associated with a change in the spreading direction in the Labrador Sea and the northward drift of Greenland. The locations of the magmatism coincide with the intersection of propagating the rifts and large-scale, pre-existing lithospheric structures, which acted as channels to the surface for the magma.[28][29]

The arrival of the proto-Iceland plume has been considered the driving mechanism for rifting in the North Atlantic. However, that rifting and initial seafloor spreading occurred prior to the arrival of the plume, large scale magmatism occurred at a distance to rifting, and that rifting propagated towards, rather than away from the plume, has led to the suggestion the plume and associated magmatism may have been a result, rather than a cause, of the plate tectonic forces that led to the propagation of rifting from the Central to the North Atlantic.[28][29]

Americas

edit

North America

edit

Mountain building continued along the North America Cordillera in response to subduction of the Farallon plate beneath the North American Plate. Along the central section of the North American margin, crustal shortening of the Cretaceous to Paleocene Sevier Orogen lessened and deformation moved eastward. The decreasing dip of the subducting Farallon Plate led to a flat-slab segment that increased friction between this and the base of the North American Plate. The resulting Laramide Orogeny, which began the development of the Rocky Mountains, was a broad zone of thick-skinned deformation, with faults extending to mid-crustal depths and the uplift of basement rocks that lay to the east of the Sevier belt, and more than 700km from the trench.[30][31] With the Laramide uplift the Western Interior Seaway was divided and then retreated.[30]

During the mid to late Eocene (50–35 Ma), plate convergence rates decreased and the dip of the Farallon slab began to steepen. Uplift ceased and the region largely levelled by erosion. By the Oligocene, convergence gave way to extension, rifting and widespread volcanism across the Laramide belt.[30][31]

South America

edit

Ocean-continent convergence accommodated by east dipping subduction zone of the Farallon Plate beneath the western edge of South America continued from the Mesozoic.[32]

Over the Paleogene, changes in plate motion and episodes of regional slab shallowing and steepening resulted in variations in the magnitude of crustal shortening and amounts of magmatism along the length of the Andes.[32] In the Northern Andes, an oceanic plateau with volcanic arc was accreted during the latest Cretaceous and Paleocene, whilst the Central Andes were dominated by the subduction of oceanic crust and the Southern Andes were impacted by the subduction of the Farallon-East Antarctic ocean ridge.[33][34]

Caribbean

edit

The Caribbean Plate is largely composed of oceanic crust of the Caribbean Large Igneous Province that formed during the Late Cretaceous.[34] During the Late Cretaceous to Paleocene, subduction of Atlantic crust was established along its northern margin, whilst to the southwest, an island arc collided with the northern Andes forming an east dipping subduction zone where Caribbean lithosphere was subducted beneath the South American margin.[35]

During the Eocene (c. 45 Ma), subduction of the Farallon Plate along the Central American subduction zone was (re)established.[34] Subduction along the northern section of the Caribbean volcanic arc ceased as the Bahamas carbonate platform collided with Cuba and was replaced by strike-slip movements as a transform fault, extending from the Mid-Atlantic Ridge, connected with the northern boundary of the Caribbean Plate. Subduction now focused along the southern Caribbean arc (Lesser Antilles).[34][36]

By the Oligocene, the intra-oceanic Central American volcanic arc began to collide with northwestern South American.[35]

Pacific Ocean

edit

At the beginning of the Paleogene, the Pacific Ocean consisted of the Pacific, Farallon, Kula and Izanagi plates. The central Pacific Plate grew by seafloor spreading as the other three plates were subducted and broken up. In the southern Pacific, seafloor spreading continued from the Late Cretaceous across the Pacific–Antarctic, Pacific-Farallon and Farallon–Antarctic mid ocean ridges.[12]

The Izanagi-Pacific spreading ridge lay nearly parallel to the East Asian subduction zone and between 60–50 Ma the spreading ridge began to be subducted. By c. 50 Ma, the Pacific Plate was no longer surrounded by spreading ridges, but had a subduction zone along its western edge. This changed the forces acting on the Pacific Plate and led to a major reorganisation of plate motions across the entire Pacific region.[37] The resulting changes in stress between the Pacific and Philippine Sea plates initiated subduction along the Izu-Bonin-Mariana and Tonga-Kermadec arcs.[37][34]

