African humid period: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
m Disambiguating links to Crater lake (link changed to Volcanic crater lake) using DisamAssist.
Expanding article
Line 363: Line 363:
The shrinkage of subtropical wetlands probably led to a drop in atmospheric [[methane]] concentrations between 5,500 and 5,000 years ago, before [[boreal ecosystem|boreal]] wetlands expanded and offset the loss of subtropical wetlands, leading to a return of higher atmospheric methane concentrations.{{sfn|Menocal|Ortiz|Guilderson|Adkins|2000|p=355}} Conversely, [[atmospheric carbon dioxide]] increased after about 7,000 years as the biosphere began releasing carbon in response to increasing aridity.<ref name="Bar-Matthews2003" />
The shrinkage of subtropical wetlands probably led to a drop in atmospheric [[methane]] concentrations between 5,500 and 5,000 years ago, before [[boreal ecosystem|boreal]] wetlands expanded and offset the loss of subtropical wetlands, leading to a return of higher atmospheric methane concentrations.{{sfn|Menocal|Ortiz|Guilderson|Adkins|2000|p=355}} Conversely, [[atmospheric carbon dioxide]] increased after about 7,000 years as the biosphere began releasing carbon in response to increasing aridity.<ref name="Bar-Matthews2003" />
[[File:Dust Storms from Africa's Bodele Depression, Natural Hazards DVIDS848674.jpg|thumb|Dust originating in the Bodele depression]]
[[File:Dust Storms from Africa's Bodele Depression, Natural Hazards DVIDS848674.jpg|thumb|Dust originating in the Bodele depression]]
A sudden increase in the amount of land-originating dust in an oceanic [[drill core]] off [[Ras Nouadhibou|Cape Blanc]], [[Mauritania]], has been interpreted as reflecting the end of the African humid period 5,500 years ago occurring in only a few centuries.{{sfn|Adkins|Menocal|Eshel|2006|p=2}} Potentially, dried up lake basins became an important source for dust.{{sfn|Cole|Goldstein|Menocal|Hemming|2009|p=264}} Today, the Sahara is the single largest source of dust in the world, with far ranging effects on ecosystems and climate.{{sfn|Zielhofer|Suchodoletz|Fletcher|Schneider|2017|p=119}}
A sudden increase in the amount of land-originating dust in an oceanic [[drill core]] off [[Ras Nouadhibou|Cape Blanc]], [[Mauritania]], has been interpreted as reflecting the end of the African humid period 5,500 years ago occurring in only a few centuries.{{sfn|Adkins|Menocal|Eshel|2006|p=2}} Potentially, dried up lake basins became an important source for dust.{{sfn|Cole|Goldstein|Menocal|Hemming|2009|p=264}} Today, the Sahara is the single largest source of dust in the world, with far ranging effects on climate and ecosystems,{{sfn|Zielhofer|Suchodoletz|Fletcher|Schneider|2017|p=119}} such as the growth of the [[Amazon rainforest]].<ref name="D'Odorico2006" />


The period 5,500–5,000 years ago also witnessed major changes in global climate, including the onset of global cooling in the form of the [[Neoglacial]].{{sfn|Muschitiello|Zhang|Sundqvist|Davies|2015|p=91}} In one climate model, the desertification of the Sahara at the end of the African humid period reduces the amount of heat transported in the atmosphere and ocean towards the poles, inducing cooling of {{convert|1|-|2|C-change}} especially in winter in the [[Arctic]] and an expansion of [[sea ice]]. Reconstructed temperatures in the Arctic indeed show a cooling, although less pronounced than in the climate model.{{sfn|Muschitiello|Zhang|Sundqvist|Davies|2015|p=93}} Further, this climate transition in the climate model is accompanied by increased negative [[Arctic Oscillation]] states, a weaker [[subpolar gyre]] and increased precipitation and [[cold air outbreak]]s in much of Europe; such changes have also been observed in paleoclimate data.{{sfn|Muschitiello|Zhang|Sundqvist|Davies|2015|pp=94-95}} These findings imply that the vegetation state of the Sahara influences the Northern Hemisphere climate.{{sfn|Muschitiello|Zhang|Sundqvist|Davies|2015|p=96}} In turn, this high latitude cooling may have further reduced precipitation over Africa.{{sfn|Schefuß|Roche|Skonieczny|Mulitza|2017|p=9}}
The period 5,500–5,000 years ago also witnessed major changes in global climate, including the onset of global cooling in the form of the [[Neoglacial]].{{sfn|Muschitiello|Zhang|Sundqvist|Davies|2015|p=91}} In one climate model, the desertification of the Sahara at the end of the African humid period reduces the amount of heat transported in the atmosphere and ocean towards the poles, inducing cooling of {{convert|1|-|2|C-change}} especially in winter in the [[Arctic]] and an expansion of [[sea ice]]. Reconstructed temperatures in the Arctic indeed show a cooling, although less pronounced than in the climate model.{{sfn|Muschitiello|Zhang|Sundqvist|Davies|2015|p=93}} Further, this climate transition in the climate model is accompanied by increased negative [[Arctic Oscillation]] states, a weaker [[subpolar gyre]] and increased precipitation and [[cold air outbreak]]s in much of Europe; such changes have also been observed in paleoclimate data.{{sfn|Muschitiello|Zhang|Sundqvist|Davies|2015|pp=94-95}} These findings imply that the vegetation state of the Sahara influences the Northern Hemisphere climate.{{sfn|Muschitiello|Zhang|Sundqvist|Davies|2015|p=96}} In turn, this high latitude cooling may have further reduced precipitation over Africa.{{sfn|Schefuß|Roche|Skonieczny|Mulitza|2017|p=9}}
Line 391: Line 391:
== References ==
== References ==
{{Reflist|refs=
{{Reflist|refs=
<ref name="D'Odorico2006">{{Cite book|url=http://link.springer.com/10.1007/1-4020-4260-4|title=Dryland Ecohydrology|date=2006|publisher=Kluwer Academic Publishers|isbn=978-1-4020-4259-1|editor-last=D'Odorico|editor-first=Paolo|location=Dordrecht|language=en|doi=10.1007/1-4020-4260-4|editor-last2=Porporato|editor-first2=Amilcare|p=589}}</ref>
<ref name="Fitchett2017">{{Cite journal|last=Fitchett|first=Jennifer M.|last2=Grab|first2=Stefan W.|last3=Bamford|first3=Marion K.|last4=Mackay|first4=Anson W.|date=2017-09-02|title=Late Quaternary research in southern Africa: progress, challenges and future trajectories|url=https://doi.org/10.1080/0035919X.2017.1297966|journal=Transactions of the Royal Society of South Africa|volume=72|issue=3|page=284|doi=10.1080/0035919X.2017.1297966|issn=0035-919X}}</ref>
<ref name="Fitchett2017">{{Cite journal|last=Fitchett|first=Jennifer M.|last2=Grab|first2=Stefan W.|last3=Bamford|first3=Marion K.|last4=Mackay|first4=Anson W.|date=2017-09-02|title=Late Quaternary research in southern Africa: progress, challenges and future trajectories|url=https://doi.org/10.1080/0035919X.2017.1297966|journal=Transactions of the Royal Society of South Africa|volume=72|issue=3|page=284|doi=10.1080/0035919X.2017.1297966|issn=0035-919X}}</ref>
<ref name="Loakes2017">{{Cite journal|last=Loakes|first=Katie|date=2017-01-02|title=Late Quaternary palaeolimnology and environmental change in the South Wollo Highlands|url=https://doi.org/10.1080/0067270X.2016.1259821|journal=Azania: Archaeological Research in Africa|volume=52|issue=1|page=131|doi=10.1080/0067270X.2016.1259821|issn=0067-270X}}</ref>
<ref name="Loakes2017">{{Cite journal|last=Loakes|first=Katie|date=2017-01-02|title=Late Quaternary palaeolimnology and environmental change in the South Wollo Highlands|url=https://doi.org/10.1080/0067270X.2016.1259821|journal=Azania: Archaeological Research in Africa|volume=52|issue=1|page=131|doi=10.1080/0067270X.2016.1259821|issn=0067-270X}}</ref>

Revision as of 15:26, 23 November 2019

The Sahara was not a desert during the African humid period. Instead, most of northern Africa was covered by grass, trees, and lakes.

The African humid period (AHP) is a climate period in Africa during the Holocene during which northern Africa was wetter than today. It was caused by changes in Earth's orbit around the Sun, and involved changes in vegetation and dust in the Sahara that altered the African monsoon, the disappearance of much of the Sahara desert which was replaced by grassy vegetation, trees and lakes and the settlement of the former desert by various animals and humans, who lived as hunter-gatherers. It has had profound effects on present-day Africa such as the birth of the Pharaonic civilization and the pyramids and potentially also the development of widespread Golden Age myths.

Before the African humid period, during the last glacial maximum, the Sahara was much larger than today and had extensive dune fields; many lakes and rivers such as Lake Victoria and the White Nile were either dry or at low levels and the Sahara mostly uninhabited. The African humid period commenced about 14,600–14,500 years ago at the end of Heinrich event 1 and concomitant to the Bølling-Allerød warming; rivers and lakes such as Lake Chad formed or expanded, glaciers grew on Mount Kilimanjaro and the Sahara retreated. Two major fluctuations occurred, one during the Younger Dryas and the other during the 8.2 kiloyear event during both of which temporarily drier conditions returned across Africa. The end of the African humid period came about 6,000–5,000 years ago during the Piora Oscillation cold period when the Sahara occupied its present position. While some evidence points to an end 5,500 years ago in the Sahara, in the Sahel, Arabia and East Africa the end of humidity appears to have taken place in several steps such as the 4.2 kiloyear event.

Earlier than the African humid period, humid periods in Africa had influenced the evolution of modern humans; the African humid period now led to a widespread settlement of the Sahara and the Arabian Deserts by humans. These at first lived on animals and plants naturally occurring in the region; later they started domesticating animals such as cattle, goats and sheep. They have left archeological sites and artifacts such as one of the oldest ships in the world; but in particular they created rock paintings such as those in the Cave of Swimmers and in the Acacus Mountains; in fact the existence of earlier wet periods was postulated after the discovery of these rock paintings in now-inhospitable parts of the Sahara. When the African humid period ended, humans gradually abandoned the desert in favour of regions with more secure water supplies, such as the Nile Valley and Mesopotamia, where they gave rise to early complex societies.

The African humid period was part of a phase where monsoon activities were stronger across the Northern Hemisphere, from the Mojave Desert in North America over Africa and the Middle East to India and China. The ultimate reason was the precession of Earth's orbit around the Sun, which shifted the season during which Earth is closest to the Sun towards Northern Hemisphere summer, increasing summer insolation and the strength of the monsoons that depend on it. This alone was not enough to make the Sahara disappear; other processes, among others the ability of vegetation and large lakes in the desert to reduce the emission of reflecting dust and to increase the amount of sunlight absorbed by the ground, did play a role in its onset. Decreased summer insolation as the Holocene progressed also brought the African humid period to an end. Increased greenhouse gases during the Holocene appear to have aided in the onset of the African humid period; this may imply that anthropogenic global warming will also result in a shrinkage of the Sahara desert.

Research history

Herodotus in 440 BC and Strabon in 23 AD discussed the existence of a greener Sahara, although their reports were at first questioned owing to their anecdotal nature. In 1850 the researcher Heinrich Barth discussed the possibility of past climate change leading to increased wetness in the Sahara after discovering petroglyphs in the Murzuq Desert, and further discoveries of petroglyphs led desert explorer László Almásy to coin the concept of a Green Sahara in the 1930s. Later in the 20th century, conclusive evidence of a past greener Sahara and the existence of lakes was increasingly reported.[1][2]

The idea that changes in Earth's orbit around the Sun influence the strength of the monsoons was already advanced in 1921, and while the original description was partly inaccurate later widespread evidence for such orbital controls on climate was found.[1] At first it was believed that humid periods in Africa correlate with glacial stages ("pluvial hypothesis") before radiocarbon dating became widespread.[3]

The development and existence of the African humid period has been investigated with archeology, climate modelling and paleoproxies,[4] with archeological sites,[5] dunes as well as lake, marine and wetland deposits playing an important role.[2] Pollen, lake deposits and former levels of lakes have been used to study the ecosystems of the African humid period,[6] and charcoal and leaf impressions have been used to identify vegetation changes.[7]

Research issues

While the precipitation changes since the last glacial cycle are well established, the magnitude and timing of the changes are often unclear.[8] The amounts of precipitation reconstructed from paleoclimate records and simulated by climate modelling are often inconsistent with each other[9] and the former between themselves;[10] the simulation of the Green Sahara in general is considered a problem for earth system models.[11] Erosion of lake sediments and reservoir effects make it difficult to date the end of their existence.[12] Vegetation changes by themselves do not necessarily indicate precipitation changes, as changes in seasonality, plant species composition and changes in land use also influence the vegetation records.[13] Isotope ratios such as the hydrogen/deuterium ratio that have been used to reconstruct past precipitation values likewise are under the influence of various physical effects, which complicates their interpretation.[14]

Terminology

Earlier humid periods are sometimes also known as "African humid periods"[15] and a number of dry/wet periods have been defined for the Central Africa region.[16] In general, such climate fluctuations between wetter and drier periods are known as "pluvials" and "interpluvials", respectively.[17] Other terms that have been applied to the African humid period or correlative climate phases specifically are:

  • "Holocene humid period", which also covers an analogous episode in Arabia and Asia.[18]
  • "Holocene Pluvial".[19]
  • "Holocene Wet Phase".[20]
  • In Central Africa as "Kibangien A".[21]
  • "Makalian" in the Neolithic of northern Sudan.[22]
  • The 14,000–6,000 humid period over the Eastern Mediterranean and Levant is also known as the "Nabtian period"[23] or the "Nabtian Wet Phase".[24]
  • "Neolithic pluvial".[25]
  • "Neolithic Subpluvial".[20]
  • In the Western Sahara 6,500 - 4,000 years before present "Nouakchottien".[26]
  • In the Central Sahara 14,000 - 7,500 years before present "Tchadien".[26]

The terms "Léopoldvillien"[27] and Ogolien [fr] have been applied to the dry period in the last glacial maximum,[28] the latter is equivalent to the "Kanemian";[29] "Kanemian dry period" refers to a dry period between 20,000–13,000 years before present in the Lake Chad area.[30]

Background and beginning

The African humid period took place in the late Pleistocene[31] and early-middle Holocene[32][33] and was characterized by increased precipitation in Northern and Western Africa[13] relative to today[34] due to a northward migration of the tropical rainbelt.[35] Within the otherwise climatically relatively stable Holocene the African humid period is considered to be a major fluctuation.[36] It is also part of the so-called Holocene climatic optimum which featured warmer summers than present-day in the Northern Hemisphere.[37] Sometimes the humid period is subdivided into a "AHP I" that lasted until 8,000 years ago and a "AHP II" from 8,000 years onwards until the end of the African humid period,[38] the former being wetter than the latter.[39]

The African humid period was not the first such phase; evidence for about 230 older such "green Sahara"/wet periods exist going back perhaps to the first appearance of the Sahara 7-8 million years ago;[1] such wet periods took place during Marine Isotope Stage 5 a and c for example.[40] Such humid periods are usually associated with interglacials while glacial stages correlated to dry periods.[15] Earlier humid periods appear to have been more intense than the African humid period of the Holocene,[41][42] including the exceptionally intense Eemian humid period which provided the pathways for early humans to cross Arabia and Northern Africa[43] and which together with later moist periods has been linked to expansions of the Aterian populations.[44]

The Bølling-Allerød warming appears to be synchronous to the onset of the African humid period[45][46][47] as well as to increased humidity in Arabia;[48] later the humid period coincides with the Atlantic period of the Blytt–Sernander sequence.[49]

Conditions before the African humid period

African vegetation during the last glacial maximum

During the Last Glacial Maximum before the African humid period, the Sahara and Sahel had been extremely dry[50] with less precipitation occurring than today[34] as reflected by the extent of dune sheets and water levels in closed lakes.[50] The southern margin of the Sahara shifted southwards by about 500–800 kilometres (310–500 mi)[51] or 5° latitude with respect to its present-day border,[52] that is a much larger desert than today.[53] Dunes were active much closer to the equator[51] in sub-Saharan Africa[54] and also over Israel and Arabia[55] and the exposed Persian Gulf[56] where dust generation increased.[48] Closer to the equator, rainforests retreated to isolated areas and were elsewhere replaced by afromontane and savannah vegetation as temperatures and rainfall decreased.[27] This drying tendency appears to be the consequence of lower sea surface temperatures and lower atmospheric water content.[57]

There is little and often equivocal evidence of human activity in the Sahara and also in Arabia at that time, reflecting its drier nature.[58][59] Episodes of subtropical drought between 11,500 and 21,000 years before present coincided with the discharge of large amounts of icebergs in the North Atlantic,[60] and exceptional dry phases are linked to Heinrich events.[61] The aridity during the Last Glacial Maximum appears to be the consequence of the colder climate and larger polar ice sheets, which squeezed the monsoon belt to the equator and weakened the West African Monsoon. The atmospheric water cycle and the Walker and Hadley circulations were weaker as well.[62]

Before the onset of the African humid period Lake Victoria, Lake Albert and Lake Edward were no longer overflowing into the White Nile;[63] Lake Turkana appears to have been partially dry.[64] The White Nile had become a seasonal river[65] whose course[66] along with that of the main Nile may have been dammed by dunes.[67] The Sudd swamps also had dried out.[65] The Nile Delta was partially dry, with sandy plains extending between ephemeral channels and exposed seafloor, and it became a source of sand for ergs[a] farther east.[69] Other lakes across Africa such as Lake Chad and Lake Tanganyika also had shrunk during the Last Glacial maximum,[70] although some lakes persisted in areas where colder temperatures had decreased evaporation.[29]

Whether some parts of the desert such as highlands like the Red Sea Hills were reached by the westerlies[71] or weather systems associated with the subtropical jet stream[72] and thus received precipitation is contentious and only clearly supported for the Maghreb in northwestern Africa,[71] river flow[54]/terrace formation[73] and lake development in the Tibesti and Jebel Marra mountains[74][75] as well as residual Nile flow may be explained in this way.[76] The highlands of Africa appear to have been less affected by drought during the last glacial maximum.[77] Finally, glaciers were active in the Bale and Semien Mountains of Ethiopia during the last glacial maximum.[51]

Onset

The humid period began about 14,600[78]-14,500 years ago,[b][31] alternatively it may have begun 15,000 years ago.[80][81] In general, the end of the glacial drought occurred between 17,000 and 11,000 years ago,[75] with an earlier beginning noted in the Saharan mountains[82] where it may have begun 18,500 years ago.[83] In southern and central Africa earlier starts 17,000 and 17,500 years ago, respectively, are possibly linked to Antarctic warming,[80][21] while Lake Malawi appears to have been low until about 10,000 years ago.[84]

