The Holocene (/ˈhɒl.əsn, --, ˈh.lə-, -l-/)[2][3] is the current geological epoch, beginning approximately 11,700 years ago.[4] It follows the Last Glacial Period, which concluded with the Holocene glacial retreat.[4] The Holocene and the preceding Pleistocene[5] together form the Quaternary period. The Holocene is an interglacial period within the ongoing glacial cycles of the Quaternary, and is equivalent to Marine Isotope Stage 1.

Holocene
0.0117 – 0 Ma
Chronology
Etymology
Name formalityFormal
Usage information
Celestial bodyEarth
Regional usageGlobal (ICS)
Time scale(s) usedICS Time Scale
Definition
Chronological unitEpoch
Stratigraphic unitSeries
Time span formalityFormal
Lower boundary definitionEnd of the Younger Dryas stadial.
Lower boundary GSSPNGRIP2 ice core, Greenland
75°06′00″N 42°19′12″W / 75.1000°N 42.3200°W / 75.1000; -42.3200
Lower GSSP ratified2008[1]
Upper boundary definitionPresent day
Upper boundary GSSPN/A
N/A
Upper GSSP ratifiedN/A

The Holocene corresponds with the rapid proliferation, growth, and impacts of the human species worldwide, including all of its written history, technological revolutions, development of major civilizations, and overall significant transition towards urban living in the present. The human impact on modern-era Earth and its ecosystems may be considered of global significance for the future evolution of living species, including approximately synchronous lithospheric evidence, or more recently hydrospheric and atmospheric evidence of the human impact. In July 2018, the International Union of Geological Sciences split the Holocene Epoch into three distinct ages based on the climate, Greenlandian (11,700 years ago to 8,200 years ago), Northgrippian (8,200 years ago to 4,200 years ago) and Meghalayan (4,200 years ago to the present), as proposed by International Commission on Stratigraphy.[6] The oldest age, the Greenlandian was characterized by a warming following the preceding ice age. The Northgrippian Age is known for vast cooling due to a disruption in ocean circulations that was caused by the melting of glaciers. The most recent age of the Holocene is the present Meghalayan, which began with extreme drought that lasted around 200 years.[6]

Etymology edit

The word Holocene was formed from two Ancient Greek words. Hólos (ὅλος) is the Greek word for "whole". "Cene" comes from the Greek word kainós (καινός), meaning "new". The concept is that this epoch is "entirely new".[7][8][9] The suffix '-cene' is used for all the seven epochs of the Cenozoic Era.

Overview edit

The International Commission on Stratigraphy has defined the Holocene as starting approximately 11,700 years before 2000 CE (11,650 cal years BP, or 9,700 BCE).[4] The Subcommission on Quaternary Stratigraphy (SQS) regards the term 'recent' as an incorrect way of referring to the Holocene, preferring the term 'modern' instead to describe current processes. It also observes that the term 'Flandrian' may be used as a synonym for Holocene, although it is becoming outdated.[10] The International Commission on Stratigraphy, however, considers the Holocene to be an epoch following the Pleistocene and specifically following the last glacial period. Local names for the last glacial period include the Wisconsinan in North America,[11] the Weichselian in Europe,[12] the Devensian in Britain,[13] the Llanquihue in Chile[14] and the Otiran in New Zealand.[15]

The Holocene can be subdivided into five time intervals, or chronozones, based on climatic fluctuations:[16][needs update?]

Note: "ka BP" means "kilo-annum Before Present", i.e. 1,000 years before 1950 (non-calibrated C14 dates)

Geologists working in different regions are studying sea levels, peat bogs, and ice-core samples, using a variety of methods, with a view toward further verifying and refining the Blytt–Sernander sequence. This is a classification of climatic periods initially defined by plant remains in peat mosses.[17] Though the method was once thought to be of little interest, based on 14C dating of peats that was inconsistent with the claimed chronozones,[18] investigators have found a general correspondence across Eurasia and North America. The scheme was defined for Northern Europe, but the climate changes were claimed to occur more widely. The periods of the scheme include a few of the final pre-Holocene oscillations of the last glacial period and then classify climates of more recent prehistory.[19]

Paleontologists have not defined any faunal stages for the Holocene. If subdivision is necessary, periods of human technological development, such as the Mesolithic, Neolithic, and Bronze Age, are usually used. However, the time periods referenced by these terms vary with the emergence of those technologies in different parts of the world.[20]

According to some scholars, a third epoch of the Quaternary, the Anthropocene, has now begun.[21] This term is used to denote the present time-interval in which many geologically significant conditions and processes have been profoundly altered by human activities. The 'Anthropocene' (a term coined by Paul J. Crutzen and Eugene Stoermer in 2000) is not a formally defined geological unit. The Subcommission on Quaternary Stratigraphy of the International Commission on Stratigraphy has a working group to determine whether it should be. In May 2019, members of the working group voted in favour of recognizing the Anthropocene as formal chrono-stratigraphic unit, with stratigraphic signals around the mid-twentieth century CE as its base. The exact criteria have still to be determined, after which the recommendation also has to be approved by the working group's parent bodies (ultimately the International Union of Geological Sciences).[22]

Geology edit

The Holocene is a geologic epoch that follows directly after the Pleistocene. Continental motions due to plate tectonics are less than a kilometre over a span of only 10,000 years. However, ice melt caused world sea levels to rise about 35 m (115 ft) in the early part of the Holocene and another 30 m in the later part of the Holocene. In addition, many areas above about 40 degrees north latitude had been depressed by the weight of the Pleistocene glaciers and rose as much as 180 m (590 ft) due to post-glacial rebound over the late Pleistocene and Holocene, and are still rising today.[23]

The sea-level rise and temporary land depression allowed temporary marine incursions into areas that are now far from the sea. For example, marine fossils from the Holocene epoch have been found in locations such as Vermont and Michigan. Other than higher-latitude temporary marine incursions associated with glacial depression, Holocene fossils are found primarily in lakebed, floodplain, and cave deposits. Holocene marine deposits along low-latitude coastlines are rare because the rise in sea levels during the period exceeds any likely tectonic uplift of non-glacial origin.[citation needed]

Post-glacial rebound in the Scandinavia region resulted in a shrinking Baltic Sea. The region continues to rise, still causing weak earthquakes across Northern Europe. An equivalent event in North America was the rebound of Hudson Bay, as it shrank from its larger, immediate post-glacial Tyrrell Sea phase, to its present boundaries.[24]

Climate edit

 
Vegetation and water bodies in northern and central Africa in the Eemian (bottom) and Holocene (top)

The climate throughout the Holocene has shown significant variability despite ice core records from Greenland suggesting a more stable climate following the preceding ice age. Marine chemical fluxes during the Holocene were lower than during the Younger Dryas, but were still considerable enough to imply notable changes in the climate.

The temporal and spatial extent of climate change during the Holocene is an area of considerable uncertainty, with radiative forcing recently proposed to be the origin of cycles identified in the North Atlantic region. Climate cyclicity through the Holocene (Bond events) has been observed in or near marine settings and is strongly controlled by glacial input to the North Atlantic.[25][26] Periodicities of ≈2500, ≈1500, and ≈1000 years are generally observed in the North Atlantic.[27][28][29] At the same time spectral analyses of the continental record, which is remote from oceanic influence, reveal persistent periodicities of 1,000 and 500 years that may correspond to solar activity variations during the Holocene Epoch.[30] A 1,500-year cycle corresponding to the North Atlantic oceanic circulation may have had widespread global distribution in the Late Holocene.[30] From 8,500 BP to 6,700 BP, North Atlantic climate oscillations were highly irregular and erratic because of perturbations from substantial ice discharge into the ocean from the collapsing Laurentide Ice Sheet.[31] The Greenland ice core records indicate that climate changes became more regional and had a larger effect on the mid-to-low latitudes and mid-to-high latitudes after ~5600 B.P.[32]

Human activity through land use changes was an important influence on Holocene climatic changes, and is believed to be why the Holocene is an atypical interglacial that has not experienced significant cooling over its course.[33] From the start of the Industrial Revolution onwards, large-scale anthropogenic greenhouse gas emissions caused the Earth to warm.[34] Likewise, climatic changes have induced substantial changes in human civilisation over the course of the Holocene.[35][36]