Subduction of the Farallon Plate beneath the American plates continued from the Late Cretaceous.[12] The Kula-Farallon spreading ridge lay to its north until the Eocene (c. 55 Ma), when the northern section of the plate split forming the Vancouver/Juan de Fuca Plate.[34] In the Oligocene (c. 28 Ma), the first segment of the Pacific–Farallon spreading ridge entered the North American subduction zone near Baja California[38] leading to major strike-slip movements and the formation of the San Andreas Fault.[12] At the Paleogene-Neogene boundary, spreading ceased between the Pacific and Farallon plates and the Farallon Plate split again forming the present date Nazca and Cocos plates.[34][38]

The Kula Plate lay between Pacific Plate and North America. To the north and northwest it was being subducted beneath the Aleutian trench.[12][34] Spreading between the Kula and Pacific and Farallon plates ceased c. 40 Ma and the Kula Plate became part of the Pacific Plate.[12][34]

Hawaii hotspot

edit

The Hawaiian-Emperor seamount chain formed above the Hawaiian hotspot. Originally thought to be stationary within the mantle, the hotspot is now considered to have drifted south during the Paleocene to early Eocene, as the Pacific Plate moved north. At c. 47 Ma, movement of the hotspot ceased and the Pacific Plate motion changed from northward to northwestward in response to the onset of subduction along its western margin. This resulted in a 60 degree bend in the seamount chain. Other seamount chains related to hotspots in the South Pacific show a similar change in orientation at this time.[39]

Antarctica

edit

Slow seafloor spreading continued between Australia and East Antarctica. Shallow water channels probably developed south of Tasmania opening the Tasmanian Passage in the Eocene and deep ocean routes opening from the mid Oligocene. Rifting between the Antarctic Peninsula and the southern tip of South America formed the Drake Passage and opened the Southern Ocean also during this time, completing the breakup of Gondwana. The opening of these passages and the creation of the Southern Ocean established the Antarctic Circumpolar Current. Glaciers began to build across the Antarctica continent that now lay isolated in the south polar region and surrounded by cold ocean waters. These changes contributed to the fall in global temperatures and the beginning of icehouse conditions.[30]

 
Paleogene flood basalts on the Ethiopian Plateau with the Afar Depression in the background.

Red Sea and East Africa

edit

Extensional stresses from the subduction zone along the northern Neotethys resulted in rifting between Africa and Arabia, forming the Gulf of Aden in the late Eocene.[40] To the west, in the early Oligocene, flood basalts erupted across Ethiopia, northeast Sudan and southwest Yemen as the Afar mantle plume began to impact the base of the African lithosphere.[12][40] Rifting across the southern Red Sea began in the mid Oligocene, and across the central and northern Red Sea regions in the late Oligocene and early Miocene.[40]

Climate

edit

The global climate of the Paleogene began with the brief but intense "impact winter" caused by the Chicxulub impact. This cold period was terminated by an abrupt warming. After temperatures stabilised, the steady cooling and drying of the Late Cretaceous-Early Paleogene Cool Interval (LKEPCI) that had spanned the last two ages of the Late Cretaceous continued.[9] About 62.2 Mya, the Latest Danian Event, a hyperthermal event, took place.[41][42][43] About 59 Ma, the LKEPCI was brought to an end by the Thanetian Thermal Event, a change from the relative cool of the Early and Middle Palaeocene and the beginning of an intense supergreenhouse effect.[9]

According to a study published in 2018, from about 56 to 48 Mya, annual air temperatures over land and at mid-latitude averaged about 23–29 °C (± 4.7 °C), which is 5–10 °C warmer than most previous estimates.[44][45][46] For comparison, this was 10 to 15 °C greater than the current annual mean temperatures in these areas.[46] At the Palaeocene-Eocene boundary, the Paleocene–Eocene Thermal Maximum (PETM) occurred,[47] one of the warmest times of the Phanerozoic eon, during which global mean surface temperatures increased to 31.6.[48] It was followed by the less severe Eocene Thermal Maximum 2 (ETM2) about 53.69 Ma.[49] Eocene Thermal Maximum 3 (ETM3) occurred about 53 Ma. The Early Eocene Climatic Optimum was brought to an end by the Azolla event, a change of climate about 48.5 Mya, believed to have been caused by a proliferation of aquatic ferns from the genus Azolla, resulting in the sequestering of large amounts of carbon dioxide by those plants. From this time until about 34 Mya, there was a slow cooling trend known as the Middle-Late Eocene Cooling (MLEC).[9] Approximately 41.5 Mya, this cooling was interrupted temporarily by the Middle Eocene Climatic Optimum (MECO).[50] Then, about 39.4 Mya, a temperature decrease termed the Late Eocene Cool Event (LECE) is detected in the oxygen isotope record.[9] A rapid decrease of global temperatures and formation of continental glaciers on Antarctica marked the end of the Eocene.[51] This sudden cooling was caused partly by the formation of the Antarctic Circumpolar Current,[52] which significantly lowered oceanic water temperatures.[53]