High lake levels occurred in the Jebel Marra and Tibesti Mountains between 15,000 and 14,000 years ago[85] and the youngest stage of glaciation in the High Atlas mountains took place at the same time as the early African humid period.[86] 14,500 years ago lakes started to appear in the arid areas, by 14,500 years before present Lake Victoria which had dried up during the preceding arid period filled again.[87] The onset of the humid period took place almost simultaneously over all of Northern and Tropical Africa,[88] while in Arabia wet conditions apparently took about two millennia to advance northward[89] with tephrochronology evidence indicating a gradual advance of humidity in northern Arabia and the Middle East.[90] Some lake level curves indicate a stepwise increase of lake levels 15,000 ± 500 and 11,500–10,800 years ago, before and after the Younger Dryas.[91]

The reappearance of Lake Victoria[87] was also accompanied by its overflow and that of Lake Albert 15,000–14,500 years ago[63] into the White Nile; Lake Tana began to overflow as well, into the Blue Nile.[85] The White Nile flooded part of its valley,[92] and it became reconnected to the main Nile.[81][c] In Egypt widespread flooding took place as part of the "Wild Nile" period[85] which led to the largest recorded floods on this river,[67] sedimentation in floodplains[94] and probably also impacted human populations along the river.[95] Even earlier, 17,000–16,800 years ago, meltwater from glaciers in Ethiopia – which were retreating at that time – may have begun to increase the flow of water and sediment in the Nile.[96] In the East African Rift water levels in lakes began to rise by about 15,500/15,000[97]-12,000 years ago;[98] Lake Kivu began overflowing into Lake Tanganyika by about 10,500 years ago.[99]

About the same time that the African humid period started, the cold glacial climate in Europe associated with Heinrich event 1 ended[87] with climate changing as far as Australasia.[85] In the Sahara mountains, the onset of a more humid climate appears to have preceded deglaciation in Europe.[57] A warming and retreat of sea ice around Antarctica may also coincide with the start of the African humid period,[100] although the Antarctic Cold Reversal also falls into this time.[21]

Causes

The African humid period was caused by a stronger West African Monsoon[101] directed by insolation changes and changes in albedo feedbacks,[9] leading to increased moisture import from the equatorial Atlantic to West Africa, but also from the North Atlantic and the Mediterranean Sea towards Mediterranean Africa.[102][103] There also were complex interactions with the atmospheric circulation of the extratropics and between moisture coming from the Atlantic Ocean and the Indian Ocean,[104] and an increased overlap between the areas wetted by the monsoon and by extratropical cyclones.[105]

Climate models indicate that changes from a dry to a green Sahara and back have threshold behaviour, with the flip occurring once a certain level of insolation is exceeded;[106] likewise, a gradual drop of insolation often leads to a sudden transition back to a dry Sahara.[107] This is due to various feedback processes which are at work[13] and in climate models there is often more than one stable climate-vegetation state.[108] As an example, sea surface temperature and greenhouse gas changes synchronized the beginning of the African humid period across Africa.[88]

Orbital changes

Milankovich cycles over the past one million years

The African humid period has been explained by increased insolation during Northern Hemisphere summer.[13] Due to precession, the season at which Earth passes closest to the Sun on its elliptical orbit – the perihelion – changes, with maximum summer insolation occurring when this happens during Northern Hemisphere summer.[109] Between 11,000 and 10,000 years ago, Earth passed through the perihelion at the time of summer solstice increasing the amount of solar radiation by about 8%,[31] resulting in the African monsoon becoming both stronger and reaching farther north.[110] The obliquity also decreased during the Holocene[111] but the effect of obliquity changes on the climate is focused on the high latitudes and its influence on the monsoon is unclear.[112]

During summer, solar heating is stronger over the North African land than over the ocean, forming a low pressure area that draws moist air and precipitation in during the summer months[31] from the Atlantic Ocean.[113] This effect was strengthened by the increased summer insolation[114] thus leading to a stronger monsoon that also reached farther north.[111] The strength of this circulation and resulting precipitation can change strongly in response to changes in summer insolation, with effects as far as the subtropics.[12]

Obliquity and precession are responsible for two of the foremost Milankovich cycles and are responsible not only for the onset and cessation of ice ages[115] but also for monsoon strength variations.[112] Southern Hemisphere monsoons are expected to have the opposite response of Northern Hemisphere monsoons to precession, seeing as the insolation changes are reversed; this observation is borne out by data from South America.[116] The precession change increased seasonality in the Northern Hemisphere while decreasing it in the Southern Hemisphere.[111]

Albedo feedbacks

According to climate modelling,[1] orbital changes by themselves cannot increase precipitation over Africa enough to explain the formation of the large desert lakes such as 330,000 square kilometres (130,000 sq mi) Lake Megachad[12] – an expanded Lake Chad[117] which had a size comparable to the Caspian Sea[118] – or the northward expansion of vegetation[119][111] unless ocean and land surface changes are factored in.[13]

Decreasing albedo resulting from vegetation changes is an important factor in the precipitation increase.[12] Specifically, increased precipitation increases the amount of vegetation; vegetation absorbs more sunlight and thus more energy is available for the monsoon. In addition, evapotranspiration from vegetation adds more moisture, although this effect is less pronounced than the albedo one.[50] Heat fluxes in the soil and evaporation are also altered by the vegetation.[120]

In addition to raw precipitation changes, changes in precipitation seasonality such as the length of dry seasons need to be considered when assessing the effects of climate change on vegetation,[121] as well as the fertilizing effects of increased carbon dioxide concentrations in the atmosphere.[120]

Other sources of albedo changes:

  • Changes in soil properties also result in changes in the monsoon; replacing desert soils with loamy ones results in increased precipitation,[122] and soils that are wet[120] or contain organic matter reflect less sunlight and accelerate the moistening process.[1] Wet soil[123] and desert sand changes also modify the albedo.[120]
  • Albedo changes caused by lakes and wetlands[9] in modelling can alter precipitation.[122]
  • Reduced dust generation from a wetter Sahara influences the climate[124] by reducing the amount of light absorbed by dust and also modifying cloud properties, making them less reflective and more efficient at inducing precipitation.[1][125] There is not universal agreement on the effects of dust on precipitation in the Sahel, however.[1]

Intertropical Convergence Zone changes

Sea surface temperatures off North Africa warmed under orbital effects and also due to weaker trade winds and direct a northward movement of the Intertropical Convergence Zone and increased moisture gradients between land and sea.[126][50] Temperature gradients between a cooler Atlantic in springs and the African continent and between warmer temperatures north of 10° latitude and cooler south of it may have assisted in this change.[127] In Eastern Africa on the other hand Intertropical Convergence Zone changes had relatively little effect on precipitation changes.[128][129] The position of the Intertropical Convergence Zone is also contentious in Arabia.[130]

Precipitation changes in East Africa

The African humid period that took place in East Africa on the other hand appears to have been caused by different mechanisms,[131] namely decreased seasonality of precipitation,[132] increased precipitation or shorter dry seasons[133] and increased inflow of moisture from the Atlantic and Indian Oceans. The Atlantic moisture inflow was in part triggered by a stronger West African and Indian monsoon, perhaps explaining why the effects of the African humid period extended into the Southern Hemisphere.[128][134] The effect of the easterly trade winds is unclear; increased moisture transport by easterly trade winds may have aided in the development of the African humid period[101] but alternatively a stronger Indian Monsoon that draws easterly winds away from East Africa may have occurred.[135]

Changes in the Congo Air Boundary[d][136] or increased convergence along this boundary may be involved here;[133][136] the Congo Air Boundary would have been shifted east by the stronger westerly winds[134] that are directed by lower atmospheric pressure over Northern Africa[137] and thus allow additional moisture from the Atlantic to reach East Africa.[138]

Different causes for increased humidity in East Africa might have dominated in the early and late African humid period,[139] and the concept of an "African humid period" extending into this part of Africa has raised criticism.[140] Finally, increased greenhouse gas concentrations may have been involved in directing the onset of the African humid period in tropical southeastern Africa;[141] there, orbital changes would be expected to lead to climate variations opposite to these in the Northern Hemisphere[142] but the sparse evidence of the former climate does not entirely match this theory.[143]

Additional factors

  • Climate change in the far northern latitudes may have contributed to the onset of the African humid period,[101] such as the shrinkage of the Scandinavian Ice Sheet and the Laurentide Ice Sheet.[120] A retreat of the ice sheets are often required to simulate the onset of the African humid period.[144] Their existence might also explain why the African humid period did not start immediately with the early insolation peak, as still existing ice sheets would have cooled the climate.[145]
  • Sea surface temperature changes in the Atlantic[101] including warmer sea surface temperatures caused by insolation changes[106] and weaker trade winds[126] also increased precipitation,[12] in part by increasing moisture gradients between land and sea[126] and landward moisture transport.[113]
  • Warming of the Mediterranean Sea increases the amount of Sahel precipitation; this effect is responsible for the recent anthropogenic global warming mediated increase in Sahel precipitation.[1] Warmer sea surface temperatures there might also explain the increased precipitation recorded in the Mediterranean during the African humid period.[130]
  • Increased precipitation during winter is correlated with a larger extent of Mediterranean precipitation and might have aided in the establishment of the African humid period, especially in North Africa,[146][147][148] around the northern Red Sea,[149] in the Tibesti[150] and in northern Arabia[130] and generally at higher latitudes where the monsoon did not arrive.[127] It has also been proposed for the Tibesti region.[151] This precipitation may have extended to other parts of the Sahara; such would have led to the areas of summer and winter precipitation overlapping.[152] Such changes in Mediterranean-derived precipitation may correlate with changes in the North Atlantic and Arctic Oscillations.[146]
  • Trough-mediated northward transport of moisture during autumn and spring has also been proposed to explain the increased precipitation and its underestimation by climate models.[9] In one climate model, increased northward moisture transport by such troughs increases autumn rainfall in the Sahara, especially in the mid-Holocene and when the climate is already moister than usual there.[153]
  • Weaker subtropical anticyclones were proposed as an explanation during the 1970s-1980s.[154]
  • In montane regions such as the Meidob volcanic field cold temperatures after the last glacial maximum may have reduced evaporation and thus allowed an early onset of humidity.[155]
  • Changes in the Earth's geomagnetic field may be linked to the humidity changes.[156]
  • Increased moisture supply from larger lakes like Lake Megachad may have increased the precipitation, although this effect is probably not adequate to explain the entire African humid period wettening.[123] A similar role has been attributed to the extensive wetlands, drainages and lakes in the Eastern Sahara.[157]
  • Two high elevation winds, the African Easterly Jet and the Tropical Easterly Jet modulate atmospheric air flows over Africa and thus also the amount of precipitation; the Tropical Easterly Jet comes from India and is powered by temperature gradients between the tropics[34] and the subtropics while the African Easterly Jet is powered by temperature gradients in the Sahel.[158] A stronger West African Monsoon resulted in a weaker African Easterly Jet and thus decreased transport of moisture out of Africa.[134]
  • Increased latent heat fluxes play a role in explaining the moistening in coupled climate models.[125]
  • Increased atmospheric carbon dioxide concentrations may have played a role in triggering the African humid period,[120] especially its extension across the equator,[159] as well as its resumption after the Younger Dryas and Heinrich event 1 through increased sea surface temperatures.[160]
  • In some parts of the Sahara increased water supply from montane regions may have assisted in the development of moist conditions.[161][162]
  • Larger forests in Eurasia may have led to a northward shift of the ITCZ.[163]
  • Other proposed mechanisms involve convection occurring above the atmospheric boundary layer,[164] changes in the solar cycles[165] and complex atmospheric flow phenomena.[166]

Manifestations

Vegetation map 7,000 years ago

The African humid period extended over the Sahara as well as eastern,[167] southeastern and equatorial Africa. In general, forests and woodlands expanded through the continent.[32] A similar wet episode took place in the tropical Americas, China, Asia,[e] the Middle East and the Arabian Peninsula[168][169][35][50] and appears to relate to the same orbital forcing as the African humid period.[168] An early Holocene monsoonal episode extended as far as the Mojave Desert in North America.[170] Conversely, a drier episode is recorded from much of South America where Lake Titicaca, Lake Junin, the discharge of the Amazon River and water availability in the Atacama were lower.[171]

The discharge of the Sanaga River[172] and other rivers[173] in Cameroon,[174] Niger River, Congo River,[172] Nile River,[175] Ntem River[19] and Rufiji River increased[176] and runoff from equatorial Africa, northeastern Africa and the western Sahara was also larger.[177] Changes in the morphology of the river systems and their alluvial plains occurred in response to the increased discharge.[21][19]

Flora and fauna of the Sahara

During the African humid period, lakes, rivers, wetlands and vegetation including grass and trees covered the Sahara and Sahel[114][178][110] and into the Red Sea Hills,[179] creating a "Green Sahara".[180] Evidence includes pollen data, archeological sites, evidence of faunal activity such as diatoms, mammals, ostracods, reptiles and snails, buried river valleys, organic-rich mats, mudstones, evaporites as well as travertines and tufas deposited in subaqueous environments.[33]

A present-day savannah

The vegetation cover then extended over almost all of the Sahara[31] and consisted of an open grass savannah with shrubs and trees.[113][181] In general, the vegetation expanded northward[35] to 27-30° northern latitude in West Africa[182][7] with a Sahel boundary at about 23° north,[37] as the Sahara was populated by plants that today often occur about 400 kilometres (250 mi)[183][184]-600 kilometres (370 mi) farther south.[185] The northward movement of vegetation took some time and some plant species moved faster than others.[186]

Forests[187] and plants from the humid tropics were concentrated around lakes and rivers.[188] The landscape during the African humid period has been described as a mosaic between various vegetation types of semi-desert and humid origin[189] rather than a simple northward displacement of plant species,[190] and some brown or yellow vegetation communities persisted.[1] Pollen data often show a dominance of grasses over humid tropics trees.[7]

Nevertheless, the climate of the Sahara did not become entirely homogeneous; the central-eastern parts of the desert were probably drier than the western and central sectors[191] and the Libyan sand sea was still a desert[1] although pure desert areas retreated or became arid/semiarid.[192] An arid belt may have existed north of 22° latitude;[193] in general conditions between 21° and 28° northern latitude are poorly known.[194] Dry areas may have persisted in the rain shadows of mountains and could have supported arid climate vegetation, explaining the presence of its pollen in sediment cores.[195] In addition, north-south gradations in vegetation patterns have been reconstructed from charcoal and pollen data.[196]

Fossils record changes in the animal fauna of the Sahara.[197] This fauna included antelopes,[31] catfish,[198] clams,[199] crocodiles, elephants, giraffes,[31] hartebeest, hippos,[198][200] molluscs, Nile perchs,[201] tilapia,[199] turtles[198] and many more animals,[202] and in Egypt spotted hyenas, warthogs, water buffaloes, wildebeest and zebra occurred.[203] Some animals expanded over the whole desert, while others were limited to places with deep water.[201] Earlier humid periods in the Sahara may have allowed species to cross the now-desert.[193] A reduction in open grasslands at the beginning of the African humid period may explain a population bottleneck in cheetahs at the start of the humid period,[204] while the humid period led to the expansion of some animal populations such as Hubert's multimammate mouse.[205]

Lakes and rivers of the Sahara

Lake Megachad, with present-day Lake Chad highlighted in green

A number of lakes formed[197] or expanded in the Sahara[154] including both mosaic-like lakes and large lakes; the largest of which was Lake Chad[206]/Lake Megachad[117] which increased to at least ten times its present-day size[207] and overflowed into the Niger River[208] during highstand through the Mayo Kebbi and the Benue River, eventually reaching the Gulf of Guinea.[209] This enlarged Lake Chad reached dimensions of 1,000 by 600 kilometres (620 mi × 370 mi) in north-south and east-west direction respectively,[187] covering the Bodélé Depression[210] and perhaps as much as 8% of the present-day Sahara desert.[211] It influenced the climate itself;[212] for example rainfall would be reduced at the centre of the lake and increased at its margins.[1] Lake Chad was possibly fed from the north by rivers draining the Hoggar (Taffassasset drainage)[213] and Tibesti Mountains and from the south by the Chari River-Logone and Komadugu;[209] the rivers draining the Tibesti formed alluvial fans[214]/the Angamma river delta at their entry into the Lake Chad basin.[215] Skeletons of elephants, hippos and hominins have been found in the Angamma river delta on the northern side of then-Lake Megachad, the dominant shoreline feature on its northern side.[187]

Among the lakes which may have formed in the Sahara are Lake Megafezzan in Libya[216] and Lake Ptolemy in Sudan.[126][211][206] In 2018, some doubts have been raised about the size and existence of some of these lakes, however,[217] especially for Lake Megafezzan.[218] Other lakes are known from I-n-Atei in the Hoggar, at Ine Sakane[219] and in Taoudenni[f] in Mali,[221] Chemchane in Mauretania,[222] at Sebkha Mellala close to Ouargla in Algeria,[223] at Bilma, Dibella, Fachi[224] and Gobero in the Ténéré[6] and at "Eight Ridges",[225] El Atrun,[226] "El Gureinat", "Ridge",[225] Selima and Oyo in Sudan.[227]

In some parts of the Sahara ephemeral lakes formed such as at Bir Kiseiba and Nabta Playa, both in Egypt and both featuring archeological sites[228] which may relate to later Egyptian religions,[229] or swamp-lakes such as at Adrar Bous close to the Air Mountains.[224] In addition, dune-contained ephemeral lakes formed in some areas of the Sahara desert,[230] and a "freshwater archipelago" appears to have existed in the Murzuq basin.[231] Finally, crater lakes formed in volcanic fields[232] and sometimes survive to this day as smaller remnant lakes such as Malha crater[233] in the Meidob volcanic field.[232] Potentially, the increased availability of water during the African humid period may have facilitated the onset of phreatomagmatic eruptions such as maar formation in the Bayuda volcanic field, although the chronology of volcanic eruptions there is not well known enough to substantiate a link to the African humid period.[234]

In addition to lakes, rivers such as the Irharhar in Algeria, Libya and Tunisia[235] and the Sahabi and Kufra rivers in Libya were active during this time[236] although there is some doubt that they carried perennial water;[237] they appear to have been more important in earlier humid periods.[238] Flow also took place in wadis[239] and in rivers discharging into endorheic basins such as Wadi Tanezzuft.[240] In the Air, Hoggar and Tibesti Mountains, the so-called "Middle Terrace" was emplaced at this time.[241] The rivers of the Sahara,[236] lakes and their watersheds may have acted as pathways for the spread of humans and animals;[242] the rivers were often connected between each other through alluvial fans.[236] Proposed examples of animals that spread through rivers are the Nile crocodile and the fish Clarias gariepinus and Tilapia zillii.[195] Some rivers discharging through the Bay of Arguin in Mauretania formed estuaries and mangroves there,[243] such as the large Tamanrasset River[244] which once formed the 12th largest watershed in the world[245] and has left a submarine canyon and riverine sediments.[238] Other rivers in the same area also formed submarine canyons,[246] and submarine landslides in the area have been related to the activity of these rivers.[247]