During the transition from the last glacial to the Holocene, the Huelmo–Mascardi Cold Reversal in the Southern Hemisphere began before the Younger Dryas, and the maximum warmth flowed south to north from 11,000 to 7,000 years ago. It appears that this was influenced by the residual glacial ice remaining in the Northern Hemisphere until the later date.[citation needed] The first major phase of Holocene climate was the Preboreal.[37] At the start of the Preboreal occurred the Preboreal Oscillation (PBO).[38] The Holocene Climatic Optimum (HCO) was a period of warming throughout the globe but was not globally synchronous and uniform.[39] Following the HCO, the global climate entered a broad trend of very gradual cooling known as Neoglaciation, which lasted from the end of the HCO to before the Industrial Revolution.[37] From the 10th-14th century, the climate was similar to that of modern times during a period known as the Mediaeval Warm Period (MWP), also known as the Mediaeval Climatic Optimum (MCO). It was found that the warming that is taking place in current years is both more frequent and more spatially homogeneous than what was experienced during the MWP. A warming of +1 degree Celsius occurs 5–40 times more frequently in modern years than during the MWP. The major forcing during the MWP was due to greater solar activity, which led to heterogeneity compared to the greenhouse gas forcing of modern years that leads to more homogeneous warming. This was followed by the Little Ice Age (LIA) from the 13th or 14th century to the mid-19th century.[40] The LIA was the coldest interval of time of the past two millennia.[41] Following the Industrial Revolution, warm decadal intervals became more common relative to before as a consequence of anthropogenic greenhouse gases, resulting in progressive global warming.[34] In the late 20th century, anthropogenic forcing superseded solar activity as the dominant driver of climate change,[42] though solar activity has continued to play a role.[43][44]

Europe edit

In Northern Germany, the Middle Holocene saw a drastic increase in the amount of raised bogs, most likely related to sea level rise. Although human activity affected geomorphology and landscape evolution in Northern Germany throughout the Holocene, it only became a dominant influence in the last four centuries.[45] In the French Alps, geochemistry and lithium isotope signatures in lake sediments have suggested gradual soil formation from the Last Glacial Period to the Holocene climatic optimum, and this soil development was altered by the settlement of human societies. Early anthropogenic activities such as deforestation and agriculture reinforced soil erosion, which peaked in the Middle Ages at an unprecedented level, marking human forcing as the most powerful factor affecting surface processes.[46]

Africa edit

North Africa, dominated by the Sahara Desert in the present, was instead a savanna dotted with large lakes during the Early and Middle Holocene,[47] regionally known as the African Humid Period (AHP).[48] The northward migration of the Intertropical Convergence Zone (ITCZ) produced increased monsoon rainfall over North Africa.[49] The lush vegetation of the Sahara brought an increase in pastoralism.[50] The AHP ended around 5,500 BP, after which the Sahara began to dry and become the desert it is today.[51]

A stronger East African Monsoon during the Middle Holocene increased precipitation in East Africa and raised lake levels.[52] Around 800 AD, or 1,150 BP, a marine transgression occurred in southeastern Africa; in the Lake Lungué basin, this sea level highstand occurred from 740 to 910 AD, or from 1,210 to 1,040 BP, as evidenced by the lake's connection to the Indian Ocean at this time. This transgression was followed by a period of transition that lasted until 590 BP, when the region experienced significant aridification and began to be extensively used by humans for livestock herding.[53]

In the Kalahari Desert, Holocene climate was overall very stable and environmental change was of low amplitude. Relatively cool conditions have prevailed since 4,000 BP.[54]

Middle East edit

During the Late Holocene, the coastline of the Levant receded westward, prompting a shift in human settlement patterns following this marine regression.[55]

Central Asia edit

In Xinjiang, long-term Holocene warming increased meltwater supply during summers, creating large lakes and oases at low altitudes and inducing enhanced moisture recycling.[56] In the Tien Shan, sedimentological evidence from Swan Lake suggests the period between 8,500 and 6,900 BP was relatively warm, with steppe meadow vegetation being predominant. An increase in Cyperaceae from 6,900 to 2,600 BP indicates cooling and humidification of the Tian Shan climate that was interrupted by a warm period between 5,500 and 4,500 BP. After 2,600 BP, an alpine steppe climate prevailed across the region.[57] Sand dune evolution in the Bayanbulak Basin shows that the region was very dry from the Holocene's beginning until around 6,500 BP, when a wet interval began.[58] In the Tibetan Plateau, the moisture optimum spanned from around 7,500 to 5,500 BP.[59]

South Asia edit

After 11,800 BP, and especially between 10,800 and 9,200 BP, Ladakh experienced tremendous moisture increase most likely related to the strengthening of the Indian Summer Monsoon (ISM). From 9,200 to 6,900 BP, relative aridity persisted in Ladakh. A second major humid phase occurred in Ladakh from 6,900 to 4,800 BP, after which the region was again arid.[60]

From 900 to 1,200 AD, during the MWP, the ISM was again strong as evidenced by low δ18O values from the Ganga Plain.[61]

The sediments of Lonar Lake in Maharashtra record dry conditions around 11,400 BP that transitioned into a much wetter climate from 11,400 to 11,100 BP due to intensification of the ISM. Over the Early Holocene, the region was very wet, but during the Middle Holocene from 6,200 to 3,900 BP, aridification occurred, with the subsequent Late Holocene being relatively arid as a whole.[62]

Coastal southwestern India experienced a stronger ISM from 9,690 to 7,560 BP, during the HCO. From 3,510 to 2,550 BP, during the Late Holocene, the ISM became weaker, although this weakening was interrupted by an interval of unusually high ISM strength from 3,400 to 3,200 BP.[63]

East Asia edit

Northern China experienced an abrupt aridification event approximately 4,000 BP.[64] From around 3,500 to 3,000 BP, northeastern China underwent a prolonged cooling, manifesting itself with the disruption of Bronze Age civilisations in the region.[65] Eastern and southern China, the monsoonal regions of China, were wetter than present in the Early and Middle Holocene.[66] Lake Huguangyan's TOC, δ13Cwax, δ13Corg, δ15N values suggest the period of peak moisture lasted from 9,200 to 1,800 BP and was attributable to a strong East Asian Summer Monsoon (EASM).[67] Late Holocene cooling events in the region were dominantly influenced by solar forcing, with many individual cold snaps linked to solar minima such as the Oort, Wolf, Spörer, and Maunder Minima.[68] Monsoonal regions of China became more arid in the Late Holocene.[66]

Southeast Asia edit

Before 7,500 BP, the Gulf of Thailand was exposed above sea level and was very arid. A marine transgression occurred from 7,500 to 6,200 BP amidst global warming.[69]

North America edit

During the Middle Holocene, western North America was drier than present, with wetter winters and drier summers.[70] After the end of the thermal maximum of the HCO around 4,500 BP, the East Greenland Current underwent strengthening.[71] A massive megadrought occurred from 2,800 to 1,850 BP in the Great Basin.[72]

Eastern North America underwent abrupt warming and humidification around 10,500 BP and then declined from 9,300 to 9,100 BP. The region has undergone a long term wettening since 5,500 BP occasionally interrupted by intervals of high aridity. A major cool event lasting from 5,500 to 4,700 BP was coeval with a major humidification before being terminated by a major drought and warming at the end of that interval.[73]

South America edit

During the Early Holocene, relative sea level rose in the Bahia region, causing a landward expansion of mangroves. During the Late Holocene, the mangroves declined as sea level dropped and freshwater supply increased.[74] In the Santa Catarina region, the maximum sea level highstand was around 2.1 metres above present and occurred about 5,800 to 5,000 BP.[75] Sea levels at Rocas Atoll were likewise higher than present for much of the Late Holocene.[76]

Australia edit

The Northwest Australian Summer Monsoon was in a strong phase from 8,500 to 6,400 BP, from 5,000 to 4,000 BP (possibly until 3,000 BP), and from 1,300 to 900 BP, with weak phases in between and the current weak phase beginning around 900 BP after the end of the last strong phase.[77]

New Zealand edit

Ice core measurements imply that the sea surface temperature (SST) gradient east of New Zealand, across the subtropical front (STF), was around 2 degrees Celsius during the HCO. This temperature gradient is significantly less than modern times, which is around 6 degrees Celsius. A study utilizing five SST proxies from 37°S to 60°S latitude confirmed that the strong temperature gradient was confined to the area immediately south of the STF, and is correlated with reduced westerly winds near New Zealand.[78] Since 7,100 BP, New Zealand experienced 53 cyclones similar in magnitude to Cyclone Bola.[79]