During the earliest Oligocene occurred the Early Oligocene Glacial Maximum (Oi1), which lasted for about 200,000 years.[54] After Oi1, global mean surface temperature continued to decrease gradually during the Rupelian Age. [9] Another major cooling event occurred at the end of the Rupelian; its most likely cause was extreme biological productivity in the Southern Ocean fostered by tectonic reorganisation of ocean currents and an influx of nutrients from Antarctica.[55] In the Late Oligocene, global temperatures began to warm slightly, though they continued to be significantly lower than during the previous epochs of the Paleogene and polar ice remained.[9]

Flora and fauna

edit

Tropical taxa diversified faster than those at higher latitudes after the Cretaceous–Paleogene extinction event, resulting in the development of a significant latitudinal diversity gradient.[56] Mammals began a rapid diversification during this period. After the Cretaceous–Paleogene extinction event, which saw the demise of the non-avian dinosaurs, mammals began to evolve from a few small and generalized forms into most of the modern varieties we see presently. Some of these mammals evolved into large forms that dominated the land, while others became capable of living in marine, specialized terrestrial, and airborne environments. Those that adapted to the oceans became modern cetaceans, while those that adapted to trees became primates, the group to which humans belong. Birds, extant dinosaurs which were already well established by the end of the Cretaceous, also experienced adaptive radiation as they took over the skies left empty by the now extinct pterosaurs. Some flightless birds such as penguins, ratites, and terror birds also filled niches left by the hesperornithes and other extinct dinosaurs.

Pronounced cooling in the Oligocene resulted in a massive floral shift, and many extant modern plants arose during this time. Grasses and herbs, such as Artemisia, began to proliferate, at the expense of tropical plants, which began to decrease. Conifer forests developed in mountainous areas. This cooling trend continued, with major fluctuation, until the end of the Pleistocene period.[57] This evidence for this floral shift is found in the palynological record.[58]