Humans of the Sahara

Conditions and resources were ripe for first hunter-gatherers, fishermen[248] and, later, pastoralists,[249] whose arrival in the Sahara coincided with a time of developing lakes[250] and may have occurred as a migration from either the north (Maghreb or Cyrenaica)[251] where the Capsian culture was located,[252] the south (Sub-Saharan Africa), or the east (Nile Valley).[251] Traces of human activity have been found in the Acacus Mountains[253] where caves and rock shelters were used as basecamps for humans,[254] such as the Uan Afuda cave[253] and the Uan Tabu and Takarkori rock shelters.[255] In Takarkori the first occupation took place in the 9th and 10th millennium before present;[256] about five millennia of human cultural evolution are recorded there.[249] At Gobero in the Ténéré desert a cemetery has been found, which has been used to reconstruct the lifestyle of these former inhabitants of the Sahara,[6] and at Lake Ptolemy in Nubia humans settled close to the lake shore, using its resources and perhaps even engaging in leisure activities.[257] Many humans at that time appear to have depended on water-bound resources, seeing as many of the tools left by the early humans are associated with fishery; hence this culture is also known as "aqualithic"[154][178] although substantial differences between the cultures of various places have been found.[258] The greening of the Sahara led to a demographic expansion[259] and especially in the Eastern Sahara human occupancy coincides with the African humid period.[260]

Humans were hunting large animals with weapons that have been found in archeological sites[261] and wild cereals occurring in the Sahara during the African humid period such as brachiaria, sorghum and urochlea were an additional source of food.[262] Humans also domesticated cattle[49] especially in the more environmentally variable Eastern Sahara,[263] goats and sheep.[264] Animal husbandry picked up in earnest around 7,000 years ago when domestic animals came to the Sahara, and a population boom may be linked to this change in cultural practice.[248] Dairying has been demonstrated in some locations[265] and cattle-husbandry is supported by the frequent depiction of cattle in rock paintings.[266] The Dufuna canoe, one of the oldest known ships in the world,[267] appears to date to the Holocene humid period and implies that the waterbodies of that time were navigated by humans.[268] In the Acacus Mountains, several cultural horizons known as Early and Late Acacus and Early, Middle, Late and Final Pastoral have been identified[269] while in Niger the Kiffian culture has been related to the beginning of the African humid period.[270] Ancient civilizations thrived,[35] with farming and animal husbandry taking place in Neolithic settlements.[222][271] Possibly, the domestication of plants in Africa was delayed by the increased food availability during the African humid period, only taking place around 2,500 BC.[272]

Images of swimming people in the Cave of Swimmers

Humans created rock art[273] such as a number of petroglyphs and rock paintings in the Sahara, perhaps the largest density of such creations in the world.[274] Scenes include animals[110] and everyday life[274] such as swimming which supports the presence of past wetter climates.[234] One well-known such petroglyph location is the Cave of Swimmers in the Gilf Kebir mountains of Egypt;[275] other well known sites are the Gabal El Uweinat mountains also of Egypt,[49] Arabia[276] and the Tassili n'Ajjer in Algeria where a number of rock paintings from this time have been discovered.[277] Humans also left artifacts such as Fesselsteine[g] and ceramics in what today are inhospitable deserts;[49] North Africa together with East Asia is one of the first places where pottery was developed[249] probably under the influence of increased availability of resources during the African humid period. The humid period also favoured its development and spread in West Africa during the 10th millennium BC;[279] the so-called "wavy line" or "dotted wavy-line" motif was widespread across Northern Africa.[258]

These populations have been described as Epipaleolithic, Mesolithic[280] or Neolithic[281] and produced a variety of lithic tools and other assemblages.[282] Genetic and archeological data indicate that these populations which exploited the resources of the African humid period Sahara probably originated in Sub-Saharan Africa and moved north after some time, after the desert got wetter;[283] this may be reflected in the northward spread of Macrohaplogroup L and Haplogroup U6 genomic lineages.[284] In return, the African humid period facilitated the movement of some Eurasian populations into Africa.[285] These favourable conditions for human populations may be reflected in paradise myths such as the Garden of Eden in The Bible and Elysium and the Golden Age in Classical Antiquity,[286] and in the spread of the Nilo-Saharan languages.[195][258]

Additional manifestations in the Sahara

The expanded vegetation stabilized previously active dunes, eventually giving rise to the present-day draa dunes in the Great Sand Sea of Egypt for example.[230] Soil development and biological activity in soils are attested in the Acacus Mountains[287] and the Mesak Settafet area of Libya,[288] but evidence of soil formation[289]/pedogenesis[41] such as the development of bog iron[290] are described from other parts of the Sahara as well.[41] The Central and Southern Sahara saw the development of alluvial deposits.[154] In Northern Africa and in the Sinai Peninsula, deposition of tufa and carbonate sediments within groundwater-fed lakes have been associated with the African humid period.[94] Finally, lightning strikes into soil left altered rocks in parts of the Central Sahara.[291]

The increased precipitation also resulted in aquifer recharge[292][280] such as the Nubian Sandstone Aquifer; presently, water from this aquifer maintains several lakes in the Sahara, such as the Lakes of Ounianga.[293] Other groundwater systems were active at that time in the Acacus Mountains, Air Mountains, in the Fezzan[294] and elsewhere in Libya.[295] Raised groundwater tables provided water to plants and was discharged in depressions.[296]

The formation of lakes[55] and vegetation reduced the export of dust from the Sahara, which is reflected in marine cores[297][124] including one core where dust export decrease dby almost half.[298] In coastal places, such as in Oman, sea level rise also reduced the production of dust.[55] In the Mediterranean, a decreased dust supply was accompanied by increased sediment input from the Nile, leading to changes in marine sediment composition.[299]

Whether the strengthening of the monsoon enhanced or reduced upwelling off Northwestern Africa is debatable,[300] with some research suggesting that the strengthening in upwelling decreased sea surface temperatures[301][302][303] and increased the biological productivity of the sea,[300] while other research suggests that the opposite occurred; less upwelling with more moisture.[50] However, regardless of whether upwelling increased or decreased, it is possible that either way, the strengthening of the monsoon still provided a boost of productivity to the coasts of Northern Africa because the increased river discharge delivered more nutrients to the sea.[301][302][303]

Arabia

Precipitation in Dhofar and southwestern Arabia is brought by the African monsoon,[304] and a change to a wetter climate resembling Africa has been noted in southern Arabia[305] and Socotra from cave and river deposits.[306] Paleolakes are recorded at Tayma, Jubbah,[307] in the Wahiba Sands of Oman[308] and Mundafan for the Holocene, and the Wadi ad-Dawasir river system became active again[309][310] with increased river runoff into the Persian Gulf.[311] In the Rub al-Khali lakes formed between 9,000 and 7,000 years ago[312] and dunes were stabilized by vegetation,[89] although the formation of lakes there was less pronounced than in the Pleistocene.[313] Episodes of increased river discharge occurred in Yemen[314] and increased precipitation is recorded in the caves of Hoti, Qunf in Oman, Mukalla in Yemen and Hoq cave in Socotra.[315] Freshwater sources in Arabia during the African humid period became focus points of human activity[316] and herding activity between mountains and lowlands occurred.[89] In addition, karstic activity took place on exposed coral reefs in the Red Sea and has left traces to this day.[317] Increased precipitation has been also invoked to explain decreased salinities in the Red Sea.[318]

However, in Arabia, deserts did not retreat as much[169] and precipitation may not have reached the central[319] and northern part of the peninsula;[320] northern Arabia remained somewhat drier than southern Arabia.[321] One study has estimated that the amount of rainfall in the Red Sea did increase to no more than 1 metre per year (39 in/year).[322] Some former lakes in Arabia have been interpreted to be marshes but these interpretations have been contested.[323]

East Africa

Nile discharge was higher than today[175] and during the early African humid period, the Nile in Egypt flooded up to 3–5 metres (9.8–16.4 ft)[175] higher than it did recently before flood control;[85] the increased flooding may explain why many archeological sites along the Nile were abandoned.[66][95] Waters from the Nile filled depressions like the Fayum Depression.[240] In addition, Nile tributaries in northwestern Sudan[324] such as Wadi Al-Malik,[175] Wadi Howar[h][326] and Valley of the Queens became active during the African humid period.[327] Wadi Howar was active until 4,500 years ago,[326] and at the time often contained dune-dammed lakes, swamps and wetlands;[328][162] it was the largest Saharan tributary of the Nile[329] and constituted an important pathway into sub-Saharian Africa.[175] Conversely it appears that Lake Victoria and Lake Albert were not overflowing into the White Nile for all of the African humid period,[330] and the White Nile would have then been sustained by overflow from Lake Turkana instead.[326] There appears to be a tendency over the course of the African humid period for the discharge of the Blue Nile to decrease relative to that of the White Nile.[331] The Blue Nile built an alluvial fan at its confluence with the White Nile, and incision by the Nile reduced flooding risk in some areas which thus became available for human use.[175]

Some lakes formed or expanded during the African humid period

Closed lakes in East Africa rose, sometimes by hundreds of metres.[281] Lake Suguta developed in the Suguta Valley, accompanied by the formation of river deltas where rivers such as the Baragoi River entered the lake,[332] and overflowed into the Kerio River, this adding water to Lake Turkana[333] where increased discharge by the Turkwel River led to the formation of a large river delta.[334] In turn, Lake Turkana overflowed through the Lotikipi Swamp into the White Nile.[335] This overflowing large lake was filled with freshwater and was populated by societies that engaged in fishery[336] but probably could also fall back on other resources in the region.[337] Deposits from this lake highstand form the Galana Boi Formation.[258] Other lakes that expanded include Lake Shala in Ethiopia[338] which joined with some neighbouring lakes and began overflowing into the Awash River,[339] Lake Hayq also in Ethiopia,[340] and Lake Bogoria, Lake Naivasha,[154] Lake Nakuru/Lake Elmenteita all in Kenya.[341] A 1,600 square kilometres (620 sq mi) large and 50 metres (160 ft) deep Lake Magadi formed in the early Holocene,[118] and in the Danakil Depression of Ethiopia freshwater conditions became established.[154] Finally, lakes formed in depressions on the mountains around Lake Kivu.[342]

A glacier on Mount Kilimanjaro. The oldest now present ice of Kilimanjaro formed during the African humid period

Glaciers on Mount Kilimanjaro expanded during the African humid period[343] after a phase during the Younger Dryas where the mountain was ice free,[344] but the tree line also rose at that time, accompanied by soil formation.[345] The wetter climate may have destabilized the neighbouring Mount Meru volcano, causing a giant landslide that removed its summit.[346]

Erosion in catchments of East Africa increased with the beginning of the humid period but then decreased even before its end,[347] as the increased weathering led to the formation of soils, these in turn to the establishment of a vegetation cover that subsequently reduced additional erosion.[348]

Surprisingly, and contrary to the patterns expected from precessional changes, the East African Rift also featured wetter climates during the African humid period,[113] which extended at least as far south as Lake Rukwa into the Southern Hemisphere.[349] In the region of the African Great Lakes, pollen evidence points to the occurrence of forests including rainforest vegetation, while today they occur only in limited areas,[350] due to the increased precipitation.[351] Denser vegetation also occurred at Lake Turkana.[352] Different types of vegetation, including dryland vegetation, existed at Lake Malawi and Lake Tanganyika however,[353] and vegetation changes there were not overly strong.[354] Development of forest vegetation around the African Great Lakes created an interconnected environment where species spread, increasing biodiversity with effects on the future when the environment became fragmented.[355]

In East Africa, the African humid period led to improved environmental conditions in terms of food and water supply from large lakes that allowed early human populations to grow in size and survive without requiring major changes in food gathering strategies.[356] Earlier wet and dry periods in East Africa may have influenced the evolution of humans[357] and allowed their spread across the Sahara[358] and into Europe.[359]

Other parts of Africa and the rainforest realm

Lake Bosumtwi in Ghana rose during the African humid period,[338] and forests expanded in the Adamawa Plateau of Cameroon.[360] On Fuerteventura in the Canary Islands there is evidence of a moister climate during the African humid period.[361] The core of the rainforest was probably unaltered by the African humid period, perhaps with some changes in species[362][363] and an expansion of their area,[47] although the peatlands of Central Congo started developing during the African humid period and peat continues to accumulate there to this day.[364]

Levant and Mediterranean

High latitude Africa did not undergo large scale changes in the past 11,700 years,[101] only a few floods in Tunisian rivers appear to correlate with the African humid period,[365] although ecosystem changes consistent with a humid period have been invoked to explain a decrease in certain rodents of Northern Africa that depend on steppe habitats.[366]

In the Pleistocene and Holocene humidity in the Mediterranean is often correlated to humidity in the Sahara,[367] and the early-mid Holocene climate of Iberia, Italy, Negev and Northern Africa was wetter than today[368] which in Sicily correlates with Intertropical Convergence Zone changes in Northern Africa.[369] Mediterranean precipitation is brought by mid-latitude cyclones[367] and either increased precipitation from the westerlies[370] or monsoonal precipitation extending into the Mediterranean may have rendered it wetter,[371] although the connection between the African Monsoon and Mediterranean precipitation is unclear.[372][367]

The Mediterranean Sea became less saline during the African humid period, in part due to increased precipitation from the westerlies[370] but also from increased river discharge in Africa, leading to the formation of sapropel layers when the increased runoff led to the Mediterranean becoming more stratified.[373][374] The S1 sapropel layer is specifically associated with the African humid period[177] and with increased discharge of the Nile and other African rivers.[238] This together with decreased dust transport by wind led to changes in the sediment patterns[375] and an increased marine food web productivity in the Mediterranean.[376]

In the Levant, wetter conditions during the African humid period are recorded from Jeita Cave in Lebanon and Soreq Cave in Israel[377] while the Dead Sea and other southern European lakes were low during this period, unlike some earlier wet periods in the Sahara; possibly the stronger winter-summer insolation gradient in these earlier wet periods played a role in making the behaviour of these water bodies distinct.[378]

Southern Africa

The effects, if any, of the African humid period on Southern Africa have been unclear. Originally it was proposed that the orbitally driven changes would imply a dry period in Southern Africa which would have given way to moister conditions as the northern African humid period ended,[78] as the Intertropical Convergence Zone should shift its average position between the two hemispheres.[101] However, the lack of paleoclimatology data with sufficient time resolution from Southern Africa has made it difficult to assess the climate there during the African humid period.[78] More recently obtained paleoclimate data have suggested however that southern Africa was actually wetter during the African humid period rather than dryer,[379][380] perhaps reaching as far as 23° south[133] and as far as the catchment of the Orange River.[381] Additionally, wetter periods unrelated to the African humid period may have occurred after deglaciation in Southern Africa.[382] Particular changes occurred in central southern Africa, where a dry period co-occurred with an expansion of Lake Makgadikgadi; presumably increased wetness over the Okavango River catchment in the Angolan Highlands due to the African humid period nourished the lake during a dry interval.[383] Conversely, and consistent with the opposite reaction pattern of the Southern Hemisphere, the Zambezi River reached its lowest discharge during the African humid period.[384] In general there is little consistency between the Northern and Southern Africa in terms of hydrological changes during the Holocene.[385]

Numerical estimates

During the African humid period, Saharan rainfall increased to no more than 500 millimetres per year (20 in/year),[386] with large uncertainty;[163] the increases have been estimated to amount to 300–400 millimetres per year (12–16 in/year),[387] and values exceeding 400 millimetres per year (16 in/year) may have spread to 19-21° northern latitude.[388] An area with less than 100 millimetres per year (3.9 in/year) may have remained in the Eastern Sahara however,[389] although its driest parts may have received 20-fold more precipitation.[296]

Other reconstructed values of the precipitation increase indicate an annual increase of about 150–320 millimetres (5.9–12.6 in) in Africa,[390] with strong regional variation.[391] From lake levels precipitation increases of 20-33%[392] or 50-100% have been inferred for East Africa,[154] with an increase of 40% reconstructed for Northern Africa.[393] In the early Holocene, there appears to have been an eastward- and northward-decreasing trend of humidity.[394] Additionally, at Tayma in Arabia a threefold increase appears to have occurred[395] while precipitation in the Wahiba Sands of Oman may have reached 250–500 millimetres per year (9.8–19.7 in/year).[396]

Global

The African humid period was accompanied by a warmer climate, the hypsithermal,[259] which has been recorded from Africa,[47] Arabia,[397] the Caribbean[398][i] and the Mediterranean for example.[371] In the Rwenzori Mountains, forests during the period may record warmer temperatures.[399] Conversely, based on a drill core taken at Saraya, Senegal, temperatures during the African humid period were 1 °C (1.8 °F) lower than today there.[400] An increase in atmospheric methane concentrations, detected in Greenland ice cores about 14,700 years ago, was probably a consequence of growing tropical wetlands.[87]

Effect on other climate modes

One climate model has indicated that a greener Sahara and reduced dust output would induce increases tropical cyclone activity, especially over the Atlantic but also in most other tropical cyclone basins due to changes in the intensity of the storms, decreases in wind shear, changes in atmospheric circulation and less dust in the atmosphere, which results in warmer oceans,[401] despite an expected decrease of tropical wave activity over the Atlantic in climate models.[402] While there are no good paleotempestology data for the time of the African humid period that could confirm or refute this theory,[403] hurricane activity[404] including past strikes in Puerto Rico[405] and in Vieques appear to correlate with the strength of the West African Monsoon.[406] Finally, the northward movement of the Intertropical Convergence Zone during the African humid period may have caused a corresponding northward movement of tropical cyclogenesis areas and storm tracks in the Atlantic Ocean.[407]

The El Nino-Southern Oscillation is a major climate variability mode. Paleoclimatology records from Ecuador and the Pacific Ocean indicate that during the early and middle Holocene ENSO variability was suppressed by about 30-60%, which can be only partially explained through orbital forcing.[408] The Green Sahara may have suppressed ENSO activity, forcing a La Nina-like climate state,[405] in a climate model this is accompanied by decreased upwelling and deepening of the thermocline in the Eastern Pacific as the Walker circulation shifts westward.[409] In addition Atlantic Nino sea surface temperature patterns develop in the Atlantic Ocean.[410]

Fluctuations

Temperatures in Greenland during the Younger Dryas

Some gaps with less precipitation took place during the late glacial and the Holocene.[181] During the Younger Dryas 12,500-11,500 years ago, the North Atlantic and Europe became much colder again and there was a phase of drought in the area of the African humid period,[411][412] extending over both East Africa[j] and West Africa and also to India[414] and the Mediterranean[415] where dune activity occurred in the Negev,[416] with lake water levels dropping in much of East Africa[417][418] and drought extending over southern Africa.[419] At the end of the Younger Dryas, precipitation, lake levels and river runoff increased again, although south of the equator the return of humid conditions was slower than the relatively abrupt change to its north.[420][395]