Pacific edit

Evidence from the Galápagos Islands shows that the El Niño–Southern Oscillation (ENSO) was significantly weaker during the Middle Holocene, but that the strength of ENSO became moderate to high over the Late Holocene.[80]

Ecological developments edit

Animal and plant life have not evolved much during the relatively short Holocene, but there have been major shifts in the richness and abundance of plants and animals. A number of large animals including mammoths and mastodons, saber-toothed cats like Smilodon and Homotherium, and giant sloths went extinct in the late Pleistocene and early Holocene. These extinctions can be mostly attributed to people.[81] In America, it coincided with the arrival of the Clovis people; this culture was known for "Clovis points" which were fashioned on spears for hunting animals. Shrubs, herbs, and mosses had also changed in relative abundance from the Pleistocene to Holocene, identified by permafrost core samples.[82]

Throughout the world, ecosystems in cooler climates that were previously regional have been isolated in higher altitude ecological "islands".[83]

The 8.2-ka event, an abrupt cold spell recorded as a negative excursion in the δ18O record lasting 400 years, is the most prominent climatic event occurring in the Holocene Epoch, and may have marked a resurgence of ice cover. It has been suggested that this event was caused by the final drainage of Lake Agassiz, which had been confined by the glaciers, disrupting the thermohaline circulation of the Atlantic.[84] This disruption was the result of an ice dam over Hudson Bay collapsing sending cold lake Agassiz water into the North Atlantic ocean.[85] Furthermore, studies show that the melting of Lake Agassiz led to sea-level rise which flooded the North American coastal landscape. The basal peat plant was then used to determine the resulting local sea-level rise of 0.20-0.56m in the Mississippi Delta.[85] Subsequent research, however, suggested that the discharge was probably superimposed upon a longer episode of cooler climate lasting up to 600 years and observed that the extent of the area affected was unclear.[86]

Human developments edit

 
Overview map of the world at the end of the 2nd millennium BC, color-coded by cultural stage:
  hunter-gatherers (Palaeolithic or Mesolithic)
  nomadic pastoralists
  simple farming societies
  complex farming societies (Bronze Age Old World, Olmecs, Andes)
  state societies (Fertile Crescent, Egypt, China)

The beginning of the Holocene corresponds with the beginning of the Mesolithic age in most of Europe. In regions such as the Middle East and Anatolia, the term Epipaleolithic is preferred in place of Mesolithic, as they refer to approximately the same time period. Cultures in this period include Hamburgian, Federmesser, and the Natufian culture, during which the oldest inhabited places still existing on Earth were first settled, such as Tell es-Sultan (Jericho) in the Middle East.[87] There is also evolving archeological evidence of proto-religion at locations such as Göbekli Tepe, as long ago as the 9th millennium BC.[88]

The preceding period of the Late Pleistocene had already brought advancements such as the bow and arrow, creating more efficient forms of hunting and replacing spear throwers. In the Holocene, however, the domestication of plants and animals allowed humans to develop villages and towns in centralized locations. Archaeological data shows that between 10,000 and 7,000 BP rapid domestication of plants and animals took place in tropical and subtropical parts of Asia, Africa, and Central America.[89] The development of farming allowed humans to transition away from hunter-gatherer nomadic cultures, which did not establish permanent settlements, to a more sustainable sedentary lifestyle. This form of lifestyle change allowed humans to develop towns and villages in centralized locations, which gave rise to the world known today. It is believed that the domestication of plants and animals began in the early part of the Holocene in the tropical areas of the planet.[89] Because these areas had warm, moist temperatures, the climate was perfect for effective farming. Culture development and human population change, specifically in South America, has also been linked to spikes in hydroclimate resulting in climate variability in the mid-Holocene (8.2 - 4.2 k cal BP).[90] Climate change on seasonality and available moisture also allowed for favorable agricultural conditions which promoted human development for Maya and Tiwanaku regions.[91] In the Korean Peninsula, climatic changes fostered a population boom during the Middle Chulmun period from 5,500 to 5,000 BP, but contributed to a subsequent bust during the Late and Final Chulmun periods, from 5,000 to 4,000 BP and from 4,000 to 3,500 BP respectively.[92]

Extinction event edit

The Holocene extinction, otherwise referred to as the sixth mass extinction or Anthropocene extinction,[93][94] is an ongoing extinction event of species during the present Holocene epoch (with the more recent time sometimes called Anthropocene) as a result of human activity.[95][96][97][98] The included extinctions span numerous families of fungi,[99] plants,[100][101] and animals, including mammals, birds, reptiles, amphibians, fish and invertebrates. With widespread degradation of highly biodiverse habitats such as coral reefs and rainforests, as well as other areas, the vast majority of these extinctions are thought to be undocumented, as the species are undiscovered at the time of their extinction, or no one has yet discovered their extinction. The current rate of extinction of species is estimated at 100 to 1,000 times higher than natural background extinction rates.[96][85][102][103]