See also

edit

References

edit
  1. ^ Zachos, J. C.; Kump, L. R. (2005). "Carbon cycle feedbacks and the initiation of Antarctic glaciation in the earliest Oligocene". Global and Planetary Change. 47 (1): 51–66. Bibcode:2005GPC....47...51Z. doi:10.1016/j.gloplacha.2005.01.001.
  2. ^ "International Chronostratigraphic Chart" (PDF). International Commission on Stratigraphy.
  3. ^ Molina, Eustoquio; Alegret, Laia; Arenillas, Ignacio; José A. Arz; Gallala, Njoud; Hardenbol, Jan; Katharina von Salis; Steurbaut, Etienne; Vandenberghe, Noel; Dalila Zaghibib-Turki (2006). "The Global Boundary Stratotype Section and Point for the base of the Danian Stage (Paleocene, Paleogene, "Tertiary", Cenozoic) at El Kef, Tunisia - Original definition and revision". Episodes. 29 (4): 263–278. doi:10.18814/epiiugs/2006/v29i4/004.
  4. ^ Steininger, Fritz F.; M. P. Aubry; W. A. Berggren; M. Biolzi; A. M. Borsetti; Julie E. Cartlidge; F. Cati; R. Corfield; R. Gelati; S. Iaccarino; C. Napoleone; F. Ottner; F. Rögl; R. Roetzel; S. Spezzaferri; F. Tateo; G. Villa; D. Zevenboom (1997). "The Global Stratotype Section and Point (GSSP) for the base of the Neogene" (PDF). Episodes. 20 (1): 23–28. doi:10.18814/epiiugs/1997/v20i1/005.
  5. ^ "GeoWhen Database – What Happened to the Tertiary?". www.stratigraphy.org.
  6. ^ Federal Geographic Data Committee. "FGDC Digital Cartographic Standard for Geologic Map Symbolization" (PDF). The National Geologic Map Database. United States Geological Survey. Retrieved 29 January 2022.
  7. ^ Orndorff, R.C. (20 July 2010). "Divisions of Geologic Time—Major Chronostratigraphic and Geochronologic Units" (PDF). United States Geological Survey. Retrieved 29 January 2022.
  8. ^ Meredith, R. W.; Janecka, J. E.; Gatesy, J.; Ryder, O. A.; Fisher, C. A.; Teeling, E. C.; Goodbla, A.; Eizirik, E.; Simao, T. L. L.; Stadler, T.; Rabosky, D. L.; Honeycutt, R. L.; Flynn, J. J.; Ingram, C. M.; Steiner, C.; Williams, T. L.; Robinson, T. J.; Burk-Herrick, A.; Westerman, M.; Ayoub, N. A.; Springer, M. S.; Murphy, W. J. (28 October 2011). "Impacts of the Cretaceous Terrestrial Revolution and KPg Extinction on Mammal Diversification". Science. 334 (6055): 521–524. Bibcode:2011Sci...334..521M. doi:10.1126/science.1211028. PMID 21940861. S2CID 38120449.
  9. ^ a b c d e f g Scotese, Christopher Robert; Song, Haijun; Mills, Benjamin J.W.; van der Meer, Douwe G. (April 2021). "Phanerozoic paleotemperatures: The earth's changing climate during the last 540 million years". Earth-Science Reviews. 215: 103503. Bibcode:2021ESRv..21503503S. doi:10.1016/j.earscirev.2021.103503. S2CID 233579194. Retrieved 23 September 2023.
  10. ^ "International Commission on Stratigraphy". stratigraphy.org. Retrieved 2024-07-15.
  11. ^ a b c d Vandenberghe, N.; Hilgen, F. J.; Speijer, R. P.; Ogg, J. G.; Gradstein, F. M.; Hammer, O.; Hollis, C. J.; Hooker, J. J. (2012-01-01), Gradstein, Felix M.; Ogg, James G.; Schmitz, Mark D.; Ogg, Gabi M. (eds.), "Chapter 28 - The Paleogene Period", The Geologic Time Scale, Boston: Elsevier, pp. 855–921, ISBN 978-0-444-59425-9, retrieved 2024-07-15
  12. ^ a b c d e f g h i j k l m n o p q Torsvik, Trond H.; Cocks, Leonard Robert Morrison (2017). Earth history and palaeogeography. Cambridge: Cambridge university press. ISBN 978-1-107-10532-4.
  13. ^ a b c d e Royden, Leigh; Faccenna, Claudio (2018-05-30). "Subduction Orogeny and the Late Cenozoic Evolution of the Mediterranean Arcs". Annual Review of Earth and Planetary Sciences. 46 (1): 261–289. Bibcode:2018AREPS..46..261R. doi:10.1146/annurev-earth-060115-012419. ISSN 0084-6597.