Another dry phase took place about 8,200 years ago, spanning East Africa[139] and Northern Africa[k] as documented by various lines of evidence,[422] accompanied by cooling in the Northern Atlantic[423] and surrounding landmasses such as Greenland;[424] it may coincide with the 8.2 kiloyear event.[411] Such abrupt changes have been particularly noticeable in dust records off northwestern Africa, where they took decades to centuries.[425] The 8,200 year event has also been noted in the Maghreb, where it is associated with a transition of the Capsian culture[426] as well as with cultural changes both in the Sahara and the Mediterranean;[254] at the Gobero cemetery a population change occurred after this dry interruption.[427] This episode appears to have been the consequence of ice-dammed lakes in North America draining[79] although a low latitude origin has also been suggested.[428]

Cooling of the Northern Atlantic during Heinrich event 1 and the Younger Dryas associated with a weaker Atlantic meridional overturning circulation leads to atmospheric pressure anomalies that shift the Tropical Easterly Jet and precipitation belts south, making Northern Africa drier. [144][160][429] Earlier Heinrich events were also accompanied by North Africa drought.[40] Likewise, a weakening of moisture transport and a less eastward position of the Congo Air Boundary contributed to reducing precipitation in East Africa[414] although some parts of southern Africa at Lake Malawi were wetter during the Younger Dryas.[430]

Many of these variations in the early Holocene appear to be caused by the discharge of meltwater from the Laurentide Ice Sheet into the Atlantic, which weakens the Atlantic meridional overturning circulation and shift storm tracks north away from the Mediterranean.[431] Some dry periods in marine cores in the Gulf of Guinea appear to coincide with events recorded in Greenland ice cores.[432] Other variations in precipitation observed in records have been attributed to solar activity changes,[10] water levels of Lake Turkana for example appear to reflect the 11-year solar cycle.[433]

In Lake Turkana, water level fluctuations took place between 8,500 and 4,500 years before present, with highstands before 8,400, around 7,000 and between 5,500 and 5,000. These appear to be controlled by sea surface temperature patterns in the Atlantic and Indian Oceans, but also by overflow of water from Lake Suguta and the Chew Bahir basins into Lake Turkana,[434] which themselves received water from additional lakes.[335] Volcanic and tectonic phenomena occur at Lake Turkana, but do not have the magnitude required to explain large changes in lake level.[435] Water level fluctuations have also been inferred for Lake Chad on the basis of pollen data, especially towards the end of the African humid period.[436] In the Taoudenni lake fluctuations of about a quarter-millennium have been recorded[437] and frequent droughts occurred in the Eastern Sahara.[438]

Other variations appear to have occurred 9,500 - 9,000 and 7,400 - 6,800[216] as well as 10,200, 8,200, 6,600 and 6,000 years before present; they were accompanied by decreased population density in parts of the Sahara,[431] and other dry interludes in Egypt have been noted 9,400 - 9,300, 8,800 - 8,600, 7,100 - 6,900 and 6,100 - 5,900 years ago.[439] During dry episodes, humans might have headed to waterbodies which still had resources.[258] Aside from fluctuations, a southward retreat of the humid period may have been underway after 8,000 years ago.[440]

End

The African humid period ended about 6,000-5,000 years ago[12][441] around 5,500 years before present.[442] After vegetation declined[34] sand claimed the Sahara which became barren,[110] accompanied by increases in dust export from the now-desert[431] and from dried up lakes[443] such as the Bodélé Basin, which today is the largest single source of dust on Earth.[444] The transition from the "green Sahara" to the present-day dry Sahara is considered to be the greatest environmental transition of the Holocene in northern Africa;[445] today almost no precipitation falls in the region.[31] The lakes dried up, mesic vegetation disappeared, and sedentary human populations were replaced by more mobile cultures.[12] The end of the African humid period but also its beginning could be considered a "climate crisis" given the strong and extended impact[423] and the drying extended as far as the Canary Islands.[446]

The Piora Oscillation cold period in the Alps[447] coincides with the end of the African humid period;[286][448] the 5,600–5,000 calibrated years ago period was characterized by widespread cooling and more variable precipitation changes around the world[448] including a cooling of sea surface temperatures on both sides of the North Atlantic,[449] with some climate changes possibly extending into southeastern Australia.[450]

A major pan-tropical environmental change took place about 4,000 calibrated years ago.[451] This change was accompanied by the collapse of ancient civilizations, severe drought in Africa, Asia and the Middle East and the retreat of glaciers on Mount Kilimanjaro[452] and Mount Kenya.[453]

Chronology

Whether the drying happened everywhere at the same time and whether it took place in centuries or millennia is unclear[35][110] in part due to disagreeing records[454] and has led to controversy,[167][183] and such a disagreement on timing also exists with respect to the expected vegetation changes.[136][32] Most recently, the idea has taken hold that the end of the African humid period occurred from north to south in a stepwise fashion[248] with its end closer to the equator occurring between 4,000 and 2,500 years ago.[88][9]

A later end in northeast Africa about 4,000 years ago may reflect the different configuration of landmasses and thus monsoon behaviour.[455] North of the present-day monsoon belt and in the (western) Sahara the drying occurred in one step 6,000-5,000 years ago while south of it precipitation decreased in a more protracted fashion.[9][88][456] Distinct environmental changes may have occurred in Central Africa, Western Africa and East Africa.[183]

Some evidence points to a two-phase change in climate with two distinct dry transitions[457] caused by the existence of two different steps of insolation decrease at which climate changes.[458] Finally, sometimes the 4.2 kiloyear event is considered to be the true end of the African humid period.[428]

Marine cores usually indicate an abrupt change[459][107] but not without exceptions[167] while pollen data do not, perhaps due to regional and local differences in vegetation.[460] Furthermore, groundwater and local vegetation can modify local conditions;[248] groundwater-fed water bodies for example persisted longer than those nourished by rain.[201]

Increased variability in precipitation may have preceded the end of the African humid period; this is a commonly observed pattern before a sudden change in climate.[461] In Gilf Kebir, between 6,300 and 5,200 years ago apparently a winter rainfall regime became established as the African humid period ended.[149]

Sahara and Sahel

After a first brief lake level drop between 5,700 and 4,700 calibrated years ago that might reflect climate variability towards the end of the African humid period,[462] water levels in Lake Megachad decreased quickly after 5,200 years before present[463] with a water body persisting in the Bodele depression until 1,000 calibrated years ago at the least.[464]

Lake Chad shrank to about 5% of its former size, with the deeper northern Bodele basin drying up entirely as it was disconnected from the southern basin where its major tributary, the Chari River, enters the lake.[187] The dried out basin was now exposed to the Harmattan winds, which blow dust out of the dry lake bed[465] which is the single largest source of dust in the world.[466]

The tropical vegetation was replaced by desert vegetation, in some places suddenly and in others the change was more gradual.[467] Along the Atlantic coast, the vegetation retreat was slowed by a stage of sea level rise that increased soil moisture levels, delaying the retreat by about two millennia.[468][469] In Libya at Wadi Tanezzuft the end of the humid period was also delayed by leftover water in dune systems and in the Tassili mountains until 2,700 years ago, when river activity finally ceased.[58][470] A brief moist pulse between 5,000 - 4,000 years ago in the Tibesti led to the development of the so-called "Lower Terrace".[471]

At Lake Yoa, which is groundwater-fed, vegetation decreased and became desert vegetation between 4,700-4,300 and 2,700 years ago, while the lake became hypersaline 4,000 years ago;[472][473][474] however the climate there may have been affected by the Tibesti Mountains and the end of the African humid period thus delayed,[463] and the lake is nourished by fossil groundwater left by the African humid period to this day.[475]

East Africa and Arabia

In northern East Africa, water levels dropped rapidly about 5,500 years before present[123] while in Hoti cave in Arabia a southward retreat of the Indian Monsoon took place about 5,900 years ago.[89] Reconstructions from Lake Abiyata in Ethiopia suggest that the end of the African humid period took the form of severe droughts rather than a gradual decrease of precipitation.[476] Drying is also documented from Oman and the Blue Nile basin[96] with a noticeable decrease of Nile discharge about 4,000 calibrated years ago.[375] Decreased discharge of the Nile led to the cessation of sapropel deposition and turbidite activity off its delta,[85] and rivers and lakes of Arabia became intermittent or entirely dry.[477]

Some data from Ethiopia and the Horn of Africa indicate that drying there may have begun already 7,000–8,000 years ago or earlier.[418][306] Drying in Arabia commenced about 7,000 calibrated years ago[316] and there are large disparities in the timing between various parts of Arabia[397] but a tendency towards an arid climate between 6,000 and 5,000 years ago has been observed[478] which continued until 2,700 years ago.[308]

Forest cover in the area of the African Great Lakes decreased between 4,700 and 3,700 years ago,[350] although drying at Lake Victoria began already 8,000 years ago,[353] at Lake Rukwa 6,700 years ago,[349] at Lake Tanganyika about 6,000 years ago[353] and at Lake Edward major changes in lake chemistry consistent with drying are noted 5,200 years ago. There a minor recovery in vegetation took place between 2,500 and 2,000 years ago, followed by a much more rapid appearance of grasses accompanied also by substantial wildfire activity. This might have been the most severe drought of the Lake Edward region in the Holocene, with many lakes such as Lake George dropping significantly or drying up altogether.[479] Other lakes such as Nakuru, Turkana, Lake Chew Bahir, Lake Abbe and Lake Zway also dropped between 5,400–4,200 years ago.[480]

The end of the African humid period at Lake Turkana occurred about 5,300 years before present,[l] accompanied by a lake level decline[482] and the cessation of overflow from other lakes in its area into Lake Turkana.[334] Between 5,000 and 4,200, Lake Turkana became more saline and its water levels decreased below the level of outflow to the Nile.[483] Towards the end of the African humid period water temperatures in the lake and in other regional lakes appear to have increased, followed by a drop after its end[484] possibly resulting from the insolation seasonality pattern that was in force at the time of the end of the African humid period.[485] The decrease of water levels in Lake Turkana also impacted the Nile and the Predynastic societies dependent on it,[486] with Nile flow decreasing.[487]

Mediterranean

Libya and the Middle Atlas became gradually more dry.[467] In Morocco, drying is observed about 6,000 radiocarbon years ago[457] while drier conditions in Iberia accompanied the end of the African humid period between 6,000 and 4,000 years ago perhaps as a consequence of increasingly frequent positive North Atlantic Oscillation episodes.[488] A 4.2 kiloyear event is recorded in dust records from the Mediterranean[489] and might have been caused by changes in the Atlantic Ocean circulations.[490]

Tropical West Africa

In Lake Bosumtwi the African humid period ended about 3,000 years ago, farther north between 5,000 and 6,000 years ago,[110] and at Lake Bosumtwi there was a brief moistening between 5,410 ± 80 years ago that ended 3,170 ± 70 years ago. This, earlier but similar changes off western Senegal and later but similar changes in the Congo Fan appear to reflect a southward shift of the precipitation zone over time.[429] Some drying occurred simultaneously between the Sahel and the Gulf of Guinea.[160] Some lakes in the Guineo-Congolian region dried out, while others were relatively unaffected.[468]

A general tendency towards a drier climate is observed in West Africa when the African humid period ended.[491] There, dense vegetation became progressively thinner between 5,000 and 3,000 years ago,[479] with the establishment of aridity between 5,200–3,600 years ago in the Sahara[492] and major perturbations around 4,200 and 3,000–2,500[493]/2,400 calibrated years ago.[494] A brief return of moister conditions took place 4,000 years ago[423] while a substantial dry phase occurred between 3,500 and 1,700 years ago.[491] In Senegal modern-type vegetation arose about 2,000 years ago.[495]

Central Africa

Farther south at the equator between 6,100 and 3,000 calibrated years before present savannah expanded at the expense of forests, with the transition possibly lasting until 2,500 calibrated years before present;[451] a different time course estimate for the area between 4° southern and 7° northern latitude states that forest cover decreased between 4,500–1,300 years ago.[468] In the Adamawa Plateau (Cameroon[496]), the Ubangui Plateau (Central African Republic[496]) and the Cameroon Volcanic Line montane forests disappeared at the end of the African humid period.[497] In the Adamawa Plateau savannah has continuously expanded since 4,000 calibrated years ago due to an increased establishment of an annual dry season.[497] Such a change took also place in Benin and Nigeria between 4,500 and 3,400 calibrated years ago.[468] Many vegetation changes in the tropical regions were probably caused by a longer dry season[498] and perhaps a smaller latitudinal range of the Intertropical Convergence Zone.[494]

Southern Hemisphere Africa

In the Southern Hemisphere at Lake Malawi drying began later – 1,000 years before present – as did the African humid period which there began only about 8,000 years ago.[484] Contrarily, increased water levels in Etosha Pan (Namibia) appear to relate to a southward movement of the Intertropical Convergence Zone at the end of the African humid period[499] although stalagmite growth data in Dante Cave also in Namibia has been interpreted as indicating a wetter climate during the African humid period.[379]

Mechanisms

The end of the humid period appears to reflect the changes in insolation during the Holocene,[88] as a progressive decrease of summer insolation caused the insolation gradients between Earth's hemispheres to decrease.[500] However, the drying appears to have been much more abrupt than the insolation changes;[107] It is not clear whether non-linear feedbacks led to abrupt changes in climate or whether the process, driven by orbital changes, was abrupt.[110] Also, the Southern Hemisphere warmed and this resulted in a southward shift of the Intertropical Convergence Zone;[501] orbitally-driven insolation has increased over the Holocene in the Southern Hemisphere.[100]

As precipitation decreased, so did vegetation, in turn increasing the albedo and further decreasing precipitation.[114] Furthermore, vegetation may have responded to increased variations in precipitation towards the end of the African humid period[111] although this view has been challenged.[502] This could have directed sudden changes in precipitation, although this view has been cast in doubt by the observation that in many places the end of the African humid period was gradual rather than sudden.[503] There might be differences between plants at higher and lower latitudes, in how they respond to climate change; for example more diverse plant communities might have slowed down the end of the African humid period.[60]

Other proposed mechanisms:

  • Decreases in polar insolation through altered cosmic ray fluxes might promote the growth of sea ice and cooling at high latitudes, which in turn result in stronger equator-to-pole temperature gradients, stronger subtropical anticyclones and more intense upwelling in for example the Benguela current.[156]
  • Changes in the circulation of high latitude oceans may have played a role as well,[500] such as the potential occurrence of another meltwater/ice rafting pulse around 5,700 years before present.[501] The decreased insolation during the mid-Holocene may have made the climate system more sensitive to changes, explaining why earlier comparable pulses did not terminate the humid period for good.[504]
  • In climate models, increased snow and ice on the Tibetan Plateau can lead to a weakening of the Indian and African monsoons, with the weakening of the former preceding that of the latter by 1,500–2,000 years.[505] There is evidence that glaciers in Tibet such as at Nanga Parbat expanded during the Holocene, especially towards the end of the African humid period.[506]
  • Decreases in sea surface temperatures of the Indian Ocean may be involved in the drying of East Africa, however the temperature records from that ocean are not agreed upon.[138] Moreover, there is no evidence of temperature changes in the Gulf of Guinea at the critical time that might explain the end of the African humid period.[123]
  • Additional feedback processes may have included the drying of soils and loss of vegetation after decreased rainfall,[110] which would have led to wind-driven deflation of the soils.[507]
  • An expansion of sea ice around Antarctica about 5,000 calibrated years ago may have provided another positive feedback.[508]
  • The expanding dry belt of the Sahara pushed the regions of cyclogenesis in the Mediterranean northwest-northward, resulting in changes of the wind[509] and precipitation regime in parts of Italy.[510]
  • High latitude climate changes have been proposed as a cause for the end of the African humid period. Specifically, about 6,000–5,000 years ago the Arctic became colder, with sea ice expanding, temperatures in Europe and off Northern Africa decreasing and the Atlantic meridional overturning circulation weakening.[123] This cooling tendency may have weakened the Tropical Easterly Jet and thus reduced the amount of precipitation falling over Africa.[511]

The orbitally-induced changes of precipitation may have been modified by the solar cycle; specifically, solar activity maxima during the ending phase of the African humid period may have offset the orbital effect and thus stabilized precipitation levels, while solar activity minima compounded the orbital effects and thus induced rapid decreases in lake water levels of Lake Turkana.[512] At Lake Victoria on the other hand, solar variations appear to sometimes lead to drough and sometimes lead to wetness, probably due to changes in the Intertropical Convergence Zone.[501]

Potentially human-mediated changes

About 2,000 years ago, major changes in vegetation in East Africa may have been caused by human activity, including large-scale deforestation for iron production during the Iron Age.[513] Similar changes have been observed on the Adamawa Plateau[514] (Cameroon[496]) but later dating of archeological sites has found no correlation between human expansion in Cameroon and environmental degradation.[515] Similar rainforest degradation across Western African took place between 3,000 and 2,000 years ago.[516] Climate-mediated processes may have increased the impact of land use changes in East Africa.[355] In the Sudanian and Sahelian savannah on the other hand human activity seems to have had little impact,[187] and in Central Africa forest changes were clearly triggered by climate change with little or no evidence of anthropogenic changes.[517] The question has led to intense debate among paleoecologists and archeologists.[518]

While humans were active in Africa during the end of the African humid period, climate models analyzed by Claussen and colleagues 1999 indicate that its end does not need any human activity as an explanation[519] although vegetation changes may have been induced by human activity.[185] Later it was suggested that overgrazing may have triggered the end of the African humid period around 5,500 years ago;[248] human influence might explain why the Sahara became a desert without the accompanying onset of an ice age; usually the existence of a Sahara desert is associated with the expansion of high latitude glaciers.[288] Later research has on the contrary suggested that human pastoralism may have actually delayed the end of the African humid period by half a millennium[520] as moving herds of animals driven by humans seeking good pasture conditions may lead to more balanced impacts of pastures on the vegetation and thus to greater vegetation quality.[521]

Global

A general drying tendency is observed in the northern tropics[522] such as in Asia more generally between 5,000 and 4,500 calibrated years ago,[523] with Asian monsoon precipitation declining between 5,000 and 4,000 years ago.[18] A drought 5,500 years ago is recorded in Mongolia[524] and eastern America, where drought conditions around 5,500–5,000 years ago occurred in places like Florida, New Hampshire and Ontario.[525][526] A drying tendency is also noted in the Caribbean and the Central Atlantic.[527]

Conversely, in South America there is evidence that the monsoon behaves in an opposite fashion consistent with precessional forcing;[522] water levels in Lake Titicaca were low during the middle Holocene and began to rise again after the end of the African humid period.[528] Likewise, a trend towards increased wetness took place in the Rocky Mountains at this time[529] although it was accompanied by a drier phase around Lake Tahoe, California and in the Western United States.[530]

Consequences

Humans

As observed in archeological sites, population in Northern Africa decreased between 6,300–5,200 years ago[110] over less than a millennium,[507] and in inner Arabia many settlements were abandoned about 5,300 years ago.[118] Some Neolithic people in the desert persisted for longer thanks to the exploitation of groundwater.[457]

Different human populations responded to the drying in diverse manners,[280] with human responses in the Western Sahara being distinct from the reactions in the Central Sahara.[6] In the Central Sahara, pastoralism replaced hunter-gatherer activity[531] and a more nomadic lifestyle replaced semi-sedentary lifestyles[532] as observed in the Acacus Mountains of Libya.[263] Nomadic lifestyles also developed in the Eastern Sahara/Red Sea Hills in response to the end of the African humid period.[533] Finally, there was a shift in domestic animal use from cattle to sheep and goats as these are more suitable for arid climates, a change reflected in rock art from which cattle disappeared at this time.[534]