Gallery edit

See also edit

Notes edit

References edit

  1. ^ Walker, Mike; Johnse, Sigfus; Rasmussen, Sune; Steffensen, Jørgen-Peder; Popp, Trevor; Gibbard, Phillip; Hoek, Wilm; Lowe, John; Andrews, John; Björck, Svante; Cwynar, Les; Hughen, Konrad; Kershaw, Peter; Kromer, Bernd; Litt, Thomas; Lowe, David; Nakagawa, Takeshi; Newnham, Rewi; Schwande, Jakob (June 2008). "The Global Stratotype Section and Point (GSSP) for the base of the Holocene Series/Epoch (Quaternary System/Period) in the NGRIP ice core". Episodes. 32 (2): 264–267. doi:10.18814/epiiugs/2008/v31i2/016. hdl:10289/920.
  2. ^ "Holocene". Merriam-Webster.com Dictionary. Retrieved 2018-02-11.
  3. ^ "Holocene". Dictionary.com Unabridged (Online). n.d. Retrieved 2018-02-11.
  4. ^ a b c Walker, Mike; Johnsen, Sigfus; Rasmussen, Sune Olander; Popp, Trevor; Steffensen, Jorgen-Peder; Gibrard, Phil; Hoek, Wim; Lowe, John; Andrews, John; Bjo Rck, Svante; Cwynar, Les C.; Hughen, Konrad; Kersahw, Peter; Kromer, Bernd; Litt, Thomas; Lowe, David J.; Nakagawa, Takeshi; Newnham, Rewi; Schwander, Jakob (2009). "Formal definition and dating of the GSSP (Global Stratotype Section and Point) for the base of the Holocene using the Greenland NGRIP ice core, and selected auxiliary records" (PDF). Journal of Quaternary Science. 24 (1): 3–17. Bibcode:2009JQS....24....3W. doi:10.1002/jqs.1227. Archived (PDF) from the original on 2013-11-04. Retrieved 2013-09-03.
  5. ^ Fan, Junxuan; Hou, Xudong. "International Chronostratigraphic Chart". International Commission on Stratigraphy. Archived from the original on January 13, 2017. Retrieved June 18, 2016.
  6. ^ a b Amos, Jonathan (2018-07-18). "Welcome to the Meghalayan Age a new phase in history". BBC News. Archived from the original on 2018-07-18. Retrieved 2018-07-18.
  7. ^ The name "Holocene" was proposed in 1850 by the French palaeontologist and entomologist Paul Gervais (1816–1879): Gervais, Paul (1850). "Sur la répartition des mammifères fossiles entre les différents étages tertiaires qui concourent à former le sol de la France" [On the distribution of mammalian fossils among the different tertiary stages which help to form the ground of France]. Académie des Sciences et Lettres de Montpellier. Section des Sciences (in French). 1: 399–413. Archived from the original on 2020-05-22. Retrieved 2018-07-15. From p. 413: Archived 2020-05-22 at the Wayback Machine "On pourrait aussi appeler Holocènes, ceux de l'époque historique, ou dont le dépôt n'est pas antérieur à la présence de l'homme; … " (One could also call "Holocene" those [deposits] of the historic era, or the deposit of which is not prior to the presence of man; … )
  8. ^ "Origin and meaning of Holocene". Online Etymology Dictionary. Archived from the original on 2019-08-08. Retrieved 2019-08-08.
  9. ^ "Origin and meaning of suffix -cene". Online Etymology Dictionary. Archived from the original on 2019-08-08. Retrieved 2019-08-08.
  10. ^ Gibbard, P. L.; Head, M. J. (2020-01-01), Gradstein, Felix M.; Ogg, James G.; Schmitz, Mark D.; Ogg, Gabi M. (eds.), "Chapter 30 - The Quaternary Period", Geologic Time Scale 2020, Elsevier, pp. 1217–1255, ISBN 978-0-12-824360-2, retrieved 2022-04-21
  11. ^ Clayton, Lee; Moran, Stephen R. (1982). "Chronology of late wisconsinan glaciation in middle North America". Quaternary Science Reviews. 1 (1): 55–82. Bibcode:1982QSRv....1...55C. doi:10.1016/0277-3791(82)90019-1.
  12. ^ Svendsen, John Inge; Astakhov, Valery I.; Bolshiyanov, Dimitri Yu.; Demidov, Igor; Dowdeswell, Julian A.; Gataullin, Valery; Hjort, Christian; Hubberten, Hans W.; Larsen, Eiliv; Mangerud, Jan; Melles, Martin; Moller, Per; Saarnisto, Matti; Siegert, Martin J. (March 1999). "Maximum extent of the Eurasian ice sheets in the Barents and Kara Sea region during the Weichselian" (PDF). Boreas. 28 (1): 234–242. Bibcode:1999Borea..28..234S. doi:10.1111/j.1502-3885.1999.tb00217.x. S2CID 34659675. Archived (PDF) from the original on 2018-02-12. Retrieved 2018-02-11.
  13. ^ Eyles, Nicholas; McCabe, A. Marshall (1989). "The Late Devensian (<22,000 BP) Irish Sea Basin: The sedimentary record of a collapsed ice sheet margin". Quaternary Science Reviews. 8 (4): 307–351. Bibcode:1989QSRv....8..307E. doi:10.1016/0277-3791(89)90034-6.
  14. ^ Denton, G.H.; Lowell, T.V.; Heusser, C.J.; Schluchter, C.; Andersern, B.G.; Heusser, Linda E.; Moreno, P.I.; Marchant, D.R. (1999). "Geomorphology, stratigraphy, and radiocarbon chronology of LlanquihueDrift in the area of the Southern Lake District, Seno Reloncavi, and Isla Grande de Chiloe, Chile" (PDF). Geografiska Annaler: Series A, Physical Geography. 81A (2): 167–229. Bibcode:1999GeAnA..81..167D. doi:10.1111/j.0435-3676.1999.00057.x (inactive 31 January 2024). S2CID 7626031. Archived from the original (PDF) on 2018-02-12.{{cite journal}}: CS1 maint: DOI inactive as of January 2024 (link)
  15. ^ Newnham, R.M.; Vandergoes, M.J.; Hendy, C.H.; Lowe, D.J.; Preusser, F. (February 2007). "A terrestrial palynological record for the last two glacial cycles from southwestern New Zealand". Quaternary Science Reviews. 26 (3–4): 517–535. Bibcode:2007QSRv...26..517N. doi:10.1016/j.quascirev.2006.05.005.
  16. ^ Mangerud, Jan; Anderson, Svend T.; Berglund, Bjorn E.; Donner, Joakim J. (October 1, 1974). "Quaternary stratigraphy of Norden: a proposal for terminology and classification" (PDF). Boreas. 3 (3): 109–128. Bibcode:1974Borea...3..109M. doi:10.1111/j.1502-3885.1974.tb00669.x. Archived (PDF) from the original on February 16, 2020. Retrieved September 15, 2013.
  17. ^ Viau, André E.; Gajewski, Konrad; Fines, Philippe; Atkinson, David E.; Sawada, Michael C. (1 May 2002). "Widespread evidence of 1500 yr climate variability in North America during the past 14 000 yr". Geology. 30 (5): 455–458. Bibcode:2002Geo....30..455V. doi:10.1130/0091-7613(2002)030<0455:WEOYCV>2.0.CO;2.
  18. ^ Blackford, J. (1993). "Peat bogs as sources of proxy climatic data: Past approaches and future research" (PDF). Climate change and human impact on the landscape. Dordrecht: Springer. pp. 47–56. doi:10.1007/978-94-010-9176-3_5. ISBN 978-0-412-61860-4. Retrieved 20 November 2020.
  19. ^ Schrøder, N.; Højlund Pedersen, L.; Juel Bitsch, R. (2004). "10,000 years of climate change and human impact on the environment in the area surrounding Lejre". The Journal of Transdisciplinary Environmental Studies. 3 (1): 1–27.
  20. ^ "Middle Ages | Definition, Dates, Characteristics, & Facts". Encyclopædia Britannica. Archived from the original on 2021-06-11. Retrieved 2021-06-04.
  21. ^ Pearce, Fred (2007). With Speed and Violence. Beacon Press. p. 21. ISBN 978-0-8070-8576-9.
  22. ^ "Working Group on the "Anthropocene"". Subcommission on Quaternary Stratigraphy. International Commission on Stratigraphy. January 4, 2016. Archived from the original on February 17, 2016. Retrieved June 18, 2017.
  23. ^ Gray, Louise (October 7, 2009). "England is sinking while Scotland rises above sea levels, according to new study". The Daily Telegraph. Archived from the original on 2022-01-11. Retrieved June 10, 2014.