  14. ^ a b Martín-Martín, Manuel; Perri, Francesco; Critelli, Salvatore (2023-08-01). "Cenozoic detrital suites from the Internal Betic-Rif Cordilleras (S Spain and N Morocco): implications for paleogeography and paleotectonics". Earth-Science Reviews. 243: 104498. Bibcode:2023ESRv..24304498M. doi:10.1016/j.earscirev.2023.104498. ISSN 0012-8252.
  15. ^ a b Brunsmann, Quentin; Rosenberg, Claudio Luca; Bellahsen, Nicolas (2024). "The Western Alpine arc: a review and new kinematic model". Comptes Rendus. Géoscience (in French). 356 (S2): 231–263. doi:10.5802/crgeos.253. ISSN 1778-7025.
  16. ^ a b Stephenson, Randell; Schiffer, Christian; Peace, Alexander; Nielsen, Søren Bom; Jess, Scott (2020-11-01). "Late Cretaceous-Cenozoic basin inversion and palaeostress fields in the North Atlantic-western Alpine-Tethys realm: Implications for intraplate tectonics". Earth-Science Reviews. 210: 103252. Bibcode:2020ESRv..21003252S. doi:10.1016/j.earscirev.2020.103252. hdl:2164/16706. ISSN 0012-8252.
  17. ^ Brombin, Valentina; Bonadiman, Costanza; Jourdan, Fred; Roghi, Guido; Coltorti, Massimo; Webb, Laura E.; Callegaro, Sara; Bellieni, Giuliano; De Vecchi, Giampaolo; Sedea, Roberto; Marzoli, Andrea (2019-05-01). "Intraplate magmatism at a convergent plate boundary: The case of the Cenozoic northern Adria magmatism". Earth-Science Reviews. 192: 355–378. Bibcode:2019ESRv..192..355B. doi:10.1016/j.earscirev.2019.03.016. hdl:11392/2403525. ISSN 0012-8252.
  18. ^ Schmid, Stefan M.; Fügenschuh, Bernhard; Kounov, Alexandre; Maţenco, Liviu; Nievergelt, Peter; Oberhänsli, Roland; Pleuger, Jan; Schefer, Senecio; Schuster, Ralf; Tomljenović, Bruno; Ustaszewski, Kamil; van Hinsbergen, Douwe J. J. (2020-02-01). "Tectonic units of the Alpine collision zone between Eastern Alps and western Turkey". Gondwana Research. 78: 308–374. Bibcode:2020GondR..78..308S. doi:10.1016/j.gr.2019.07.005. hdl:1874/394073. ISSN 1342-937X.
  19. ^ a b c Koshnaw, Renas I.; Schlunegger, Fritz; Stockli, Daniel F. (2021-11-03). "Detrital zircon provenance record of the Zagros mountain building from the Neotethys obduction to the Arabia–Eurasia collision, NW Zagros fold–thrust belt, Kurdistan region of Iraq". Solid Earth. 12 (11): 2479–2501. Bibcode:2021SolE...12.2479K. doi:10.5194/se-12-2479-2021. ISSN 1869-9510.
  20. ^ a b c Fu, Xiaofei; Feng, Zhiqiang; Zhang, Faqiang; Zhang, Zhongmin; Guo, Jinrui; Cao, Zhe; Kor, Ting; Cheng, Ming; Yan, Jianzhao; Zhou, Yu (2024-03-01). "Wilson cycles of the Zagros fold and thrust belt: A comprehensive review". Journal of Asian Earth Sciences. 262: 105993. Bibcode:2024JAESc.26205993F. doi:10.1016/j.jseaes.2023.105993. ISSN 1367-9120.
  21. ^ Metcalfe, Ian (2021-12-01). "Multiple Tethyan ocean basins and orogenic belts in Asia". Gondwana Research. SPECIAL ISSUE: GR-100. 100: 87–130. Bibcode:2021GondR.100...87M. doi:10.1016/j.gr.2021.01.012. ISSN 1342-937X.
  22. ^ a b Martin, Craig R.; Jagoutz, Oliver; Upadhyay, Rajeev; Royden, Leigh H.; Eddy, Michael P.; Bailey, Elizabeth; Nichols, Claire I. O.; Weiss, Benjamin P. (2020-11-24). "Paleocene latitude of the Kohistan–Ladakh arc indicates multistage India–Eurasia collision". Proceedings of the National Academy of Sciences. 117 (47): 29487–29494. Bibcode:2020PNAS..11729487M. doi:10.1073/pnas.2009039117. ISSN 0027-8424. PMC 7703637. PMID 33148806.
  23. ^ van Hinsbergen, Douwe J. J.; Lippert, Peter C.; Li, Shihu; Huang, Wentao; Advokaat, Eldert L.; Spakman, Wim (2019-06-05). "Reconstructing Greater India: Paleogeographic, kinematic, and geodynamic perspectives". Tectonophysics. Linking Plate Tectonics and Volcanism to Deep Earth Dynamics – a tribute to Trond H. Torsvik. 760: 69–94. Bibcode:2019Tectp.760...69V. doi:10.1016/j.tecto.2018.04.006. hdl:1874/380963. ISSN 0040-1951.