The development of irrigation systems in Arabia may have been an adaptation to the drying tendency.[316] The decreased availability of resources forced human populations to adapt,[535] in general fishing and hunting declined in favour of farming and herding.[536] However, the effects of the end of the African humid period on human food production have been subject to controversy.[537]

The pyramids of Giza, the most recognizable trace left by the Egyptian civilization

The warm episode and coinciding drought may have triggered animal and human migration to less inhospitable areas[485] and the appearance of pastoralists where previously fishery-dependent societies had existed as happened at Lake Turkana.[336] Humans moved to the Nile, where the society of Ancient Egypt with pharaohs and pyramids was eventually forged by these climate refugees[538][507][539] perhaps reflecting renewed exuberance;[286] thus the end of the African humid period can be considered responsible for the birth of Ancient Egypt.[539][1] Lower water levels in the Nile also aided the settlement of its valley as has been observed at Kerma.[540] A similar process may have led to the development of the Garamantian civilization.[541] Such human migrations towards more hospitable conditions along rivers and the development of irrigation also took place along the Euphrates River, Tigris River and Indus River, leading to the development of the Sumerian and Harappan civilizations.[542][60] Population shifts into mountain areas have also been reported for the Air Mountains, Hoggar and Tibesti.[393] In other places, such as the Acacus Mountains populations conversely remained in oases.[543]

The Nile itself was not totally unaffected however;[330] the end of the African humid period may also be linked to the collapse of the Old Kingdom in Egypt[35] when the Nile floods failed for three decades around 4,160 years before present.[544] The ongoing decrease of precipitation after the end of the African humid period may also be the cause of the end of the Akkadian Kingdom in Mesopotamia.[545] The end of the Garamantian civilization may also relate to climate change although other historical events were probably more important;[546] at Tanezzuft oasis it certainly relates to the drying trend after 1,600 years ago.[543]

In Central Africa, forests fragmented and savannahs formed in some places, which facilitated the movement and growth of Bantu speaking populations;[503] these in turn may have affected the ecosystem.[547] The vegetation changes may have aided in the establishment of agriculture.[517] The relatively slow decline of precipitation gave humans more time to adapt to the changing climate conditions.[356]

Cultural changes may also have occurred as a consequence of climate change, such as[548] changes in gender roles, the development of elites,[549] the increased presence of human burials where formerly cattle burials predominated,[550] as well as an increase of monumental architecture in the Sahara may have also been a response to increasingly adverse climates.[531] A spread in cattle domestication at the time of climate change[263] may also relate to these events, although its role is controversial.[548] Finally, changes in agricultural practices at the end of the African humid period may be associated with the propagation of malaria and one of its causative pathogens Plasmodium falciparum; in turn these may correlate with the origin of human genome variants such as sickle cell disease that are linked to malaria resistance.[551]

Non-human

In the Sahara, animal and plant populations were fragmented and restricted to certain favoured areas such as moist areas of mountain ranges; this happened for example to fish and crocodiles which only persist in isolated water bodies and to Mediterranean plants[552][553] such as cypresses which persist only in mountains,[554] along with some reptiles that may have also been stranded there by the drying.[555] The buffalo species Syncerus antiquus probably went extinct from the increased competition of pastoralists triggered by the climate drying.[556] Gorilla populations became split in West African and East African populations by the drying of the African Great Lakes region,[351] and a similar population split between the insect species Chalinus albitibialis and Chalinus timnaensis in Northern Africa and the Middle East may have also been caused by the expansion of deserts there.[557] Giraffes, widespread in the Sahara during the African humid period, may have been forced to migrate into the Sahel; this together with the separating effect of Lake Megachad may also have influenced the development of giraffe subspecies.[558] Climate change together with human impacts may have led to the extinction of a number of large mammals in Egypt.[559]

The Dahomey Gap[m] formed 4,500–3,200 years before present, correlative to the end of the African humid period.[561] The harbour porpoise declined in the Mediterranean due to a switch to oligotrophic conditions as discharge from African rivers decreased.[376] Desert varnish formed on some exposed rocks in the Sahara.[562]

Global climate

The shrinkage of subtropical wetlands probably led to a drop in atmospheric methane concentrations between 5,500 and 5,000 years ago, before boreal wetlands expanded and offset the loss of subtropical wetlands, leading to a return of higher atmospheric methane concentrations.[423] Conversely, atmospheric carbon dioxide increased after about 7,000 years as the biosphere began releasing carbon in response to increasing aridity.[545]

Dust originating in the Bodele depression

A sudden increase in the amount of land-originating dust in an oceanic drill core off Cape Blanc, Mauritania, has been interpreted as reflecting the end of the African humid period 5,500 years ago occurring in only a few centuries.[563] Potentially, dried up lake basins became an important source for dust.[474] Today, the Sahara is the single largest source of dust in the world, with far ranging effects on climate and ecosystems,[564] such as the growth of the Amazon rainforest.[565]

The period 5,500–5,000 years ago also witnessed major changes in global climate, including the onset of global cooling in the form of the Neoglacial.[566] In one climate model, the desertification of the Sahara at the end of the African humid period reduces the amount of heat transported in the atmosphere and ocean towards the poles, inducing cooling of 1–2 °C (1.8–3.6 °F) especially in winter in the Arctic and an expansion of sea ice. Reconstructed temperatures in the Arctic indeed show a cooling, although less pronounced than in the climate model.[567] Further, this climate transition in the climate model is accompanied by increased negative Arctic Oscillation states, a weaker subpolar gyre and increased precipitation and cold air outbreaks in much of Europe; such changes have also been observed in paleoclimate data.[568] These findings imply that the vegetation state of the Sahara influences the Northern Hemisphere climate.[569] In turn, this high latitude cooling may have further reduced precipitation over Africa.[511]

Present-day situation

Presently, the African Monsoon still influences the climate between 5° south and 25° north latitude; the latitudes around 10° north receive the bulk of their precipitation from the monsoon[n] during summer, with smaller amounts of rainfall occurring farther north. Thus farther north deserts can be found while the moister areas are vegetated.[111] In the Central Sahara, annual precipitation reaches no more than 50–100 millimetres per year (2.0–3.9 in/year).[571] Even farther north, the margin of the desert coincides with the area where the westerlies bring precipitation;[2] they also influence southernmost Africa.[572] The existence of the deserts is because of the subsidence of air over parts of Northern Africa, which is further increased by the radiative cooling over the desert.[1] In East Africa the monsoon leads to two rain seasons in the equatorial area, the so-called "long rains" in March–May and the "short rains" in October–November[573] when the Intertropical Convergence Zone moves northward and southward over the region, respectively;[574] in addition to the Indian Ocean-sourced precipitation there is also Atlantic- and Congo-sourced precipitation west of the Congo Air Boundary.[570][573] Climate variability exists to this day, with the Sahel suffering from droughts in the 1970s and 1980s when precipitation decreased by 30% and the flow of the Niger River and Senegal River even more,[575] followed by an increase of precipitation.[1] In Arabia, the monsoon does not penetrate far from the Arabian Sea and some areas are under the influence of winter precipitation brought by cyclones from the Mediterranean Sea.[576] East Africa is also under the influence of monsoon circulations.[577]

Implications for future global warming

Greening of the Sahel between 1982 and 1999

Some simulations of man-made global warming and increased carbon dioxide concentrations in the atmosphere have observed a substantial increase in precipitation in the Sahel/Sahara. This could lead to an expansion of vegetation into present-day desert, although it would be less extensive than during the mid-Holocene[108] and perhaps accompanied by a northward shift of the desert, i.e. a drying of northernmost Africa.[578] Such a precipitation increase may also reduce the amount of dust originating in Northern Africa,[579] with effects on hurricane activity in the Atlantic and increased threats of hurricane strikes in the Caribbean, the Gulf of Mexico and the East Coast of the United States of America.[403]

The Special Report on Global Warming of 1.5 °C and the IPCC Fifth Assessment Report indicate that global warming will likely result in increased precipitation across most of East Africa, parts of Central Africa and the principal wet season of West Africa, although there is significant uncertainty related to these projections especially for West Africa.[580] On the other hand, West Africa[581] and parts of East Africa may become drier during given seasons and months.[581][582] In addition, the end of the 20th century drying trend may be due to global warming.[582] Currently, the Sahel is becoming greener but precipitation has not fully recovered to levels reached in the mid-20th century.[578]

Climate models have yielded equivocal results about the effects of anthropogenic global warming on the Sahara/Sahel precipitation, and it has to be considered that man-made climate change occurs through different mechanisms than the natural climate change that led to the African humid period and that vegetation feedbacks often are not taken into account.[583] One study in 2003 showed that vegetation intrusions in the Sahara can occur within decades after strong rises in atmospheric carbon dioxide[584] but would not cover more than about 45% of the Sahara.[37] That climate study also indicated that vegetation expansion can only occur if grazing or other perturbations to vegetation growth do not hamper it.[585]

A greening of the Sahara on the one hand may allow agriculture and pastoralism to expand into hitherto unsuitable areas, but increased precipitation can also lead to increased water borne diseases and flooding.[586] Also, expanded human activity may be vulnerable to climate reversals as demonstrated by the droughts that followed the mid-20th century wet period.[587]

See also

Notes

  1. ^ Dune-covered areas.[68]
  2. ^ Earlier it was thought that it had started about 9,000 years ago, before it was found that it probably began earlier and was interrupted by the Younger Dryas.[50] Alternatively, wet conditions in the Sahara and Sahel began about 10,900–10,500 years ago.[79]
  3. ^ This was originally believed to have occurred 7,000 or 13,000 years before present,[81] but a more recent suggestion indicates a reconnection of the Nile 14,000–15,000 years ago.[93]
  4. ^ The Congo Air Boundary is the point at which moisture bearing winds from the Indian Ocean collide with these from the Atlantic Ocean.[133]
  5. ^ Where the Indian Summer Monsoon penetrated farther inland[10] and was more intense starting about 14,800 years ago.[80]
  6. ^ Salt deposits left there were mined beginning in the 16th century.[220]
  7. ^ Fesselsteine are stony artifacts, that are interpreted as tools for restraining animals.[278]
  8. ^ Also known as the Yellow Nile[325]
  9. ^ In the Caribbean, a wet period has been identified in the mid-Holocene which correlated with the African wet period and was preceded and followed by drier conditions.[398]
  10. ^ There is conflicting evidence on whether the Younger Dryas was wetter or drier in tropical southeastern Africa.[413]
  11. ^ Whether it also took place in Asia is unclear; perhaps it was too short to trigger climate changes recognizable in records.[421]
  12. ^ Three different conceptual ideas for its termination exist.[481]
  13. ^ The Dahomey Gap is a region without forests in southern Benin, Ghana and Togo[560] that forms a gap in the Guineo-Congolian forest belt.[468]
  14. ^ The main area of monsoon rains does not coincide with the Intertropical Convergence Zone.[570]