  24. ^ Lajeuness, Patrick; Allard, Michael (2003). "The Nastapoka drift belt, eastern Hudson Bay: implications of a stillstand of the Quebec-Labrador ice margin in the Tyrrell Sea at 8 ka BP" (PDF). Canadian Journal of Earth Sciences. 40 (1): 65–76. Bibcode:2003CaJES..40...65L. doi:10.1139/e02-085. Archived from the original (PDF) on 2004-03-22.
  25. ^ Bond, G.; et al. (1997). "A Pervasive Millennial-Scale Cycle in North Atlantic Holocene and Glacial Climates" (PDF). Science. 278 (5341): 1257–1266. Bibcode:1997Sci...278.1257B. doi:10.1126/science.278.5341.1257. S2CID 28963043. Archived from the original (PDF) on 2008-02-27.
  26. ^ Bond, G.; et al. (2001). "Persistent Solar Influence on North Atlantic Climate During the Holocene". Science. 294 (5549): 2130–2136. Bibcode:2001Sci...294.2130B. doi:10.1126/science.1065680. PMID 11739949. S2CID 38179371.
  27. ^ Bianchi, G.G.; McCave, I.N. (1999). "Holocene periodicity in North Atlantic climate and deep-ocean flow south of Iceland". Nature. 397 (6719): 515–517. Bibcode:1999Natur.397..515B. doi:10.1038/17362. S2CID 4304638.
  28. ^ Viau, A.E.; Gajewski, K.; Sawada, M.C.; Fines, P. (2006). "Millennial-scale temperature variations in North America during the Holocene". Journal of Geophysical Research. 111 (D9): D09102. Bibcode:2006JGRD..111.9102V. doi:10.1029/2005JD006031.
  29. ^ Debret, M.; Sebag, D.; Crosta, X.; Massei, N.; Petit, J.-R.; Chapron, E.; Bout-Roumazeilles, V. (2009). "Evidence from wavelet analysis for a mid-Holocene transition in global climate forcing" (PDF). Quaternary Science Reviews. 28 (25): 2675–2688. Bibcode:2009QSRv...28.2675D. doi:10.1016/j.quascirev.2009.06.005. S2CID 117917422. Archived (PDF) from the original on 2018-12-28. Retrieved 2018-12-16.
  30. ^ a b Kravchinsky, V.A.; Langereis, C.G.; Walker, S.D.; Dlusskiy, K.G.; White, D. (2013). "Discovery of Holocene millennial climate cycles in the Asian continental interior: Has the sun been governing the continental climate?". Global and Planetary Change. 110: 386–396. Bibcode:2013GPC...110..386K. doi:10.1016/j.gloplacha.2013.02.011.
  31. ^ Martin-Puertas, Celia; Hernandez, Armand; Pardo-Igúzquiza, Eulogio; Boyall, Laura; Brierley, Chris; Jiang, Zhiyi; Tjallingii, Rik; Blockley, Simon P. E.; Rodríguez-Tovar, Francisco Javier (23 March 2023). "Dampened predictable decadal North Atlantic climate fluctuations due to ice melting". Nature Geoscience. 16 (4): 357–362. Bibcode:2023NatGe..16..357M. doi:10.1038/s41561-023-01145-y. ISSN 1752-0908. S2CID 257735721. Retrieved 22 September 2023.
  32. ^ O'Brien, S. R.; Mayewski, P. A.; Meeker, L. D.; Meese, D. A.; Twickler, M. S.; Whitlow, S. I. (1995-12-22). "Complexity of Holocene Climate as Reconstructed from a Greenland Ice Core". Science. 270 (5244): 1962–1964. Bibcode:1995Sci...270.1962O. doi:10.1126/science.270.5244.1962. ISSN 0036-8075. S2CID 129199142.
  33. ^ Ruddiman, W. F.; Fuller, D. Q.; Kutzbach, J. E.; Tzedakis, P. C.; Kaplan, J. O.; Ellis, E. C.; Vavrus, S. J.; Roberts, C. N.; Fyfe, R.; He, F.; Lemmen, C.; Woodbridge, J. (15 February 2016). "Late Holocene climate: Natural or anthropogenic?". Reviews of Geophysics. 54 (1): 93–118. Bibcode:2016RvGeo..54...93R. doi:10.1002/2015RG000503. hdl:10026.1/8204. ISSN 8755-1209. S2CID 46451944.
  34. ^ a b Seip, Knut Lehre; Wang, Hui (3 March 2023). "Maximum Northern Hemisphere warming rates before and after 1880 during the Common Era". Theoretical and Applied Climatology. 152 (1–2): 307–319. Bibcode:2023ThApC.152..307S. doi:10.1007/s00704-023-04398-0. hdl:11250/3071271. ISSN 0177-798X. S2CID 257338719.
  35. ^ Degroot, Dagomar; Anchukaitis, Kevin J; Tierney, Jessica E; Riede, Felix; Manica, Andrea; Moesswilde, Emma; Gauthier, Nicolas (1 October 2022). "The history of climate and society: a review of the influence of climate change on the human past". Environmental Research Letters. 17 (10): 103001. Bibcode:2022ERL....17j3001D. doi:10.1088/1748-9326/ac8faa. hdl:10852/100641. ISSN 1748-9326. S2CID 252130680.
  36. ^ Zhang, David D.; Brecke, Peter; Lee, Harry F.; He, Yuan-Qing; Zhang, Jane (4 December 2007). "Global climate change, war, and population decline in recent human history". Proceedings of the National Academy of Sciences of the United States of America. 104 (49): 19214–19219. Bibcode:2007PNAS..10419214Z. doi:10.1073/pnas.0703073104. ISSN 0027-8424. PMC 2148270. PMID 18048343.
  37. ^ a b Wanner, Heinz; Beer, Jürg; Bütikofer, Jonathan; Crowley, Thomas J.; Cubasch, Ulrich; Flückiger, Jacqueline; Goosse, Hugues; Grosjean, Martin; Joos, Fortunat; Kaplan, Jed O.; Küttel, Marcel; Müller, Simon A.; Prentice, I. Colin; Solomina, Olga; Stocker, Thomas F. (October 2008). "Mid- to Late Holocene climate change: an overview". Quaternary Science Reviews. 27 (19): 1791–1828. Bibcode:2008QSRv...27.1791W. doi:10.1016/j.quascirev.2008.06.013. ISSN 0277-3791. Retrieved 27 September 2023.
  38. ^ Hoek, Wim Z.; Bos, Johanna A. A. (August 2007). "Early Holocene climate oscillations—causes and consequences". Quaternary Science Reviews. Early Holocene climate oscillations - causes and consequences. 26 (15): 1901–1906. Bibcode:2007QSRv...26.1901H. doi:10.1016/j.quascirev.2007.06.008. ISSN 0277-3791. Retrieved 27 September 2023.
  39. ^ Gao, Fuyuan; Jia, Jia; Xia, Dunsheng; Lu, Caichen; Lu, Hao; Wang, Youjun; Liu, Hao; Ma, Yapeng; Li, Kaiming (15 March 2019). "Asynchronous Holocene Climate Optimum across mid-latitude Asia". Palaeogeography, Palaeoclimatology, Palaeoecology. 518: 206–214. Bibcode:2019PPP...518..206G. doi:10.1016/j.palaeo.2019.01.012. S2CID 135199089. Retrieved 10 September 2023.
  40. ^ Guiot, Joël (March 2012). "A robust spatial reconstruction of April to September temperature in Europe: Comparisons between the medieval period and the recent warming with a focus on extreme values". Global and Planetary Change. 84–85: 14–22. Bibcode:2012GPC....84...14G. doi:10.1016/j.gloplacha.2011.07.007.
  41. ^ Wanner, H.; Mercolli, L.; Grosjean, M.; Ritz, S. P. (17 October 2014). "Holocene climate variability and change; a data-based review". Journal of the Geological Society. 172 (2): 254–263. doi:10.1144/jgs2013-101. ISSN 0016-7649. S2CID 73548216. Retrieved 27 September 2023.
  42. ^ Duan, Jianping; Zhang, Qi-Bin (27 October 2014). "A 449 year warm season temperature reconstruction in the southeastern Tibetan Plateau and its relation to solar activity: Temperature reconstruction in the Tibet". Journal of Geophysical Research: Atmospheres. 119 (20): 11, 578–11, 592. doi:10.1002/2014JD022422. S2CID 128906290.
  43. ^ Benestad, R. E.; Schmidt, G. A. (27 July 2009). "Solar trends and global warming". Journal of Geophysical Research: Atmospheres. 114 (D14). Bibcode:2009JGRD..11414101B. doi:10.1029/2008JD011639. ISSN 0148-0227.
  44. ^ Perry, Charles A.; Hsu, Kenneth J. (7 November 2000). "Geophysical, archaeological, and historical evidence support a solar-output model for climate change". Proceedings of the National Academy of Sciences of the United States of America. 97 (23): 12433–12438. Bibcode:2000PNAS...9712433P. doi:10.1073/pnas.230423297. ISSN 0027-8424. PMC 18780. PMID 11050181.