  24. ^ Meng, Jun; Gilder, Stuart A.; Tan, Xiaodong; Li, Xin; Li, Yalin; Luo, Hui; Suzuki, Noritoshi; Wang, Zihao; Chi, Yuchen; Zhang, Chunyang; Wang, Chengshan (2023-08-15). "Strengthening the argument for a large Greater India". Proceedings of the National Academy of Sciences. 120 (33): e2305928120. Bibcode:2023PNAS..12005928M. doi:10.1073/pnas.2305928120. ISSN 0027-8424. PMC 10433724. PMID 37552758.
  25. ^ a b c Metcalfe, Ian (2021-12-01). "Multiple Tethyan ocean basins and orogenic belts in Asia". Gondwana Research. SPECIAL ISSUE: GR-100. 100: 87–130. Bibcode:2021GondR.100...87M. doi:10.1016/j.gr.2021.01.012. ISSN 1342-937X.
  26. ^ a b c Morley, C. K. (2012-10-01). "Late Cretaceous–Early Palaeogene tectonic development of SE Asia". Earth-Science Reviews. 115 (1): 37–75. Bibcode:2012ESRv..115...37M. doi:10.1016/j.earscirev.2012.08.002. ISSN 0012-8252.
  27. ^ Gaina, Carmen; Jakob, Johannes (2019-06-05). "Global Eocene tectonic unrest: Possible causes and effects around the North American plate". Tectonophysics. Linking Plate Tectonics and Volcanism to Deep Earth Dynamics – a tribute to Trond H. Torsvik. 760: 136–151. Bibcode:2019Tectp.760..136G. doi:10.1016/j.tecto.2018.08.010. ISSN 0040-1951.
  28. ^ a b c d e f Abdelmalak, M. M.; Planke, S.; Polteau, S.; Hartz, E. H.; Faleide, J. I.; Tegner, C.; Jerram, D. A.; Millett, J. M.; Myklebust, R. (2019-06-05). "Breakup volcanism and plate tectonics in the NW Atlantic". Tectonophysics. Linking Plate Tectonics and Volcanism to Deep Earth Dynamics – a tribute to Trond H. Torsvik. 760: 267–296. Bibcode:2019Tectp.760..267A. doi:10.1016/j.tecto.2018.08.002. hdl:2164/12816. ISSN 0040-1951.
  29. ^ a b Peace, Alexander L.; Phethean, J. J. J.; Franke, D.; Foulger, G. R.; Schiffer, C.; Welford, J. K.; McHone, G.; Rocchi, S.; Schnabel, M.; Doré, A. G. (2020-07-01). "A review of Pangaea dispersal and Large Igneous Provinces – In search of a causative mechanism". Earth-Science Reviews. A new paradigm for the North Atlantic Realm. 206: 102902. Bibcode:2020ESRv..20602902P. doi:10.1016/j.earscirev.2019.102902. hdl:11568/992336. ISSN 0012-8252.
  30. ^ a b c d Stanley, Steven; Luczaj, John (2015). Earth System Science (4th ed.). New York: W.H.Freeman and Company. ISBN 978-1-319-15402-8.
  31. ^ a b Yonkee, W. A.; Weil, A. B.; Wells, M. L. (2024-07-01). "Integrating structural, paleomagnetic, and thermo/geochronologic studies to understand evolution of the Sevier and Laramide belts, northern Utah to Wyoming". Journal of Structural Geology. 184: 105104. Bibcode:2024JSG...18405104Y. doi:10.1016/j.jsg.2024.105104. ISSN 0191-8141.
  32. ^ a b Horton, Brian K. (February 2018). "Tectonic Regimes of the Central and Southern Andes: Responses to Variations in Plate Coupling During Subduction". Tectonics. 37 (2): 402–429. Bibcode:2018Tecto..37..402H. doi:10.1002/2017TC004624. ISSN 0278-7407.
  33. ^ Pfiffner, O. Adrian; Gonzalez, Laura (June 2013). "Mesozoic–Cenozoic Evolution of the Western Margin of South America: Case Study of the Peruvian Andes". Geosciences. 3 (2): 262–310. Bibcode:2013Geosc...3..262P. doi:10.3390/geosciences3020262. ISSN 2076-3263.
  34. ^ a b c d e f g h i Seton, M.; Müller, R. D.; Zahirovic, S.; Gaina, C.; Torsvik, T.; Shephard, G.; Talsma, A.; Gurnis, M.; Turner, M.; Maus, S.; Chandler, M. (2012-07-01). "Global continental and ocean basin reconstructions since 200 Ma". Earth-Science Reviews. 113 (3): 212–270. Bibcode:2012ESRv..113..212S. doi:10.1016/j.earscirev.2012.03.002. ISSN 0012-8252.