References

  1. ^ a b c d e f g h i j k l m n Bader, Jürgen; Dallmeyer, Anne; Claussen, Martin (29 March 2017). "Theory and Modeling of the African Humid Period and the Green Sahara". Oxford Research Encyclopedia of Climate Science. 1. doi:10.1093/acrefore/9780190228620.013.532.
  2. ^ a b c Hoelzmann & Holmes 2017, p. 3.
  3. ^ Wendorf, Karlén & Schild 2007, p. 190.
  4. ^ Timm et al. 2010, p. 2612.
  5. ^ Hoelzmann et al. 2001, p. 193.
  6. ^ a b c d Stivers et al. 2008, p. 2.
  7. ^ a b c Watrin, Lézine & Hély 2009, p. 657.
  8. ^ Lézine, Duplessy & Cazet 2005, p. 227.
  9. ^ a b c d e f Skinner & Poulsen 2016, p. 349.
  10. ^ a b c Junginger et al. 2014, p. 1.
  11. ^ Hopcroft et al. 2017, p. 6805.
  12. ^ a b c d e f g Menocal et al. 2000, p. 348.
  13. ^ a b c d e Peck et al. 2015, p. 140.
  14. ^ Hoelzmann & Holmes 2017, p. 11.
  15. ^ a b Krüger et al. 2017, p. 1.
  16. ^ Sangen 2012, p. 144.
  17. ^ Médail et al. 2013, p. 1.
  18. ^ a b Lézine et al. 2017, p. 68.
  19. ^ a b c Runge 2013, p. 81.
  20. ^ a b Olsen 2017, p. 90.
  21. ^ a b c d Sangen 2012, p. 213.
  22. ^ Spinage 2012, p. 71.
  23. ^ Revel et al. 2010, p. 1357.
  24. ^ Said 1993, p. 128.
  25. ^ Brass, Michael (1 March 2018). "Early North African Cattle Domestication and Its Ecological Setting: A Reassessment". Journal of World Prehistory. 31 (1): 86. doi:10.1007/s10963-017-9112-9. ISSN 1573-7802.
  26. ^ a b Baumhauer & Runge 2009, p. 10.
  27. ^ a b Sangen 2012, p. 211.
  28. ^ Soriano et al. 2009, p. 2.
  29. ^ a b Pachur & Altmann 2006, p. 32.
  30. ^ Sepulchre et al. 2008, p. 42.
  31. ^ a b c d e f g h Menocal et al. 2000, p. 347.
  32. ^ a b c Russell & Ivory 2018, p. 1.
  33. ^ a b Quade et al. 2018, p. 1.
  34. ^ a b c d Schefuß et al. 2017, p. 2.
  35. ^ a b c d e f Costa et al. 2014, p. 58.
  36. ^ Blanchet et al. 2013, p. 98.
  37. ^ a b c Petoukhov et al. 2003, p. 99.
  38. ^ Chiotis 2018, p. 17.
  39. ^ Chiotis 2018, p. 20.
  40. ^ a b Röhl et al. 2008, p. 671.
  41. ^ a b c Zerboni, Trombino & Cremaschi 2011, p. 331.
  42. ^ Jones & Stewart 2016, p. 126.
  43. ^ Krüger et al. 2017, pp. 12–13.
  44. ^ Jones & Stewart 2016, p. 117.
  45. ^ Timm et al. 2010, p. 2627.
  46. ^ Hoelzmann & Holmes 2017, p. 10.
  47. ^ a b c Runge 2013, p. 65.
  48. ^ a b Petraglia & Rose 2010, p. 45.
  49. ^ a b c d Blümel 2002, p. 8.
  50. ^ a b c d e f g Adkins, Menocal & Eshel 2006, p. 1.
  51. ^ a b c Williams et al. 2010, p. 1131.
  52. ^ Riemer 2006, pp. 554–555.
  53. ^ Brooks et al. 2007, p. 255.
  54. ^ a b Baumhauer & Runge 2009, p. 28.
  55. ^ a b c Muhs et al. 2013, p. 29.
  56. ^ Kennett & Kennett 2007, p. 235.
  57. ^ a b Pachur & Altmann 2006, p. 6.
  58. ^ a b Brooks et al. 2007, pp. 258–259.
  59. ^ Petraglia & Rose 2010, p. 197.
  60. ^ a b c Bard 2013, p. 808.
  61. ^ Krüger et al. 2017, p. 14.
  62. ^ Sangen 2012, p. 212.
  63. ^ a b Williams et al. 2010, p. 1129.
  64. ^ Morrissey & Scholz 2014, p. 95.
  65. ^ a b Williams et al. 2010, p. 1134.
  66. ^ a b Castañeda et al. 2016, p. 54.
  67. ^ a b Runge 2010, p. 237.
  68. ^ Perego, Zerboni & Cremaschi 2011, p. 465.
  69. ^ Muhs et al. 2013, p. 42,44.
  70. ^ Gasse 2000, p. 195.
  71. ^ a b Brookes 2003, p. 164.
  72. ^ Maley 2000, p. 133.
  73. ^ Runge 2010, p. 234.
  74. ^ Maley 2000, p. 122.
  75. ^ a b Zerboni & Gatto 2015, p. 307.
  76. ^ Maley 2000, p. 127.
  77. ^ Moeyersons et al. 2006, p. 166.
  78. ^ a b c Reimer et al. 2010, p. 36.
  79. ^ a b Cremaschi et al. 2010, p. 89.
  80. ^ a b c Junginger et al. 2014, p. 12.
  81. ^ a b c Williams et al. 2006, p. 2652.
  82. ^ Pachur & Altmann 2006, p. 11.
  83. ^ Pachur & Altmann 2006, p. 601.
  84. ^ Talbot et al. 2007, p. 4.
  85. ^ a b c d e f Williams et al. 2010, p. 1132.
  86. ^ Hughes, Fenton & Gibbard 2011, pp. 1066–1068.
  87. ^ a b c d Menocal et al. 2000, p. 354.
  88. ^ a b c d e Peck et al. 2015, p. 142.
  89. ^ a b c d Petraglia & Rose 2010, p. 46.
  90. ^ Neugebauer, Ina; Wulf, Sabine; Schwab, Markus J.; Serb, Johanna; Plessen, Birgit; Appelt, Oona; Brauer, Achim (August 2017). "Implications of S1 tephra findings in Dead Sea and Tayma palaeolake sediments for marine reservoir age estimation and palaeoclimate synchronisation". Quaternary Science Reviews. 170: 274. Bibcode:2017QSRv..170..269N. doi:10.1016/j.quascirev.2017.06.020. ISSN 0277-3791.
  91. ^ Battarbee, Gasse & Stickley 2004, p. 242.
  92. ^ Williams et al. 2010, p. 1127.
  93. ^ Williams et al. 2006, p. 2664.
  94. ^ a b Hamdan & Brook 2015, p. 184.
  95. ^ a b Kuper 2006, p. 412.
  96. ^ a b Revel et al. 2010, p. 1358.
  97. ^ Barker et al. 2002, p. 302.
  98. ^ Moeyersons et al. 2006, p. 177.
  99. ^ Gasse 2000, p. 203.
  100. ^ a b Guilderson et al. 2001, p. 196.
  101. ^ a b c d e f Burrough & Thomas 2013, p. 29.
  102. ^ Vermeersch, Linseele & Marinova 2008, p. 395.
  103. ^ Röhl et al. 2008, p. 673.
  104. ^ Mercuri et al. 2018, p. 219.
  105. ^ Baumhauer 2004, p. 290.
  106. ^ a b Menocal et al. 2000, p. 356.
  107. ^ a b c Renssen et al. 2003, p. 1.
  108. ^ a b Renssen et al. 2003, p. 4.
  109. ^ Shi & Liu 2009, p. 3721.
  110. ^ a b c d e f g h i Menocal 2015, p. 1.
  111. ^ a b c d e f Hély et al. 2009, p. 672.
  112. ^ a b Shi & Liu 2009, p. 3722.
  113. ^ a b c d Tierney et al. 2011, p. 103.
  114. ^ a b c Renssen et al. 2006, p. 95.
  115. ^ Shi & Liu 2009, pp. 3720–3721.
  116. ^ Shi & Liu 2009, p. 3723.
  117. ^ a b Armitage, Bristow & Drake 2015, p. 8543.
  118. ^ a b c Beer et al. 2002, p. 591.
  119. ^ Battarbee, Gasse & Stickley 2004, p. 243.
  120. ^ a b c d e f Timm et al. 2010, p. 2613.
  121. ^ Servant, Buchet & Vincens 2010, p. 290.
  122. ^ a b Menocal et al. 2000, p. 357.
  123. ^ a b c d e Schefuß et al. 2017, p. 7.
  124. ^ a b Donnelly et al. 2017, p. 6222.
  125. ^ a b Gaetani et al. 2017, p. 7622.
  126. ^ a b c d Quade et al. 2018, p. 2.
  127. ^ a b The Hadley circulation : present, past and future. Advances in Global Change Research. Vol. 21. Kluwer academic Publishers. 2004. p. 339. doi:10.1007/978-1-4020-2944-8. ISBN 978-1-4020-2944-8.
  128. ^ a b Tierney et al. 2011, p. 110.
  129. ^ Cohen et al. 2008, p. 254.
  130. ^ a b c Vahrenholt & Lüning 2019, p. 529.
  131. ^ Burrough & Thomas 2013, pp. 29–30.
  132. ^ Tierney et al. 2011, p. 109.
  133. ^ a b c d Burrough & Thomas 2013, p. 30.
  134. ^ a b c Junginger et al. 2014, p. 13.
  135. ^ Costa et al. 2014, p. 64.
  136. ^ a b c Costa et al. 2014, p. 59.
  137. ^ Castañeda et al. 2016, p. 53.
  138. ^ a b Liu et al. 2017, p. 130.
  139. ^ a b Liu et al. 2017, p. 131.
  140. ^ Johnson, Thomas C.; Werne, Josef P.; Castañeda, Isla S. (1 September 2007). "Wet and arid phases in the southeast African tropics since the Last Glacial Maximum". Geology. 35 (9): 825. Bibcode:2007Geo....35..823C. doi:10.1130/G23916A.1. ISSN 0091-7613.
  141. ^ Hoelzmann & Holmes 2017, p. 31.
  142. ^ Barker et al. 2002, p. 295.
  143. ^ Barker et al. 2002, p. 296.
  144. ^ a b Timm et al. 2010, p. 2629.
  145. ^ Hoelzmann & Holmes 2017, p. 25.
  146. ^ a b Hamdan & Brook 2015, p. 185.
  147. ^ Phillipps et al. 2012, p. 72.
  148. ^ Petit-Maire 1989, p. 648.
  149. ^ a b Williams et al. 2010, p. 1133.
  150. ^ Prasad & Negendank 2004, pp. 219–220.
  151. ^ Baumhauer & Runge 2009, p. 6.
  152. ^ Linstädter & Kröpelin 2004, p. 763.
  153. ^ Skinner & Poulsen 2016, pp. 355–356.
  154. ^ a b c d e f g Bowman, D.; Nyamweru, C. K. (1 January 1989). "Climatic changes in the Chalbi Desert, North Kenya". Journal of Quaternary Science. 4 (2): 137. Bibcode:1989JQS.....4..131N. doi:10.1002/jqs.3390040204. ISSN 1099-1417.
  155. ^ Pachur & Altmann 2006, p. 276.
  156. ^ a b Reimer et al. 2010, p. 42.
  157. ^ Pachur & Altmann 2006, p. 556.
  158. ^ Schefuß et al. 2017, p. 3.
  159. ^ Hoelzmann & Holmes 2017, pp. 25–26.
  160. ^ a b c Schefuß et al. 2017, p. 5.
  161. ^ Mercuri et al. 2018, p. 225.
  162. ^ a b Prasad & Negendank 2004, p. 221.
  163. ^ a b Hopcroft et al. 2017, p. 6804.
  164. ^ Dixit et al. 2018, p. 234.
  165. ^ Pachur & Altmann 2006, p. 9.
  166. ^ Dixit et al. 2018, p. 247.
  167. ^ a b c Liu et al. 2017, p. 123.
  168. ^ a b Huang et al. 2008, p. 1459.
  169. ^ a b Engel et al. 2012, p. 131.
  170. ^ Hiner, Christine A.; Silveira, Emily; Arevalo, Andrea; Murrieta, Rosa; Lucero, Ricardo; Eeg, Holly; Palermo, Jennifer; Lachniet, Matthew S.; Anderson, William T.; Knell, Edward J.; Kirby, Matthew E. (2015). "Evidence for insolation and Pacific forcing of late glacial through Holocene climate in the Central Mojave Desert (Silver Lake, CA)". Quaternary Research. 84 (2): 9. Bibcode:2015QuRes..84..174K. doi:10.1016/j.yqres.2015.07.003. ISSN 1096-0287.
  171. ^ Huang et al. 2008, p. 1461.
  172. ^ a b Flögel, S.; Beckmann, B.; Hofmann, P.; Bornemann, A.; Westerhold, T.; Norris, R.D.; Dullo, C.; Wagner, T. (September 2008). "Evolution of tropical watersheds and continental hydrology during the Late Cretaceous greenhouse; impact on marine carbon burial and possible implications for the future". Earth and Planetary Science Letters. 274 (1–2): 10. Bibcode:2008E&PSL.274....1F. doi:10.1016/j.epsl.2008.06.011. ISSN 0012-821X.
  173. ^ Sangen 2012, p. 214.
  174. ^ Sangen 2012, p. 138.
  175. ^ a b c d e f Usai, Donatella (2 June 2016). A Picture of Prehistoric Sudan. Vol. 1. Oxford University Press. doi:10.1093/oxfordhb/9780199935413.013.56.
  176. ^ Liu et al. 2017, p. 127.
  177. ^ a b Wu et al. 2017, p. 95.
  178. ^ a b Stojanowski, Carver & Miller 2014, p. 80.
  179. ^ Vermeersch, Linseele & Marinova 2008, p. 396.
  180. ^ Chiotis 2018, p. 187.
  181. ^ a b Bristow et al. 2018, p. 182.
  182. ^ Hély et al. 2009, p. 685.
  183. ^ a b c Sylvestre et al. 2013, p. 224.
  184. ^ Lézine 2017, p. 4.
  185. ^ a b Baumhauer 2004, p. 291.
  186. ^ Watrin, Lézine & Hély 2009, p. 663.
  187. ^ a b c d e Bristow et al. 2018, p. 183.
  188. ^ Watrin, Lézine & Hély 2009, p. 668.
  189. ^ Lézine 2017, p. 5.
  190. ^ Watrin, Lézine & Hély 2009, p. 667.
  191. ^ Linstädter & Kröpelin 2004, p. 762.
  192. ^ Brookes 2003, p. 163.
  193. ^ a b White et al. 2011, p. 458.
  194. ^ Prasad & Negendank 2004, p. 225.
  195. ^ a b c White et al. 2011, p. 460.
  196. ^ Hopcroft et al. 2017, p. 6808.
  197. ^ a b Cole et al. 2009, p. 257.
  198. ^ a b c Stivers et al. 2008, p. 4.
  199. ^ a b Stivers et al. 2008, p. 11.
  200. ^ Petit-Maire 1989, p. 641.
  201. ^ a b c Mercuri et al. 2018, p. 221.
  202. ^ Pachur & Altmann 2006, p. 528.
  203. ^ Gross et al. 2014, p. 14472.
  204. ^ Cooper, Alan; Llamas, Bastien; Breen, James; Burns, James A.; Kosintsev, Pavel; Jahren, A. Hope; Shute, Elen; Zazula, Grant D.; Wooller, Matthew J.; Rabanus-Wallace, M. Timothy (May 2017). "Megafaunal isotopes reveal role of increased moisture on rangeland during late Pleistocene extinctions". Nature Ecology & Evolution. 1 (5): 4. doi:10.1038/s41559-017-0125. ISSN 2397-334X. PMID 28812683.
  205. ^ Mouline, Karine; Granjon, Laurent; Galan, Maxime; Tatard, Caroline; Abdoullaye, Doukary; Atteyine, Solimane Ag; Duplantier, Jean-Marc; Cosson, Jean-François (2008). "Phylogeography of a Sahelian rodent species Mastomys huberti: a Plio-Pleistocene story of emergence and colonization of humid habitats". Molecular Ecology. 17 (4): 1036–1053. doi:10.1111/j.1365-294X.2007.03610.x. ISSN 1365-294X. PMID 18261047.
  206. ^ a b Duringer, Philippe; Marsaleix, Patrick; Moussa, Abderamane; Roquin, Claude; Denamiel, Cléa; Ghienne, Jean-François; Schuster, Mathieu; Bouchette, Frédéric (2010). "Hydrodynamics in Holocene Lake Mega-Chad". Quaternary Research. 73 (2): 226. Bibcode:2010QuRes..73..226B. doi:10.1016/j.yqres.2009.10.010. ISSN 1096-0287.
  207. ^ Bard 2013, p. 809.
  208. ^ Lézine, Duplessy & Cazet 2005, p. 234.
  209. ^ a b Sylvestre et al. 2013, pp. 232–233.
  210. ^ Armitage, Bristow & Drake 2015, p. 8544.
  211. ^ a b Drake & Bristow 2006, p. 906.
  212. ^ Sepulchre et al. 2008, p. 43.
  213. ^ Pachur & Altmann 2006, p. 26.
  214. ^ Pachur & Altmann 2006, p. 23.
  215. ^ Runge 2010, p. 239.
  216. ^ a b Runge 2010, p. 238.
  217. ^ Quade et al. 2018, p. 19.
  218. ^ Pachur & Altmann 2006, p. 83.
  219. ^ Vahrenholt & Lüning 2019, pp. 518–519.
  220. ^ Petit-Maire 1989, p. 645.
  221. ^ Wendorf, Karlén & Schild 2007, p. 196.
  222. ^ a b Hillaire-Marcel, Claude; Casanova, Joël; Lézine, Anne-Marie (1 March 1990). "Across an early Holocene humid phase in western Sahara:Pollen and isotope stratigraphy". Geology. 18 (3): 264. Bibcode:1990Geo....18..264L. doi:10.1130/0091-7613(1990)018<0264:AAEHHP>2.3.CO;2. ISSN 0091-7613.
  223. ^ Gasse 2000, p. 204.
  224. ^ a b Gasse & Van Campo 1994, p. 447.
  225. ^ a b Pachur & Altmann 2006, p. 246.
  226. ^ Jahns 1995, p. 23.
  227. ^ Wendorf, Karlén & Schild 2007, p. 206.
  228. ^ Wendorf, Karlén & Schild 2007, pp. 206–207.
  229. ^ Wendorf, Karlén & Schild 2007, p. 216.
  230. ^ a b Bubenzer, Olaf; Bolten, Andreas (December 2008). "The use of new elevation data (SRTM/ASTER) for the detection and morphometric quantification of Pleistocene megadunes (draa) in the eastern Sahara and the southern Namib". Geomorphology. 102 (2): 225. Bibcode:2008Geomo.102..221B. doi:10.1016/j.geomorph.2008.05.003. ISSN 0169-555X.
  231. ^ Pachur & Altmann 2006, p. 80.
  232. ^ a b Franz, Gerhard; Breitkreuz, Christoph; Coyle, David A.; El Hur, Bushra; Heinrich, Wilhelm; Paulick, Holger; Pudlo, Dieter; Smith, Robyn; Steiner, Gesine (August 1997). "The alkaline Meidob volcanic field (Late Cenozoic, northwest Sudan)". Journal of African Earth Sciences. 25 (2): 7. Bibcode:1997JAfES..25..263F. doi:10.1016/S0899-5362(97)00103-6. ISSN 1464-343X.
  233. ^ Wendorf, Karlén & Schild 2007, p. 204.
  234. ^ a b Lenhardt, Nils; Borah, Suranjana B.; Lenhardt, Sukanya Z.; Bumby, Adam J.; Ibinoof, Montasir A.; Salih, Salih A. (May 2018). "The monogenetic Bayuda Volcanic Field, Sudan – New insights into geology and volcanic morphology". Journal of Volcanology and Geothermal Research. 356: 222. Bibcode:2018JVGR..356..211L. doi:10.1016/j.jvolgeores.2018.03.010. ISSN 0377-0273.
  235. ^ Wu et al. 2017, p. 106.
  236. ^ a b c White et al. 2011, p. 459.
  237. ^ Quade et al. 2018, p. 18.
  238. ^ a b c Wu et al. 2017, p. 96.
  239. ^ Perego, Zerboni & Cremaschi 2011, p. 472.
  240. ^ a b Zerboni & Gatto 2015, p. 309.
  241. ^ Maley 2000, p. 125.
  242. ^ Drake & Bristow 2006, p. 909.
  243. ^ Lecomte, Frédéric; Dodson, Julian J.; Guinand, Bruno; Durand, Jean-Dominique (9 October 2013). "Pelagic Life and Depth: Coastal Physical Features in West Africa Shape the Genetic Structure of the Bonga Shad, Ethmalosa fimbriata". PLOS ONE. 8 (10): 2. Bibcode:2013PLoSO...877483D. doi:10.1371/journal.pone.0077483. ISSN 1932-6203. PMC 3793960. PMID 24130890.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  244. ^ Armitage, S.J.; Pinder, R.C. (April 2017). "Testing the applicability of optically stimulated luminescence dating to Ocean Drilling Program cores". Quaternary Geochronology. 39: 125. doi:10.1016/j.quageo.2017.02.008. ISSN 1871-1014.
  245. ^ Lewin, John; Ashworth, Philip J.; Strick, Robert J. P. (February 2017). "Spillage sedimentation on large river floodplains: Spillage sedimentation on large river floodplains". Earth Surface Processes and Landforms. 42 (2): 301. doi:10.1002/esp.3996.
  246. ^ Ramos, Ramil & Sanz 2017, p. 95.
  247. ^ Ramos, Ramil & Sanz 2017, p. 101.
  248. ^ a b c d e Maslin, Manning & Brierley 2018, p. 1.
  249. ^ a b c Lernia et al. 2017, p. 1.
  250. ^ Riemer 2006, p. 555.
  251. ^ a b Stojanowski, Carver & Miller 2014, pp. 80–82.
  252. ^ Mercuri, Anna Maria; Sadori, Laura (2014), Goffredo, Stefano; Dubinsky, Zvy (eds.), "Mediterranean Culture and Climatic Change: Past Patterns and Future Trends", The Mediterranean Sea, Springer Netherlands, p. 519, doi:10.1007/978-94-007-6704-1_30, ISBN 9789400767034, retrieved 5 October 2019