  45. ^ Gerdes, G; Petzelberger, B. E. M; Scholz-Böttcher, B. M; Streif, H (1 January 2003). "The record of climatic change in the geological archives of shallow marine, coastal, and adjacent lowland areas of Northern Germany". Quaternary Science Reviews. Environmental response to climate and human impact in central Eur ope during the last 15000 years - a German contribution to PAGES-PEPIII. 22 (1): 101–124. Bibcode:2003QSRv...22..101G. doi:10.1016/S0277-3791(02)00183-X. ISSN 0277-3791. Retrieved 27 October 2023.
  46. ^ Zhang, Xu (Yvon); Bajard, Manon; Bouchez, Julien; Sabatier, Pierre; Poulenard, Jérôme; Arnaud, Fabien; Crouzet, Christian; Kuessner, Marie; Dellinger, Mathieu; Gaillardet, Jérôme (2023-12-15). "Evolution of the alpine Critical Zone since the Last Glacial Period using Li isotopes from lake sediments". Earth and Planetary Science Letters. 624: 118463. Bibcode:2023E&PSL.62418463Z. doi:10.1016/j.epsl.2023.118463. ISSN 0012-821X.
  47. ^ Armitage, Simon J.; Bristow, Charlie S.; Drake, Nick A. (14 July 2015). "West African monsoon dynamics inferred from abrupt fluctuations of Lake Mega-Chad". Proceedings of the National Academy of Sciences of the United States of America. 112 (28): 8543–8548. Bibcode:2015PNAS..112.8543A. doi:10.1073/pnas.1417655112. ISSN 0027-8424. PMC 4507243. PMID 26124133.
  48. ^ Depreux, Bruno; Lefèvre, David; Berger, Jean-François; Segaoui, Fatima; Boudad, Larbi; El Harradji, Abderrahmane; Degeai, Jean-Philippe; Limondin-Lozouet, Nicole (1 March 2021). "Alluvial records of the African Humid Period from the NW African highlands (Moulouya basin, NE Morocco)". Quaternary Science Reviews. 255: 106807. Bibcode:2021QSRv..25506807D. doi:10.1016/j.quascirev.2021.106807. ISSN 0277-3791. S2CID 233792780.
  49. ^ Sha, Lijuan; Ait Brahim, Yassine; Wassenburg, Jasper A.; Yin, Jianjun; Peros, Matthew; Cruz, Francisco W.; Cai, Yanjun; Li, Hanying; Du, Wenjing; Zhang, Haiwei; Edwards, R. Lawrence; Cheng, Hai (16 December 2019). "How Far North Did the African Monsoon Fringe Expand During the African Humid Period? Insights From Southwest Moroccan Speleothems". Geophysical Research Letters. 46 (23): 14093–14102. Bibcode:2019GeoRL..4614093S. doi:10.1029/2019GL084879. ISSN 0094-8276. S2CID 213015081.
  50. ^ Manning, Katie; Timpson, Adrian (October 2014). "The demographic response to Holocene climate change in the Sahara". Quaternary Science Reviews. 101: 28–35. Bibcode:2014QSRv..101...28M. doi:10.1016/j.quascirev.2014.07.003. S2CID 54923700.
  51. ^ Adkins, Jess; deMenocal, Peter; Eshel, Gidon (20 October 2006). "The "African humid period" and the record of marine upwelling from excess 230 Th in Ocean Drilling Program Hole 658C: Th NORMALIZED FLUXES OFF NORTH AFRICA". Paleoceanography and Paleoclimatology. 21 (4). doi:10.1029/2005PA001200.
  52. ^ Forman, Steven L.; Wright, David K.; Bloszies, Christopher (1 August 2014). "Variations in water level for Lake Turkana in the past 8500 years near Mt. Porr, Kenya and the transition from the African Humid Period to Holocene aridity". Quaternary Science Reviews. 97: 84–101. Bibcode:2014QSRv...97...84F. doi:10.1016/j.quascirev.2014.05.005. ISSN 0277-3791. Retrieved 22 September 2023.
  53. ^ Sitoe, Sandra Raúl; Risberg, Jan; Norström, Elin; Westerberg, Lars-Ove (1 November 2017). "Late Holocene sea-level changes and paleoclimate recorded in Lake Lungué, southern Mozambique". Palaeogeography, Palaeoclimatology, Palaeoecology. 485: 305–315. Bibcode:2017PPP...485..305S. doi:10.1016/j.palaeo.2017.06.022. ISSN 0031-0182. Retrieved 22 November 2023.
  54. ^ Lancaster, N. (1 May 1989). "Late Quaternary paleoenvironments in the southwestern Kalahari". Palaeogeography, Palaeoclimatology, Palaeoecology. 70 (4): 367–376. Bibcode:1989PPP....70..367L. doi:10.1016/0031-0182(89)90114-4. ISSN 0031-0182. Retrieved 15 September 2023.
  55. ^ Giaime, Matthieu; Artzy, Michal; Jol, Harry M.; Salmon, Yossi; López, Gloria I.; Abu Hamid, Amani (1 May 2022). "Refining Late-Holocene environmental changes of the Akko coastal plain and its impacts on the settlement and anchorage patterns of Tel Akko (Israel)". Marine Geology. 447: 106778. Bibcode:2022MGeol.44706778G. doi:10.1016/j.margeo.2022.106778. ISSN 0025-3227. S2CID 247636727.
  56. ^ Rao, Zhiguo; Wu, Dandan; Shi, Fuxi; Guo, Haichun; Cao, Jiantao; Chen, Fahu (1 April 2019). "Reconciling the 'westerlies' and 'monsoon' models: A new hypothesis for the Holocene moisture evolution of the Xinjiang region, NW China". Earth-Science Reviews. 191: 263–272. Bibcode:2019ESRv..191..263R. doi:10.1016/j.earscirev.2019.03.002. ISSN 0012-8252. S2CID 134712945. Retrieved 15 September 2023.
  57. ^ Huang, Xiao-zhong; Chen, Chun-zhu; Jia, Wan-na; An, Cheng-bang; Zhou, Ai-feng; Zhang, Jia-wu; Jin, Ming; Xia, Dun-sheng; Chen, Fa-hu; Grimm, Eric C. (15 August 2015). "Vegetation and climate history reconstructed from an alpine lake in central Tienshan Mountains since 8.5ka BP". Palaeogeography, Palaeoclimatology, Palaeoecology. 432: 36–48. Bibcode:2015PPP...432...36H. doi:10.1016/j.palaeo.2015.04.027. ISSN 0031-0182. Retrieved 10 September 2023.
  58. ^ Long, Hao; Shen, Ji; Chen, Jianhui; Tsukamoto, Sumiko; Yang, Linhai; Cheng, Hongyi; Frechen, Manfred (15 October 2017). "Holocene moisture variations over the arid central Asia revealed by a comprehensive sand-dune record from the central Tian Shan, NW China". Quaternary Science Reviews. 174: 13–32. Bibcode:2017QSRv..174...13L. doi:10.1016/j.quascirev.2017.08.024. ISSN 0277-3791. Retrieved 10 September 2023.
  59. ^ Wünnemann, Bernd; Yan, Dada; Andersen, Nils; Riedel, Frank; Zhang, Yongzhan; Sun, Qianli; Hoelzmann, Philipp (15 November 2018). "A 14 ka high-resolution δ18O lake record reveals a paradigm shift for the process-based reconstruction of hydroclimate on the northern Tibetan Plateau". Quaternary Science Reviews. 200: 65–84. Bibcode:2018QSRv..200...65W. doi:10.1016/j.quascirev.2018.09.040. ISSN 0277-3791. S2CID 134520306. Retrieved 10 September 2023.
  60. ^ Demske, Dieter; Tarasov, Pavel E.; Wünnemann, Bernd; Riedel, Frank (15 August 2009). "Late glacial and Holocene vegetation, Indian monsoon and westerly circulation in the Trans-Himalaya recorded in the lacustrine pollen sequence from Tso Kar, Ladakh, NW India". Palaeogeography, Palaeoclimatology, Palaeoecology. 279 (3–4): 172–185. Bibcode:2009PPP...279..172D. doi:10.1016/j.palaeo.2009.05.008. Retrieved 27 September 2023.
  61. ^ Singh, Dhruv Sen; Gupta, Anil K.; Sangode, S. J.; Clemens, Steven C.; Prakasam, M.; Srivastava, Priyeshu; Prajapati, Shailendra K. (12 June 2015). "Multiproxy record of monsoon variability from the Ganga Plain during 400–1200 A.D." Quaternary International. Updated Quaternary Climatic Research in parts of the Third Pole Selected papers from the HOPE-2013 conference, Nainital, India. 371: 157–163. Bibcode:2015QuInt.371..157S. doi:10.1016/j.quaint.2015.02.040. ISSN 1040-6182. Retrieved 10 September 2023.
  62. ^ Menzel, Philip; Gaye, Birgit; Mishra, Praveen K.; Anoop, Ambili; Basavaiah, Nathani; Marwan, Norbert; Plessen, Birgit; Prasad, Sushma; Riedel, Nils; Stebich, Martina; Wiesner, Martin G. (15 September 2014). "Linking Holocene drying trends from Lonar Lake in monsoonal central India to North Atlantic cooling events". Palaeogeography, Palaeoclimatology, Palaeoecology. 410: 164–178. Bibcode:2014PPP...410..164M. doi:10.1016/j.palaeo.2014.05.044. ISSN 0031-0182. Retrieved 15 September 2023.