  35. ^ a b Montes, Camilo; Rodriguez-Corcho, Andres Felipe; Bayona, German; Hoyos, Natalia; Zapata, Sebastian; Cardona, Agustin (2019-11-01). "Continental margin response to multiple arc-continent collisions: The northern Andes-Caribbean margin". Earth-Science Reviews. 198: 102903. Bibcode:2019ESRv..19802903M. doi:10.1016/j.earscirev.2019.102903. ISSN 0012-8252.
  36. ^ van Benthem, Steven; Govers, Rob; Spakman, Wim; Wortel, Rinus (2013). "Tectonic evolution and mantle structure of the Caribbean". Journal of Geophysical Research Solid Earth. 118 (6): 3019–3036. Bibcode:2013JGRB..118.3019V. doi:10.1002/jgrb.50235. ISSN 2169-9313.
  37. ^ a b Seton, Maria; Flament, Nicolas; Whittaker, Joanne; Müller, R. Dietmar; Gurnis, Michael; Bower, Dan J. (2015-03-28). "Ridge subduction sparked reorganization of the Pacific plate-mantle system 60–50 million years ago". Geophysical Research Letters. 42 (6): 1732–1740. Bibcode:2015GeoRL..42.1732S. doi:10.1002/2015GL063057. ISSN 0094-8276.
  38. ^ a b Wright, Nicky M.; Seton, Maria; Williams, Simon E.; Müller, R. Dietmar (2016-03-01). "The Late Cretaceous to recent tectonic history of the Pacific Ocean basin". Earth-Science Reviews. 154: 138–173. Bibcode:2016ESRv..154..138W. doi:10.1016/j.earscirev.2015.11.015. ISSN 0012-8252.
  39. ^ Jiang, Zhaoxia; Li, Sanzhong; Liu, Qingsong; Zhang, Jianli; Zhou, Zaizheng; Zhang, Yuzhen (2021-04-01). "The trials and tribulations of the Hawaii hot spot model". Earth-Science Reviews. 215: 103544. Bibcode:2021ESRv..21503544J. doi:10.1016/j.earscirev.2021.103544. ISSN 0012-8252.
  40. ^ a b c Boone, Samuel C.; Balestrieri, Maria-Laura; Kohn, Barry (2021). "Tectono-Thermal Evolution of the Red Sea Rift". Frontiers in Earth Science. 9: 588. Bibcode:2021FrEaS...9..588B. doi:10.3389/feart.2021.713448. ISSN 2296-6463.
  41. ^ Jehle, Sofie; Bornemann, André; Lägel, Anna Friederike; Deprez, Arne; Speijer, Robert P. (1 July 2019). "Paleoceanographic changes across the Latest Danian Event in the South Atlantic Ocean and planktic foraminiferal response". Palaeogeography, Palaeoclimatology, Palaeoecology. 525: 1–13. Bibcode:2019PPP...525....1J. doi:10.1016/j.palaeo.2019.03.024. S2CID 134929774. Retrieved 30 December 2022.
  42. ^ Jehle, Sofie; Bornemann, André; Deprez, Arne; Speijer, Robert P. (25 November 2015). "The Impact of the Latest Danian Event on Planktic Foraminiferal Faunas at ODP Site 1210 (Shatsky Rise, Pacific Ocean)". PLOS ONE. 10 (11): e0141644. Bibcode:2015PLoSO..1041644J. doi:10.1371/journal.pone.0141644. PMC 4659543. PMID 26606656.
  43. ^ Sprong, M.; Youssef, J. A.; Bornemann, André; Schulte, P.; Steurbaut, E.; Stassen, P.; Kouwenhoven, T. J.; Speijer, Robert P. (September 2011). "A multi-proxy record of the Latest Danian Event at Gebel Qreiya, Eastern Desert, Egypt" (PDF). Journal of Micropalaeontology. 30 (2): 167–182. Bibcode:2011JMicP..30..167S. doi:10.1144/0262-821X10-023. S2CID 55038043. Archived from the original (PDF) on 28 June 2023. Retrieved 30 December 2022.
  44. ^ Naafs, B. D. A.; Rohrssen, M.; Inglis, G. N.; Lähteenoja, O.; Feakins, S. J.; Collinson, M. E.; Kennedy, E. M.; Singh, P. K.; Singh, M. P.; Lunt, D. J.; Pancost, R. D. (2018). "High temperatures in the terrestrial mid-latitudes during the early Palaeogene" (PDF). Nature Geoscience. 11 (10): 766–771. Bibcode:2018NatGe..11..766N. doi:10.1038/s41561-018-0199-0. hdl:1983/82e93473-2a5d-4a6d-9ca1-da5ebf433d8b. S2CID 135045515.
  45. ^ University of Bristol (30 July 2018). "Ever-increasing CO2 levels could take us back to the tropical climate of Paleogene period". ScienceDaily.
  46. ^ a b "Ever-increasing CO2 levels could take us back to the tropical climate of Paleogene period". University of Bristol. 2018.