  253. ^ a b Cremaschi et al. 2010, p. 88.
  254. ^ a b Cremaschi et al. 2010, p. 91.
  255. ^ Lernia et al. 2013, p. 122.
  256. ^ Chiotis 2018, p. 16.
  257. ^ Hoelzmann et al. 2001, p. 210.
  258. ^ a b c d e Smith 2018, p. 243.
  259. ^ a b Badino, Federica; Ravazzi, Cesare; Vallè, Francesca; Pini, Roberta; Aceti, Amelia; Brunetti, Michele; Champvillair, Elena; Maggi, Valter; Maspero, Francesco; Perego, Renata; Orombelli, Giuseppe (April 2018). "8800 years of high-altitude vegetation and climate history at the Rutor Glacier forefield, Italian Alps. Evidence of middle Holocene timberline rise and glacier contraction". Quaternary Science Reviews. 185: 41. Bibcode:2018QSRv..185...41B. doi:10.1016/j.quascirev.2018.01.022. ISSN 0277-3791.
  260. ^ Phillipps et al. 2012, p. 71.
  261. ^ White et al. 2011, pp. 460–461.
  262. ^ Tafuri et al. 2006, p. 390.
  263. ^ a b c Brooks et al. 2007, p. 260.
  264. ^ Riemer 2006, p. 556.
  265. ^ Lernia et al. 2012, pp. 391–392.
  266. ^ Lernia et al. 2013, p. 121.
  267. ^ Breunig, Neumann & Van Neer 1996, p. 116.
  268. ^ Breunig, Neumann & Van Neer 1996, p. 117.
  269. ^ Lernia et al. 2013, pp. 123–124.
  270. ^ Stojanowski, Christopher M. (30 November 2018), "Persistence or Pastoralism: The Challenges of Studying Hunter-Gatherer Resilience in Africa", in Temple, Daniel H.; Stojanowski, Christopher M. (eds.), Hunter-Gatherer Adaptation and Resilience (1 ed.), Cambridge University Press, p. 195, doi:10.1017/9781316941256.009, ISBN 9781316941256, retrieved 22 July 2019
  271. ^ Lézine 2017, p. 3.
  272. ^ Lernia et al. 2017, p. 5.
  273. ^ Marinova, Meckler & McKay 2014, p. 50.
  274. ^ a b Lernia et al. 2012, p. 390.
  275. ^ Marinova, Meckler & McKay 2014, p. 54.
  276. ^ Olsen 2017, p. 107.
  277. ^ Olsen 2017, p. 93.
  278. ^ Pachur & Altmann 2006, p. 533.
  279. ^ Soriano et al. 2009, p. 8.
  280. ^ a b c Cremaschi & Zerboni 2009, p. 690.
  281. ^ a b Gasse 2000, p. 189.
  282. ^ Pirie et al. 2009, p. 930.
  283. ^ Brooks et al. 2007, p. 259.
  284. ^ Calderón, Rosario; Pereira, Luisa; Baali, Abdellatif; Melhaoui, Mohammed; Oliveira, Marisa; Rito, Teresa; Rodríguez, Juan N.; Novelletto, Andrea; Dugoujon, Jean M.; Soares, Pedro; Hernández, Candela L. (28 October 2015). "Early Holocenic and Historic mtDNA African Signatures in the Iberian Peninsula: The Andalusian Region as a Paradigm". PLOS ONE. 10 (10): 16. Bibcode:2015PLoSO..1039784H. doi:10.1371/journal.pone.0139784. ISSN 1932-6203. PMC 4624789. PMID 26509580.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  285. ^ Haber, Marc; Mezzavilla, Massimo; Bergström, Anders; Prado-Martinez, Javier; Hallast, Pille; Saif-Ali, Riyadh; Al-Habori, Molham; Dedoussis, George; Zeggini, Eleftheria; Blue-Smith, Jason; Wells, R. Spencer; Xue, Yali; Zalloua, Pierre A.; Tyler-Smith, Chris (1 December 2016). "Chad Genetic Diversity Reveals an African History Marked by Multiple Holocene Eurasian Migrations". The American Journal of Human Genetics. 99 (6): 1316–1324. doi:10.1016/j.ajhg.2016.10.012. ISSN 0002-9297. PMC 5142112. PMID 27889059.
  286. ^ a b c Blümel 2002, p. 12.
  287. ^ Zerboni, Trombino & Cremaschi 2011, p. 321.
  288. ^ a b Zerboni, Trombino & Cremaschi 2011, p. 332.
  289. ^ Sponholz, Baumhauer & Felix-Henningsen 2016, pp. 97–98.
  290. ^ Baumhauer 2004, p. 296.
  291. ^ Sponholz, Baumhauer & Felix-Henningsen 2016, p. 103.
  292. ^ Perego, Zerboni & Cremaschi 2011, p. 466.
  293. ^ Eggermont et al. 2008, p. 2411.
  294. ^ Cremaschi et al. 2010, p. 87.
  295. ^ Pachur & Altmann 2006, p. 153.
  296. ^ a b Pachur & Altmann 2006, p. 2.
  297. ^ Hély et al. 2009, p. 680.
  298. ^ Goudie, Andrew S.; Middleton, Nicholas J. (2006), "Quaternary Dust Loadings", Desert Dust in the Global System, Springer Berlin Heidelberg, p. 202, doi:10.1007/3-540-32355-4_9, ISBN 9783540323549, retrieved 5 October 2019
  299. ^ Muhs et al. 2013, p. 43.
  300. ^ a b Kohn, Marion; Steinke, Stephan; Baumann, Karl-Heinz; Donner, Barbara; Meggers, Helge; Zonneveld, Karin A.F. (March 2011). "Stable oxygen isotopes from the calcareous-walled dinoflagellate Thoracosphaera heimii as a proxy for changes in mixed layer temperatures off NW Africa during the last 45,000yr". Palaeogeography, Palaeoclimatology, Palaeoecology. 302 (3–4): 319. Bibcode:2011PPP...302..311K. doi:10.1016/j.palaeo.2011.01.019. ISSN 0031-0182.
  301. ^ a b Zarriess, Michelle; Mackensen, Andreas (September 2010). "The tropical rainbelt and productivity changes off northwest Africa: A 31,000-year high-resolution record". Marine Micropaleontology. 76 (3–4): 87. Bibcode:2010MarMP..76...76Z. doi:10.1016/j.marmicro.2010.06.001. ISSN 0377-8398.
  302. ^ a b Haslett, Simon K.; Smart, Christopher W. (2006). "Late Quaternary upwelling off tropical NW Africa: new micropalaeontological evidence from ODP Hole 658C". Journal of Quaternary Science. 21 (3): 267. Bibcode:2006JQS....21..259H. doi:10.1002/jqs.970. ISSN 1099-1417.
  303. ^ a b Haslett, Simon K; Davies, Catherine F C (1 March 2006). "Late Quaternary climate–ocean changes in western North Africa: offshore geochemical evidence". Transactions of the Institute of British Geographers. 31 (1): 37. doi:10.1111/j.1475-5661.2006.00193.x. ISSN 0020-2754.
  304. ^ Matter et al. 2016, p. 88.
  305. ^ Radies et al. 2005, p. 111.
  306. ^ a b Damme, Kay Van; Benda, Petr; Damme, Dirk Van; Geest, Peter De; Hajdas, Irka (26 August 2018). "The first vertebrate fossil from Socotra Island (Yemen) is an early Holocene Egyptian fruit bat". Journal of Natural History. 52 (31–32): 2017. doi:10.1080/00222933.2018.1510996. ISSN 0022-2933 – via Academia.edu.
  307. ^ Vahrenholt & Lüning 2019, p. 524.
  308. ^ a b Radies et al. 2005, p. 122.
  309. ^ Vahrenholt & Lüning 2019, p. 527.
  310. ^ Matter et al. 2016, p. 99.
  311. ^ Kennett & Kennett 2007, p. 236.
  312. ^ Petraglia & Rose 2010, p. 28.
  313. ^ Matter et al. 2016, p. 89.
  314. ^ Petraglia & Rose 2010, p. 219.
  315. ^ Vahrenholt & Lüning 2019, pp. 525–527.
  316. ^ a b c Lézine et al. 2010, p. 427.
  317. ^ Renaud et al. 2010, p. 230.
  318. ^ Kennett & Kennett 2007, p. 237.
  319. ^ Matter et al. 2016, p. 98.
  320. ^ Lézine et al. 2010, p. 426.
  321. ^ Prasad & Negendank 2004, p. 213.
  322. ^ Renaud et al. 2010, p. 228.
  323. ^ Matter et al. 2016, p. 89,98.
  324. ^ Gasse, Françoise (January 2005). "Continental palaeohydrology and palaeoclimate during the Holocene". Comptes Rendus Geoscience. 337 (1–2): 81. Bibcode:2005CRGeo.337...79G. doi:10.1016/j.crte.2004.10.006. ISSN 1631-0713.
  325. ^ Mercuri et al. 2018, p. 226.
  326. ^ a b c Morrissey & Scholz 2014, p. 98.
  327. ^ Graham, Angus; Strutt, Kristian D.; Peeters, Jan; Toonen, Willem H. J.; Pennington, Benjamin T.; Emery, Virginia L.; Barker, Dominic S.; Johansson, Carolin (30 June 2017). "Theban Harbours and Waterscapes Survey, Spring 2016". The Journal of Egyptian Archaeology. 102 (1): 19. doi:10.1177/030751331610200103.
  328. ^ Wendorf, Karlén & Schild 2007, p. 205.
  329. ^ Hoelzmann et al. 2001, p. 212.
  330. ^ a b Morrissey & Scholz 2014, p. 96.
  331. ^ Blanchet et al. 2013, p. 105.
  332. ^ Garcin et al. 2017, p. 60.
  333. ^ Junginger et al. 2014, p. 2.
  334. ^ a b van der Lubbe et al. 2017, p. 8.
  335. ^ a b Bloszies, Forman & Wright 2015, p. 66.
  336. ^ a b van der Lubbe et al. 2017, p. 3.
  337. ^ Smith 2018, p. 249.
  338. ^ a b Hoelzmann & Holmes 2017, p. 12.
  339. ^ Gasse & Van Campo 1994, p. 445.
  340. ^ Loakes, Katie (2 January 2017). "Late Quaternary palaeolimnology and environmental change in the South Wollo Highlands". Azania: Archaeological Research in Africa. 52 (1): 131. doi:10.1080/0067270X.2016.1259821. ISSN 0067-270X.
  341. ^ Hoelzmann & Holmes 2017, p. 17.
  342. ^ Runge, Jürgen (12 October 2017). Runge, Jürgen; Eisenberg, Joachim (eds.). The African Neogene – Climate, Environments and People (1 ed.). CRC Press. p. 145. doi:10.1201/9781315161808. ISBN 9781315161808.
  343. ^ Beer et al. 2002, p. 593.
  344. ^ Gabrielli, P.; Hardy, D.R.; Kehrwald, N.; Davis, M.; Cozzi, G.; Turetta, C.; Barbante, C.; Thompson, L.G. (June 2014). "Deglaciated areas of Kilimanjaro as a source of volcanic trace elements deposited on the ice cap during the late Holocene". Quaternary Science Reviews. 93: 3. Bibcode:2014QSRv...93....1G. doi:10.1016/j.quascirev.2014.03.007. ISSN 0277-3791.
  345. ^ Zech, Michael (December 2006). "Evidence for Late Pleistocene climate changes from buried soils on the southern slopes of Mt. Kilimanjaro, Tanzania". Palaeogeography, Palaeoclimatology, Palaeoecology. 242 (3–4): 310. Bibcode:2006PPP...242..303Z. doi:10.1016/j.palaeo.2006.06.008. ISSN 0031-0182.
  346. ^ Kervyn, M.; Macheyeki, A.; Kwelwa, S.; Delvaux, D.; Delcamp, A. (1 January 2016). "Sector collapse events at volcanoes in the North Tanzanian divergence zone and their implications for regional tectonics". GSA Bulletin. 128 (1–2): 15. Bibcode:2013GSABu.125..299B. doi:10.1130/B31119.1. ISSN 0016-7606.
  347. ^ Garcin et al. 2017, p. 67.
  348. ^ Garcin et al. 2017, p. 68.
  349. ^ a b Barker et al. 2002, p. 303.
  350. ^ a b Russell & Ivory 2018, p. 7.
  351. ^ a b Russell & Ivory 2018, p. 8.
  352. ^ Jahns 1995, p. 28.
  353. ^ a b c Russell & Ivory 2018, p. 9.
  354. ^ Tierney et al. 2011, p. 106.
  355. ^ a b Russell & Ivory 2018, p. 12.
  356. ^ a b Junginger & Trauth 2013, p. 186.
  357. ^ Junginger & Trauth 2013, p. 174.
  358. ^ White et al. 2011, p. 461.
  359. ^ Müller, Ulrich C.; Pross, Jörg; Tzedakis, Polychronis C.; Gamble, Clive; Kotthoff, Ulrich; Schmiedl, Gerhard; Wulf, Sabine; Christanis, Kimon (February 2011). "The role of climate in the spread of modern humans into Europe". Quaternary Science Reviews. 30 (3–4): 273–279. Bibcode:2011QSRv...30..273M. doi:10.1016/j.quascirev.2010.11.016. ISSN 0277-3791.
  360. ^ Nguetsop, Victor François; Bentaleb, Ilham; Favier, Charly; Bietrix, Sophie; Martin, Céline; Servant-Vildary, Simone; Servant, Michel (July 2013). "A late Holocene palaeoenvironmental record from Lake Tizong, northern Cameroon using diatom and carbon stable isotope analyses". Quaternary Science Reviews. 72: 50. Bibcode:2013QSRv...72...49N. doi:10.1016/j.quascirev.2013.04.005. ISSN 0277-3791.
  361. ^ La Roche, Francisco; Genise, Jorge F.; Castillo, Carolina; Quesada, María Luisa; García-Gotera, Cristo M.; De la Nuez, Julio (September 2014). "Fossil bee cells from the Canary Islands. Ichnotaxonomy, palaeobiology and palaeoenvironments of Palmiraichnus castellanosi". Palaeogeography, Palaeoclimatology, Palaeoecology. 409: 262. Bibcode:2014PPP...409..249L. doi:10.1016/j.palaeo.2014.05.012. ISSN 0031-0182. {{cite journal}}: Invalid |ref=harv (help)
  362. ^ Hély et al. 2009, p. 683.
  363. ^ Tropical rainforest responses to climatic change. Environmental Sciences (2nd ed.). Springer. 2011. p. 166. ISBN 978-3-642-05383-2.
  364. ^ Ifo, Suspense A.; Bocko, Yannick E.; Page, Susan E.; Mitchard, Edward T. A.; Lawson, Ian T.; Lewis, Simon L.; Dargie, Greta C. (February 2017). "Age, extent and carbon storage of the central Congo Basin peatland complex". Nature. 542 (7639): 86–90. Bibcode:2017Natur.542...86D. doi:10.1038/nature21048. ISSN 1476-4687. PMID 28077869.
  365. ^ Zielhofer, Christoph; Faust, Dominik (March 2008). "Mid- and Late Holocene fluvial chronology of Tunisia". Quaternary Science Reviews. 27 (5–6): 586. Bibcode:2008QSRv...27..580Z. doi:10.1016/j.quascirev.2007.11.019. ISSN 0277-3791.
  366. ^ Stoetzel, Emmanuelle (1 December 2017). "Adaptations and Dispersals of Anatomically Modern Humans in the Changing Environments of North Africa: the Contribution of Microvertebrates". African Archaeological Review. 34 (4): 9. doi:10.1007/s10437-017-9272-0. ISSN 1572-9842.
  367. ^ a b c Zielhofer et al. 2016, p. 858.
  368. ^ Yanes, Yurena; Romanek, Christopher S.; Molina, Fernando; Cámara, Juan Antonio; Delgado, Antonio (November 2011). "Holocene paleoenvironment (∼7200–4000 cal BP) of the Los Castillejos archaeological site (SE Spain) inferred from the stable isotopes of land snail shells". Quaternary International. 244 (1): 73–74. Bibcode:2011QuInt.244...67Y. doi:10.1016/j.quaint.2011.04.031. ISSN 1040-6182.
  369. ^ Censi, P.; Incarbona, A.; Oliveri, E.; Bonomo, S.; Tranchida, G. (June 2010). "Yttrium and REE signature recognized in Central Mediterranean Sea (ODP Site 963) during the MIS 6–MIS 5 transition". Palaeogeography, Palaeoclimatology, Palaeoecology. 292 (1–2): 206. Bibcode:2010PPP...292..201C. doi:10.1016/j.palaeo.2010.03.045. ISSN 0031-0182.
  370. ^ a b Spötl, Christoph; Nicolussi, Kurt; Patzelt, Gernot; Boch, Ronny (April 2010). "Humid climate during deposition of sapropel 1 in the Mediterranean Sea: Assessing the influence on the Alps". Global and Planetary Change. 71 (3–4): 242. Bibcode:2010GPC....71..242S. doi:10.1016/j.gloplacha.2009.10.003. ISSN 0921-8181.
  371. ^ a b Sbaffi, Laura; Wezel, Forese Carlo; Curzi, Giuseppe; Zoppi, Ugo (January 2004). "Millennial- to centennial-scale palaeoclimatic variations during Termination I and the Holocene in the central Mediterranean Sea". Global and Planetary Change. 40 (1–2): 203. Bibcode:2004GPC....40..201S. doi:10.1016/S0921-8181(03)00111-5. ISSN 0921-8181.
  372. ^ Incarbona, Alessandro; Zarcone, Giuseppe; Agate, Mauro; Bonomo, Sergio; Stefano, Enrico; Masini, Federico; Russo, Fabio; Sineo, Luca (2010). "A multidisciplinary approach to reveal the Sicily Climate and Environment over the last 20 000 years". Open Geosciences. 2 (2): 71. Bibcode:2010CEJG....2...71I. doi:10.2478/v10085-010-0005-8. ISSN 2391-5447.
  373. ^ Hamann et al. 2017, p. 453.
  374. ^ Williams et al. 2010, p. 1117.
  375. ^ a b Hamann et al. 2017, p. 461.
  376. ^ a b Fontaine, M.C. (1 January 2016). Harbour Porpoises, Phocoena phocoena, in the Mediterranean Sea and Adjacent Regions: Biogeographic Relicts of the Last Glacial Period. Vol. 75. pp. 333–358. doi:10.1016/bs.amb.2016.08.006. ISBN 9780128051528. ISSN 0065-2881. PMID 27770989. {{cite book}}: |journal= ignored (help)
  377. ^ Vahrenholt & Lüning 2019, p. 522.
  378. ^ Kiro, Yael; Goldstein, Steven L.; Garcia-Veigas, Javier; Levy, Elan; Kushnir, Yochanan; Stein, Mordechai; Lazar, Boaz (April 2017). "Relationships between lake-level changes and water and salt budgets in the Dead Sea during extreme aridities in the Eastern Mediterranean". Earth and Planetary Science Letters. 464: 221. Bibcode:2017E&PSL.464..211K. doi:10.1016/j.epsl.2017.01.043. ISSN 0012-821X.
  379. ^ a b Sletten, Hillary R.; Railsback, L. Bruce; Liang, Fuyuan; Brook, George A.; Marais, Eugene; Hardt, Benjamin F.; Cheng, Hai; Edwards, R. Lawrence (April 2013). "A petrographic and geochemical record of climate change over the last 4600years from a northern Namibia stalagmite, with evidence of abruptly wetter climate at the beginning of southern Africa's Iron Age". Palaeogeography, Palaeoclimatology, Palaeoecology. 376: 158. Bibcode:2013PPP...376..149S. doi:10.1016/j.palaeo.2013.02.030. ISSN 0031-0182.
  380. ^ Reimer et al. 2010, p. 40.
  381. ^ Ramisch, Arne; Bens, Oliver; Buylaert, Jan-Pieter; Eden, Marie; Heine, Klaus; Hürkamp, Kerstin; Schwindt, Daniel; Völkel, Jörg (March 2017). "Fluvial landscape development in the southwestern Kalahari during the Holocene – Chronology and provenance of fluvial deposits in the Molopo Canyon". Geomorphology. 281: 104. Bibcode:2017Geomo.281...94R. doi:10.1016/j.geomorph.2016.12.021. ISSN 0169-555X.
  382. ^ Fitchett, Jennifer M.; Grab, Stefan W.; Bamford, Marion K.; Mackay, Anson W. (2 September 2017). "Late Quaternary research in southern Africa: progress, challenges and future trajectories". Transactions of the Royal Society of South Africa. 72 (3): 284. doi:10.1080/0035919X.2017.1297966. ISSN 0035-919X.
  383. ^ Burrough & Thomas 2013, p. 43.