  63. ^ Shaji, Jithu; Banerji, Upasana S.; Maya, K.; Joshi, Kumar Batuk; Dabhi, Ankur J.; Bharti, Nisha; Bhushan, Ravi; Padmalal, D. (30 December 2022). "Holocene monsoon and sea-level variability from coastal lowlands of Kerala, SW India". Quaternary International. 642: 48–62. Bibcode:2022QuInt.642...48S. doi:10.1016/j.quaint.2022.03.005. S2CID 247553867. Retrieved 10 September 2023.
  64. ^ Guo, Zhengtang; Petit-Maire, Nicole; Kröpelin, Stefan (November 2000). "Holocene non-orbital climatic events in present-day arid areas of northern Africa and China". Global and Planetary Change. 26 (1–3): 97–103. Bibcode:2000GPC....26...97G. doi:10.1016/S0921-8181(00)00037-0. Retrieved 10 September 2023.
  65. ^ Zheng, Yanhong; Yu, Shi-Yong; Fan, Tongyu; Oppenheimer, Clive; Yu, Xuefeng; Liu, Zhao; Xian, Feng; Liu, Zhen; Li, Jianyong; Li, Jiahao (15 July 2021). "Prolonged cooling interrupted the Bronze Age cultures in northeastern China 3500 years ago". Palaeogeography, Palaeoclimatology, Palaeoecology. 574: 110461. Bibcode:2021PPP...57410461Z. doi:10.1016/j.palaeo.2021.110461. ISSN 0031-0182. S2CID 236229299. Retrieved 15 October 2023.
  66. ^ a b Zhao, Yan; Yu, Zicheng; Chen, Fahu; Zhang, Jiawu; Yang, Bao (1 December 2009). "Vegetation response to Holocene climate change in monsoon-influenced region of China". Earth-Science Reviews. 97 (1): 242–256. Bibcode:2009ESRv...97..242Z. doi:10.1016/j.earscirev.2009.10.007. ISSN 0012-8252. Retrieved 10 September 2023.
  67. ^ Jia, Guodong; Bai, Yang; Yang, Xiaoqiang; Xie, Luhua; Wei, Gangjian; Ouyang, Tingping; Chu, Guoqiang; Liu, Zhonghui; Peng, Ping'an (1 March 2015). "Biogeochemical evidence of Holocene East Asian summer and winter monsoon variability from a tropical maar lake in southern China". Quaternary Science Reviews. 111: 51–61. doi:10.1016/j.quascirev.2015.01.002. ISSN 0277-3791. Retrieved 10 September 2023.
  68. ^ Park, Jungjae (1 March 2017). "Solar and tropical ocean forcing of late-Holocene climate change in coastal East Asia". Palaeogeography, Palaeoclimatology, Palaeoecology. 469: 74–83. Bibcode:2017PPP...469...74P. doi:10.1016/j.palaeo.2017.01.005. ISSN 0031-0182. Retrieved 15 September 2023.
  69. ^ Zhang, Hui; Liu, Shengfa; Wu, Kaikai; Cao, Peng; Pan, Hui-Juan; Wang, Hongmin; Cui, Jingjing; Li, Jingrui; Khokiattiwong, Somkiat; Kornkanitnan, Narumol; Shi, Xuefa (20 August 2022). "Evolution of sedimentary environment in the Gulf of Thailand since the last deglaciation". Quaternary International. Understanding the Late Quaternary Paleomonsoon and Paleoenvironmental Shifts of Asia. 629: 36–43. Bibcode:2022QuInt.629...36Z. doi:10.1016/j.quaint.2021.02.018. ISSN 1040-6182. S2CID 233897984. Retrieved 15 September 2023.
  70. ^ Steinman, Byron A.; Pompeani, David P.; Abbott, Mark B.; Ortiz, Joseph D.; Stansell, Nathan D.; Finkenbinder, Matthew S.; Mihindukulasooriya, Lorita N.; Hillman, Aubrey L. (15 June 2016). "Oxygen isotope records of Holocene climate variability in the Pacific Northwest". Quaternary Science Reviews. 142: 40–60. Bibcode:2016QSRv..142...40S. doi:10.1016/j.quascirev.2016.04.012. ISSN 0277-3791.
  71. ^ Perner, Kerstin; Moros, Matthias; Lloyd, Jeremy M.; Jansen, Eystein; Stein, Rüdiger (1 December 2015). "Mid to late Holocene strengthening of the East Greenland Current linked to warm subsurface Atlantic water". Quaternary Science Reviews. 129: 296–307. Bibcode:2015QSRv..129..296P. doi:10.1016/j.quascirev.2015.10.007. ISSN 0277-3791. S2CID 129732336. Retrieved 11 September 2023.
  72. ^ Mensing, Scott A.; Sharpe, Saxon E.; Tunno, Irene; Sada, Don W.; Thomas, Jim M.; Starratt, Scott; Smith, Jeremy (15 October 2013). "The Late Holocene Dry Period: multiproxy evidence for an extended drought between 2800 and 1850 cal yr BP across the central Great Basin, USA". Quaternary Science Reviews. 78: 266–282. Bibcode:2013QSRv...78..266M. doi:10.1016/j.quascirev.2013.08.010. ISSN 0277-3791. Retrieved 10 September 2023.
  73. ^ Shuman, Bryan N.; Marsicek, Jeremiah (1 June 2016). "The structure of Holocene climate change in mid-latitude North America". Quaternary Science Reviews. 141: 38–51. Bibcode:2016QSRv..141...38S. doi:10.1016/j.quascirev.2016.03.009. ISSN 0277-3791.
  74. ^ Fontes, Neuza Araújo; Moraes, Caio A.; Cohen, Marcelo C L; Alves, Igor Charles C.; França, Marlon Carlos; Pessenda, Luiz C R; Francisquini, Mariah Izar; Bendassolli, José Albertino; Macario, Kita; Mayle, Francis (February 2017). "The Impacts of the Middle Holocene High Sea-Level Stand and Climatic Changes on Mangroves of the Jucuruçu River, Southern Bahia – Northeastern Brazil". Radiocarbon. 59 (1): 215–230. Bibcode:2017Radcb..59..215F. doi:10.1017/RDC.2017.6. ISSN 0033-8222. S2CID 133047191.
  75. ^ Angulo, Rodolfo J.; Lessa, Guilherme C.; Souza, Maria Cristina de (1 March 2006). "A critical review of mid- to late-Holocene sea-level fluctuations on the eastern Brazilian coastline". Quaternary Science Reviews. 25 (5): 486–506. Bibcode:2006QSRv...25..486A. doi:10.1016/j.quascirev.2005.03.008. ISSN 0277-3791. Retrieved 17 September 2023.
  76. ^ Angulo, Rodolfo José; de Souza, Maria Cristina; da Camara Rosa, Maria Luiza Correa; Caron, Felipe; Barboza, Eduardo G.; Costa, Mirella Borba Santos Ferreira; Macedo, Eduardo; Vital, Helenice; Gomes, Moab Praxedes; Garcia, Khalil Bow Ltaif (1 May 2022). "Paleo-sea levels, Late-Holocene evolution, and a new interpretation of the boulders at the Rocas Atoll, southwestern Equatorial Atlantic". Marine Geology. 447: 106780. Bibcode:2022MGeol.44706780A. doi:10.1016/j.margeo.2022.106780. ISSN 0025-3227. S2CID 247822701. Retrieved 17 September 2023.
  77. ^ Eroglu, Deniz; McRobie, Fiona H.; Ozken, Ibrahim; Stemler, Thomas; Wyrwoll, Karl-Heinz; Breitenbach, Sebastian F. M.; Marwan, Norbert; Kurths, Jürgen (26 September 2016). "See–saw relationship of the Holocene East Asian–Australian summer monsoon". Nature Communications. 7 (1): 12929. Bibcode:2016NatCo...712929E. doi:10.1038/ncomms12929. ISSN 2041-1723. PMC 5052686. PMID 27666662.
  78. ^ Prebble, J. G.; Bostock, H. C.; Cortese, G.; Lorrey, A. M.; Hayward, B. W.; Calvo, E.; Northcote, L. C.; Scott, G. H.; Neil, H. L. (August 2017). "Evidence for a Holocene Climatic Optimum in the southwest Pacific: A multiproxy study: Holocene Optimum in SW Pacific". Paleoceanography. 32 (8): 763–779. doi:10.1002/2016PA003065. hdl:10261/155815.
  79. ^ Orpin, A. R.; Carter, L.; Page, M. J.; Cochran, U. A.; Trustrum, N. A.; Gomez, B.; Palmer, A. S.; Mildenhall, D. C.; Rogers, K. M.; Brackley, H. L.; Northcote, L. (15 April 2010). "Holocene sedimentary record from Lake Tutira: A template for upland watershed erosion proximal to the Waipaoa Sedimentary System, northeastern New Zealand". Marine Geology. From mountain source to ocean sink – the passage of sediment across an active margin, Waipaoa Sedimentary System, New Zealand. 270 (1): 11–29. Bibcode:2010MGeol.270...11O. doi:10.1016/j.margeo.2009.10.022. ISSN 0025-3227. Retrieved 11 September 2023.