  47. ^ Wing, S. L. (11 November 2005). "Transient Floral Change and Rapid Global Warming at the Paleocene-Eocene Boundary". Science. 310 (5750): 993–996. Bibcode:2005Sci...310..993W. doi:10.1126/science.1116913. ISSN 0036-8075. PMID 16284173. S2CID 7069772. Retrieved 23 September 2023.
  48. ^ Inglis, Gordon N.; Bragg, Fran; Burls, Natalie J.; Cramwinckel, Margot J.; Evans, David; Foster, Gavin L.; Huber, Matthew; Lunt, Daniel J.; Siler, Nicholas; Steinig, Sebastian; Tierney, Jessica E.; Wilkinson, Richard; Anagnostou, Eleni; de Boer, Agatha M.; Dunkley Jones, Tom (26 October 2020). "Global mean surface temperature and climate sensitivity of the early Eocene Climatic Optimum (EECO), Paleocene–Eocene Thermal Maximum (PETM), and latest Paleocene". Climate of the Past. 16 (5): 1953–1968. Bibcode:2020CliPa..16.1953I. doi:10.5194/cp-16-1953-2020. hdl:1983/24a30f12-51a6-4544-9db8-b2009e33c02a. ISSN 1814-9332. Retrieved 23 September 2023.
  49. ^ Stap, L.; Lourens, L.J.; Thomas, E.; Sluijs, A.; Bohaty, S.; Zachos, J.C. (1 July 2010). "High-resolution deep-sea carbon and oxygen isotope records of Eocene Thermal Maximum 2 and H2". Geology. 38 (7): 607–610. Bibcode:2010Geo....38..607S. doi:10.1130/G30777.1. hdl:1874/385773. S2CID 41123449.
  50. ^ Bohaty, Steven M.; Zachos, James C. (1 November 2003). "Significant Southern Ocean warming event in the late middle Eocene". Geology. 31 (11): 1017. Bibcode:2003Geo....31.1017B. doi:10.1130/G19800.1. ISSN 0091-7613. Retrieved 23 September 2023.
  51. ^ Pearson, Paul N.; Foster, Gavin L.; Wade, Bridget S. (13 September 2009). "Atmospheric carbon dioxide through the Eocene–Oligocene climate transition". Nature. 461 (7267): 1110–1113. Bibcode:2009Natur.461.1110P. doi:10.1038/nature08447. ISSN 0028-0836. PMID 19749741. S2CID 205218274. Retrieved 23 September 2023.
  52. ^ Sauermilch, Isabel; Whittaker, Joanne M.; Klocker, Andreas; Munday, David R.; Hochmuth, Katharina; Bijl, Peter K.; LaCasce, Joseph H. (9 November 2021). "Gateway-driven weakening of ocean gyres leads to Southern Ocean cooling". Nature Communications. 12 (1): 6465. Bibcode:2021NatCo..12.6465S. doi:10.1038/s41467-021-26658-1. ISSN 2041-1723. PMC 8578591. PMID 34753912.
  53. ^ Barker, P.F.; Thomas, E. (June 2004). "Origin, signature and palaeoclimatic influence of the Antarctic Circumpolar Current". Earth-Science Reviews. 66 (1–2): 143–162. Bibcode:2004ESRv...66..143B. doi:10.1016/j.earscirev.2003.10.003. Retrieved 23 September 2023.
  54. ^ Zachos, James C.; Lohmann, Kyger C.; Walker, James C. G.; Wise, Sherwood W. (March 1993). "Abrupt Climate Change and Transient Climates during the Paleogene: A Marine Perspective". The Journal of Geology. 101 (2): 191–213. Bibcode:1993JG....101..191Z. doi:10.1086/648216. ISSN 0022-1376. PMID 11537739. S2CID 29784731. Retrieved 23 September 2023.
  55. ^ Hochmuth, Katharina; Whittaker, Joanne M.; Sauermilch, Isabel; Klocker, Andreas; Gohl, Karsten; LaCasce, Joseph H. (9 November 2022). "Southern Ocean biogenic blooms freezing-in Oligocene colder climates". Nature Communications. 13 (1): 6785. Bibcode:2022NatCo..13.6785H. doi:10.1038/s41467-022-34623-9. ISSN 2041-1723. PMC 9646741. PMID 36351905.
  56. ^ Crame, J. Alistair (March 2020). "Early Cenozoic evolution of the latitudinal diversity gradient". Earth-Science Reviews. 202: 103090. Bibcode:2020ESRv..20203090C. doi:10.1016/j.earscirev.2020.103090. S2CID 214219923.
  57. ^ Traverse, Alfred (1988). Paleopalynology. Unwin Hyman. ISBN 978-0-04-561001-3. OCLC 17674795.
  58. ^ Muller, Jan (January 1981). "Fossil pollen records of extant angiosperms". The Botanical Review. 47 (1): 1–142. Bibcode:1981BotRv..47....1M. doi:10.1007/bf02860537. ISSN 0006-8101. S2CID 10574478.
edit