  80. ^ Zhang, Zhaohui; Leduc, Guillaume; Sachs, Julian P. (15 October 2014). "El Niño evolution during the Holocene revealed by a biomarker rain gauge in the Galápagos Islands". Earth and Planetary Science Letters. 404: 420–434. Bibcode:2014E&PSL.404..420Z. doi:10.1016/j.epsl.2014.07.013. ISSN 0012-821X.
  81. ^ Lemoine, Rhys Taylor; Buitenwerf, Robert; Svenning, Jens-Christian (2023-12-01). "Megafauna extinctions in the late-Quaternary are linked to human range expansion, not climate change". Anthropocene. 44: 100403. Bibcode:2023Anthr..4400403L. doi:10.1016/j.ancene.2023.100403. ISSN 2213-3054.
  82. ^ Willerslev, Eske; Hansen, Anders J.; Binladen, Jonas; Brand, Tina B.; Gilbert, M. Thomas P.; Shapiro, Beth; Bunce, Michael; Wiuf, Carsten; Gilichinsky, David A.; Cooper, Alan (2 May 2003). "Diverse Plant and Animal Genetic Records from Holocene and Pleistocene Sediments". Science. 300 (5620): 791–795. Bibcode:2003Sci...300..791W. doi:10.1126/science.1084114. ISSN 0036-8075. PMID 12702808. S2CID 1222227.
  83. ^ Singh, Ashbindu (2005). One Planet, Many People: Atlas of Our Changing Environment. United Nations Environment Programme. p. 4. ISBN 978-9280725711. Archived from the original on 2020-01-02. Retrieved 2017-06-28.
  84. ^ Barber, D.C; Dyke, A.; Hillaire-Marcel, C.; Jennings, A.E.; Andrews, J.T.; Kerwin, M.W.; Bilodeau, G.; McNeely, R.; Southon, J.; Morehead, M.D.; Gagnon, J.-M. (July 22, 1999). "Forcing of the cold event of 8,200 years ago by catastrophic drainage of Laurentide lakes". Nature. 400 (6742): 344–348. Bibcode:1999Natur.400..344B. doi:10.1038/22504. S2CID 4426918. Retrieved 11 September 2023.
  85. ^ a b c Li, Yong-Xiang; Törnqvist, Torbjörn E.; Nevitt, Johanna M.; Kohl, Barry (15 January 2012). "Synchronizing a sea-level jump, final Lake Agassiz drainage, and abrupt cooling 8200years ago". Earth and Planetary Science Letters. Sea Level and Ice Sheet Evolution: A PALSEA Special Edition. 315–316: 41–50. Bibcode:2012E&PSL.315...41L. doi:10.1016/j.epsl.2011.05.034. ISSN 0012-821X. Retrieved 15 October 2023.
  86. ^ Rohling, Eelco J.; Pälike, Heiko (21 April 2005). "Centennial-scale climate cooling with a sudden event around 8,200 years ago". Nature. 434 (7036): 975–979. Bibcode:2005Natur.434..975R. doi:10.1038/nature03421. PMID 15846336. S2CID 4394638. Retrieved 15 October 2023.
  87. ^ Chisholm, Hugh, ed. (1911). "Jericho" . Encyclopædia Britannica (11th ed.). Cambridge University Press.
  88. ^ Curry, Andrew (November 2008). "Göbekli Tepe: The World's First Temple?". Smithsonian Magazine. Archived from the original on March 17, 2009. Retrieved March 14, 2009.
  89. ^ a b Gupta, Anil K. (10 July 2004). "Origin of agriculture and domestication of plants and animals linked to early Holocene climate amelioration". Current Science. 87 (1): 54–59. ISSN 0011-3891. JSTOR 24107979. Retrieved 11 September 2023.
  90. ^ Riris, Philip; Arroyo-Kalin, Manuel (9 May 2019). "Widespread population decline in South America correlates with mid-Holocene climate change". Scientific Reports. 9 (1): 6850. Bibcode:2019NatSR...9.6850R. doi:10.1038/s41598-019-43086-w. ISSN 2045-2322. PMC 6509208. PMID 31073131. Retrieved 15 October 2023.
  91. ^ Brenner, Mark; Hodell, David A.; Rosenmeier, Michael F.; Curtis, Jason H.; Binford, Michael W.; Abbott, Mark B. (2001-01-01), Markgraf, Vera (ed.), "Chapter 6 - Abrupt Climate Change and Pre-Columbian Cultural Collapse", Interhemispheric Climate Linkages, San Diego: Academic Press, pp. 87–103, doi:10.1016/b978-012472670-3/50009-4, ISBN 978-0-12-472670-3, retrieved 2022-04-23
  92. ^ Kim, Habeom; Lee, Gyoung-Ah; Crema, Enrico R. (10 December 2021). "Bayesian analyses question the role of climate in Chulmun demography". Scientific Reports. 11 (1): 23797. Bibcode:2021NatSR..1123797K. doi:10.1038/s41598-021-03180-4. ISSN 2045-2322. PMC 8664936. PMID 34893660.
  93. ^ Wagler, Ron (2011). "The Anthropocene Mass Extinction: An Emerging Curriculum Theme for Science Educators". The American Biology Teacher. 73 (2): 78–83. doi:10.1525/abt.2011.73.2.5. S2CID 86352610.
  94. ^ Walsh, Alistair (January 11, 2022). "What to expect from the world's sixth mass extinction". Deutsche Welle. Retrieved February 5, 2022.
  95. ^ Ripple WJ, Wolf C, Newsome TM, Galetti M, Alamgir M, Crist E, Mahmoud MI, Laurance WF (13 November 2017). "World Scientists' Warning to Humanity: A Second Notice" (PDF). BioScience. 67 (12): 1026–1028. doi:10.1093/biosci/bix125. Archived from the original (PDF) on 15 December 2019. Retrieved 4 October 2022. Moreover, we have unleashed a mass extinction event, the sixth in roughly 540 million years, wherein many current life forms could be annihilated or at least committed to extinction by the end of this century.
  96. ^ a b Ceballos, Gerardo; Ehrlich, Paul R. (8 June 2018). "The misunderstood sixth mass extinction". Science. 360 (6393): 1080–1081. Bibcode:2018Sci...360.1080C. doi:10.1126/science.aau0191. OCLC 7673137938. PMID 29880679. S2CID 46984172.
  97. ^ Dirzo, Rodolfo; Young, Hillary S.; Galetti, Mauro; Ceballos, Gerardo; Isaac, Nick J. B.; Collen, Ben (2014). "Defaunation in the Anthropocene" (PDF). Science. 345 (6195): 401–406. Bibcode:2014Sci...345..401D. doi:10.1126/science.1251817. PMID 25061202. S2CID 206555761. In the past 500 years, humans have triggered a wave of extinction, threat, and local population declines that may be comparable in both rate and magnitude with the five previous mass extinctions of Earth's history.
  98. ^ Cowie, Robert H.; Bouchet, Philippe; Fontaine, Benoît (2022). "The Sixth Mass Extinction: fact, fiction or speculation?". Biological Reviews. 97 (2): 640–663. doi:10.1111/brv.12816. PMC 9786292. PMID 35014169. S2CID 245889833.
  99. ^ Guy, Jack (September 30, 2020). "Around 40% of the world's plant species are threatened with extinction". CNN. Retrieved September 1, 2021.
  100. ^ Hollingsworth, Julia (June 11, 2019). "Almost 600 plant species have become extinct in the last 250 years". CNN. Retrieved January 14, 2020. The research -- published Monday in Nature, Ecology & Evolution journal -- found that 571 plant species have disappeared from the wild worldwide, and that plant extinction is occurring up to 500 times faster than the rate it would without human intervention.
  101. ^ Watts, Jonathan (August 31, 2021). "Up to half of world's wild tree species could be at risk of extinction". The Guardian. Retrieved September 1, 2021.
  102. ^ De Vos, Jurriaan M.; Joppa, Lucas N.; Gittleman, John L.; Stephens, Patrick R.; Pimm, Stuart L. (2014-08-26). "Estimating the normal background rate of species extinction" (PDF). Conservation Biology (in Spanish). 29 (2): 452–462. doi:10.1111/cobi.12380. ISSN 0888-8892. PMID 25159086. S2CID 19121609.
  103. ^ Pimm, S. L.; Jenkins, C. N.; Abell, R.; Brooks, T. M.; Gittleman, J. L.; Joppa, L. N.; Raven, P. H.; Roberts, C. M.; Sexton, J. O. (30 May 2014). "The biodiversity of species and their rates of extinction, distribution, and protection" (PDF). Science. 344 (6187): 1246752. doi:10.1126/science.1246752. PMID 24876501. S2CID 206552746. The overarching driver of species extinction is human population growth and increasing per capita consumption.

Further reading edit

External links edit