  384. ^ Lubbe, H. J. L. van der; Frank, Martin; Tjallingii, Rik; Schneider, Ralph R. (2016). "Neodymium isotope constraints on provenance, dispersal, and climate-driven supply of Zambezi sediments along the Mozambique Margin during the past ∼45,000 years" (PDF). Geochemistry, Geophysics, Geosystems. 17 (1): 195. Bibcode:2016GGG....17..181V. doi:10.1002/2015GC006080. ISSN 1525-2027.
  385. ^ Battarbee, Gasse & Stickley 2004, p. 572.
  386. ^ Quade et al. 2018, p. 16.
  387. ^ Dixit et al. 2018, p. 233.
  388. ^ Lézine, Duplessy & Cazet 2005, pp. 226–227.
  389. ^ Pachur & Altmann 2006, p. 564.
  390. ^ Hoelzmann & Holmes 2017, p. 15.
  391. ^ Hoelzmann & Holmes 2017, pp. 16–18.
  392. ^ Junginger & Trauth 2013, p. 178.
  393. ^ a b Baumhauer & Runge 2009, p. 29.
  394. ^ Baumhauer & Runge 2009, p. 11.
  395. ^ a b Engel et al. 2012, p. 139.
  396. ^ Radies et al. 2005, p. 123.
  397. ^ a b Vahrenholt & Lüning 2019, p. 507.
  398. ^ a b Greer, Lisa; Swart, Peter K. (2006). "Decadal cyclicity of regional mid-Holocene precipitation: Evidence from Dominican coral proxies". Paleoceanography. 21 (2): 2. Bibcode:2006PalOc..21.2020G. doi:10.1029/2005PA001166. ISSN 1944-9186.
  399. ^ Russell & Ivory 2018, p. 2.
  400. ^ Rolandone, F.; Lucazeau, F. (20 August 2012). "Heat-flow and subsurface temperature history at the site of Saraya (eastern Senegal)". Solid Earth. 3 (2): 216. Bibcode:2012SolE....3..213L. doi:10.5194/se-3-213-2012. ISSN 1869-9510.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  401. ^ Donnelly et al. 2017, p. 6223.
  402. ^ Gaetani et al. 2017, p. 7639.
  403. ^ a b Donnelly et al. 2017, p. 6225.
  404. ^ Toomey et al. 2013, p. 31.
  405. ^ a b Gaetani et al. 2017, p. 7640.
  406. ^ Donnelly et al. 2017, p. 6224.
  407. ^ Toomey et al. 2013, p. 39.
  408. ^ Liu et al. 2017, p. 2.
  409. ^ Liu et al. 2017, p. 3.
  410. ^ Liu et al. 2017, p. 9.
  411. ^ a b Niedermeyer et al. 2010, p. 3003.
  412. ^ Menocal et al. 2000, pp. 354–355.
  413. ^ Cohen et al. 2008, p. 252.
  414. ^ a b Junginger et al. 2014, p. 14.
  415. ^ Zielhofer et al. 2016, p. 857.
  416. ^ Muhs et al. 2013, p. 34.
  417. ^ Wendorf, Karlén & Schild 2007, p. 191.
  418. ^ a b Bloszies, Forman & Wright 2015, p. 65.
  419. ^ Talbot et al. 2007, pp. 9–10.
  420. ^ Talbot et al. 2007, p. 10.
  421. ^ Morrill, Overpeck & Cole 2016, p. 469.
  422. ^ Zerboni & Gatto 2015, p. 310.
  423. ^ a b c d Menocal et al. 2000, p. 355.
  424. ^ Zielhofer et al. 2016, p. 851.
  425. ^ Cole et al. 2009, p. 258.
  426. ^ Lubell, David; Jackes, Mary (1 June 2008). "Early and Middle Holocene Environments and Capsian Cultural Change: Evidence from the Télidjène Basin, Eastern Algeria". African Archaeological Review. 25 (1–2): 53. CiteSeerX 10.1.1.518.2283. doi:10.1007/s10437-008-9024-2. ISSN 1572-9842.
  427. ^ Stivers et al. 2008, p. 1.
  428. ^ a b Blanchet et al. 2013, p. 108.
  429. ^ a b Peck et al. 2015, p. 141.
  430. ^ Garcin, Yannick; Vincens, Annie; Williamson, David; Guiot, Joël; Buchet, Guillaume (2006). "Wet phases in tropical southern Africa during the last glacial period". Geophysical Research Letters. 33 (7): 3. Bibcode:2006GeoRL..33.7703G. doi:10.1029/2005GL025531. ISSN 1944-8007.
  431. ^ a b c Zielhofer et al. 2017, p. 131.
  432. ^ Lézine, Duplessy & Cazet 2005, p. 236.
  433. ^ Schuster & Nutz 2016, p. 1615.
  434. ^ Junginger et al. 2014, pp. 98–99.
  435. ^ Schuster & Nutz 2016, pp. 1614–1615.
  436. ^ Sylvestre et al. 2013, p. 237.
  437. ^ Wendorf, Karlén & Schild 2007, p. 197.
  438. ^ Wendorf, Karlén & Schild 2007, p. 203.
  439. ^ Said 1993, p. 131.
  440. ^ Chiotis 2018, p. 18.
  441. ^ Zerboni & Gatto 2015, p. 312.
  442. ^ Huang et al. 2008, p. 1460.
  443. ^ Krüger et al. 2017, p. 10.
  444. ^ Armitage, Bristow & Drake 2015, p. 8547.
  445. ^ Sylvestre et al. 2013, p. 223.
  446. ^ Nogué, Sandra; Nascimento, Lea de; Fernández‐Palacios, José María; Whittaker, Robert J.; Willis, Kathy J. (2013). "The ancient forests of La Gomera, Canary Islands, and their sensitivity to environmental change". Journal of Ecology. 101 (2): 374. Bibcode:2009JEcol..98...74B. doi:10.1111/1365-2745.12051. ISSN 1365-2745.
  447. ^ Blümel 2002, p. 11.
  448. ^ a b Magny & Haas 2004, p. 425.
  449. ^ Marsicek et al. 2013, p. 140.
  450. ^ Mooney, Scott D.; Black, Manu P. (1 March 2006). "Holocene fire history from the Greater Blue Mountains World Heritage Area, New South Wales, Australia: the climate, humans and fire nexus". Regional Environmental Change. 6 (1–2): 48–49. Bibcode:2013REC..2013....1J. doi:10.1007/s10113-005-0003-8. ISSN 1436-378X.
  451. ^ a b Lebamba et al. 2016, p. 130.
  452. ^ Beer et al. 2002, p. 592.
  453. ^ Wendorf, Karlén & Schild 2007, p. 201.
  454. ^ Liu et al. 2014, p. 2024.
  455. ^ Jung et al. 2004, p. 35.
  456. ^ Chalié & Roubeix 2018, p. 3.
  457. ^ a b c Claussen et al. 1999, p. 2037.
  458. ^ Jung et al. 2004, pp. 34–35.
  459. ^ Zielhofer et al. 2017, p. 120.
  460. ^ Hély et al. 2009, p. 673.
  461. ^ Chalié & Roubeix 2018, pp. 11–12.
  462. ^ Bristow et al. 2018, p. 194.
  463. ^ a b Schefuß et al. 2017, p. 6.
  464. ^ Bristow et al. 2018, p. 186.
  465. ^ Hoelzmann & Holmes 2017, pp. 26–27.
  466. ^ Drake & Bristow 2006, p. 908.
  467. ^ a b Mercuri et al. 2018, p. 222.
  468. ^ a b c d e Lézine 2009, p. 751.
  469. ^ Petit-Maire 1989, p. 649.
  470. ^ Zerboni, Andrea; Mori, Lucia; Bosi, Giovanna; Buldrini, Fabrizio; Bernasconi, Andrea; Gatto, Maria Carmela; Mercuri, Anna Maria (September 2017). "Domestic firing activities and fuel consumption in a Saharan oasis: Micromorphological and archaeobotanical evidence from the Garamantian site of Fewet (Central Sahara, SW Libya)". Journal of Arid Environments. 144: 124. Bibcode:2017JArEn.144..123Z. doi:10.1016/j.jaridenv.2017.03.012. hdl:11380/1135660. ISSN 0140-1963.
  471. ^ Pachur & Altmann 2006, p. 34.
  472. ^ Eggermont et al. 2008, p. 2423.
  473. ^ Lézine 2009, p. 753.
  474. ^ a b Cole et al. 2009, p. 264.
  475. ^ Krinner et al. 2012, p. 2.
  476. ^ Chalié & Roubeix 2018, p. 13.
  477. ^ Olsen 2017, p. 91.
  478. ^ Kennett & Kennett 2007, p. 240.
  479. ^ a b Russell & Ivory 2018, p. 10.
  480. ^ Junginger et al. 2014, pp. 14–15.
  481. ^ Schaebitz, F.; Trauth, M. H.; Brown, M. C.; Frank, U.; Rethemeyer, J.; Weber, M.; Lamb, H. F.; Asrat, A.; Junginger, A.; Foerster, V. (7 March 2014). "46 000 years of alternating wet and dry phases on decadal to orbital timescales in the cradle of modern humans: the Chew Bahir project, southern Ethiopia". Climate of the Past Discussions. 10 (2): 980. Bibcode:2014CliPD..10..977F. doi:10.5194/cpd-10-977-2014. ISSN 1814-9324.{{cite journal}}: CS1 maint: unflagged free DOI (link)
  482. ^ van der Lubbe et al. 2017, p. 1.
  483. ^ Berke et al. 2012, p. 99.
  484. ^ a b Berke et al. 2012, p. 100.
  485. ^ a b Berke et al. 2012, p. 103.
  486. ^ Morrissey & Scholz 2014, p. 89.
  487. ^ Morrissey & Scholz 2014, p. 99.
  488. ^ Costas, Susana; Jerez, Sonia; Trigo, Ricardo M.; Goble, Ronald; Rebêlo, Luís (May 2012). "Sand invasion along the Portuguese coast forced by westerly shifts during cold climate events" (PDF). Quaternary Science Reviews. 42: 24. Bibcode:2012QSRv...42...15C. doi:10.1016/j.quascirev.2012.03.008. hdl:10400.9/1848. ISSN 0277-3791.
  489. ^ Zielhofer et al. 2017, p. 132.
  490. ^ Hoelzmann & Holmes 2017, p. 26.
  491. ^ a b Sangen 2012, p. 215.
  492. ^ Pirie et al. 2009, p. 924.
  493. ^ Servant, Buchet & Vincens 2010, p. 291.
  494. ^ a b Lebamba et al. 2016, p. 136.
  495. ^ Niedermeyer et al. 2010, p. 3002.
  496. ^ a b c Lézine et al. 2013, p. 329.
  497. ^ a b Lézine et al. 2013, p. 328.
  498. ^ Lézine 2017, p. 20.
  499. ^ Hipondoka, M.H.T.; Mauz, B.; Kempf, J.; Packman, S.; Chiverrell, R.C.; Bloemendal, J. (January 2014). "Chronology of sand ridges and the Late Quaternary evolution of the Etosha Pan, Namibia". Geomorphology. 204: 561–562. Bibcode:2014Geomo.204..553H. doi:10.1016/j.geomorph.2013.08.034. ISSN 0169-555X.
  500. ^ a b Forman, Wright & Bloszies 2014, p. 85.
  501. ^ a b c Meeker, L. David; Cumming, Brian F.; Stager, J. Curt (2003). "A 10,000-year high-resolution diatom record from Pilkington Bay, Lake Victoria, East Africa". Quaternary Research. 59 (2): 180. Bibcode:2003QuRes..59..172S. doi:10.1016/S0033-5894(03)00008-5. ISSN 1096-0287.
  502. ^ Krinner et al. 2012, pp. 1–2.
  503. ^ a b Servant, Buchet & Vincens 2010, p. 282.
  504. ^ Brooks et al. 2007, p. 257.
  505. ^ Ganopolski et al. 2009, p. 466.
  506. ^ Ganopolski et al. 2009, p. 458.
  507. ^ a b c Menocal 2015, p. 2.
  508. ^ Guilderson et al. 2001, p. 197.
  509. ^ Vincenzo & Massimo 2015, p. 15.
  510. ^ Vincenzo & Massimo 2015, p. 13.
  511. ^ a b Schefuß et al. 2017, p. 9.
  512. ^ Schuster & Nutz 2016, p. 1616.
  513. ^ Russell & Ivory 2018, p. 11.
  514. ^ Lebamba et al. 2016, p. 137.
  515. ^ Lézine et al. 2013, p. 334.
  516. ^ Sachse et al. 2018, p. 3261.
  517. ^ a b Lézine 2017, p. 19.
  518. ^ Sachse et al. 2018, p. 3262.
  519. ^ Claussen et al. 1999, p. 2040.
  520. ^ Maslin, Manning & Brierley 2018, p. 4.
  521. ^ Maslin, Manning & Brierley 2018, p. 5.
  522. ^ a b Reimer et al. 2010, p. 41.
  523. ^ Morrill, Overpeck & Cole 2016, p. 473.
  524. ^ Fedotov, A.P; Chebykin, E.P; Yu, Semenov M; Vorobyova, S.S; Yu, Osipov E; Golobokova, L.P; Pogodaeva, T.V; Zheleznyakova, T.O; Grachev, M.A; Tomurhuu, D; Oyunchimeg, Ts; Narantsetseg, Ts; Tomurtogoo, O; Dolgikh, P.T; Arsenyuk, M.I; De Batist, M (July 2004). "Changes in the volume and salinity of Lake Khubsugul (Mongolia) in response to global climate changes in the upper Pleistocene and the Holocene". Palaeogeography, Palaeoclimatology, Palaeoecology. 209 (1–4): 256. doi:10.1016/j.palaeo.2003.12.022. ISSN 0031-0182.
  525. ^ Marsicek et al. 2013, p. 130.
  526. ^ Aharon, Paul; Dhungana, Rajesh (August 2017). "Ocean-atmosphere interactions as drivers of mid-to-late Holocene rapid climate changes: Evidence from high-resolution stalagmite records at DeSoto Caverns, Southeast USA". Quaternary Science Reviews. 170: 78. Bibcode:2017QSRv..170...69A. doi:10.1016/j.quascirev.2017.06.023. ISSN 0277-3791.
  527. ^ Wahl, David; Byrne, Roger; Anderson, Lysanna (November 2014). "An 8700 year paleoclimate reconstruction from the southern Maya lowlands". Quaternary Science Reviews. 103: 21. Bibcode:2014QSRv..103...19W. doi:10.1016/j.quascirev.2014.08.004. ISSN 0277-3791.
  528. ^ Rowe, Harold D; Guilderson, Thomas P; Dunbar, Robert B; Southon, John R; Seltzer, Geoffrey O; Mucciarone, David A; Fritz, Sherilyn C; Baker, Paul A (September 2003). "Late Quaternary lake-level changes constrained by radiocarbon and stable isotope studies on sediment cores from Lake Titicaca, South America". Global and Planetary Change. 38 (3–4): 287. Bibcode:2003GPC....38..273R. doi:10.1016/S0921-8181(03)00031-6. ISSN 0921-8181.
  529. ^ Shuman, Bryan N.; Serravezza, Marc (October 2017). "Patterns of hydroclimatic change in the Rocky Mountains and surrounding regions since the last glacial maximum". Quaternary Science Reviews. 173: 74. Bibcode:2017QSRv..173...58S. doi:10.1016/j.quascirev.2017.08.012. ISSN 0277-3791.
  530. ^ Shinker, Jacqueline J.; Powers, Kristine; Hougardy, Devin D.; Carter, Grace E.; Shuman, Bryan N. (1 March 2014). "A north–south moisture dipole at multi-century scales in the Central and Southern Rocky Mountains, U.S.A., during the late Holocene". Rocky Mountain Geology. 49 (1): 45. doi:10.2113/gsrocky.49.1.33. ISSN 1555-7332.
  531. ^ a b Pirie et al. 2009, p. 931.
  532. ^ Lernia et al. 2013, p. 120.
  533. ^ Andersen, Gidske L.; Krzywinski, Knut; Talib, Mohamed; Saadallah, Ahmed E.M.; Hobbs, Joseph J.; Pierce, Richard H. (July 2014). "Traditional nomadic tending of trees in the Red Sea Hills". Journal of Arid Environments. 106: 36. Bibcode:2014JArEn.106...36A. doi:10.1016/j.jaridenv.2014.02.009. ISSN 0140-1963.
  534. ^ Tafuri et al. 2006, p. 392.
  535. ^ Schuster & Nutz 2016, p. 1609.
  536. ^ Junginger & Trauth 2013, p. 176.
  537. ^ Junginger & Trauth 2013, p. 175.
  538. ^ Kuper 2006, p. 415.
  539. ^ a b Linstädter & Kröpelin 2004, p. 765.
  540. ^ Mercuri et al. 2018, p. 228.
  541. ^ Brooks et al. 2007, pp. 262–263.
  542. ^ Magny & Haas 2004, p. 428.
  543. ^ a b Cremaschi & Zerboni 2009, p. 700.
  544. ^ Castañeda et al. 2016, p. 47.
  545. ^ a b Bar-Matthews, Miryam; Ayalon, Avner; Gilmour, Mabs; Matthews, Alan; Hawkesworth, Chris J. (September 2003). "Sea–land oxygen isotopic relationships from planktonic foraminifera and speleothems in the Eastern Mediterranean region and their implication for paleorainfall during interglacial intervals". Geochimica et Cosmochimica Acta. 67 (17): 3195. Bibcode:2003GeCoA..67.3181B. doi:10.1016/S0016-7037(02)01031-1. ISSN 0016-7037.
  546. ^ Cremaschi & Zerboni 2009, p. 699.
  547. ^ Sachse et al. 2018, p. 3264.
  548. ^ a b Brooks et al. 2007, p. 261.
  549. ^ Tafuri et al. 2006, p. 399.
  550. ^ Brooks et al. 2007, p. 262.
  551. ^ Smith, Alison J. (27 July 2016). "Century-scale Holocene processes as a source of natural selection pressure in human evolution: Holocene climate and the Human Genome Project". The Holocene. 17 (5): 692–693. Bibcode:2007Holoc..17..689S. doi:10.1177/0959683607079003.
  552. ^ Spinage 2012, p. 58.
  553. ^ Médail et al. 2013, p. 2.
  554. ^ Boratyński, Adam; Ok, Tolga; Boratyńska, Krystyna; Dagher-Kharrat, Magda Bou; Romo, Angel; Dering, Monika; Sękiewicz, Katarzyna (28 September 2018). "Phylogenetic and biogeographic insights into long-lived Mediterranean Cupressus taxa with a schizo-endemic distribution and Tertiary origin". Botanical Journal of the Linnean Society. 188 (2): 15. doi:10.1093/botlinnean/boy049. ISSN 0024-4074.
  555. ^ Escoriza, Daniel; Bakhouche, Badis (2017). "11. Genus Malpolon: New distribution area in Algeria". The Herpetological Bulletin (140): 35.
  556. ^ Faith, J. Tyler (January 2014). "Late Pleistocene and Holocene mammal extinctions on continental Africa". Earth-Science Reviews. 128: 115. Bibcode:2014ESRv..128..105F. doi:10.1016/j.earscirev.2013.10.009. ISSN 0012-8252.
  557. ^ Vilhelmsen, Lars (7 March 2005). "Chalinus albitibialis , a new species of Orussidae (Insecta, Hymenoptera) from Morocco". Zootaxa. 880 (1): 6. doi:10.11646/zootaxa.880.1.1. ISSN 1175-5334.
  558. ^ Hassanin, Alexandre; Ropiquet, Anne; Gourmand, Anne-Laure; Chardonnet, Bertrand; Rigoulet, Jacques (March 2007). "Mitochondrial DNA variability in Giraffa camelopardalis: consequences for taxonomy, phylogeography and conservation of giraffes in West and central Africa". Comptes Rendus Biologies. 330 (3): 265–74. doi:10.1016/j.crvi.2007.02.008. ISSN 1631-0691. PMID 17434121.
  559. ^ Gross et al. 2014, p. 14473.
  560. ^ Salzmann, Ulrich; Hoelzmann, Philipp (1 February 2005). "The Dahomey Gap: an abrupt climatically induced rain forest fragmentation in West Africa during the late Holocene". The Holocene. 15 (2): 190. Bibcode:2005Holoc..15..190S. doi:10.1191/0959683605hl799rp. ISSN 0959-6836.
  561. ^ Hély et al. 2009, p. 684.
  562. ^ White et al. 2011, p. 472.
  563. ^ Adkins, Menocal & Eshel 2006, p. 2.
  564. ^ Zielhofer et al. 2017, p. 119.
  565. ^ D'Odorico, Paolo; Porporato, Amilcare, eds. (2006). Dryland Ecohydrology. Dordrecht: Kluwer Academic Publishers. p. 589. doi:10.1007/1-4020-4260-4. ISBN 978-1-4020-4259-1.
  566. ^ Muschitiello et al. 2015, p. 91.
  567. ^ Muschitiello et al. 2015, p. 93.
  568. ^ Muschitiello et al. 2015, pp. 94–95.
  569. ^ Muschitiello et al. 2015, p. 96.
  570. ^ a b Hoelzmann & Holmes 2017, p. 5.
  571. ^ Baumhauer & Runge 2009, p. 25.
  572. ^ Gasse 2000, p. 190.
  573. ^ a b Junginger et al. 2014, p. 4.
  574. ^ Forman, Wright & Bloszies 2014, p. 88.
  575. ^ Lézine, Duplessy & Cazet 2005, p. 225.
  576. ^ Lézine et al. 2017, p. 69.
  577. ^ Spinage 2012, p. 60.
  578. ^ a b Brooks et al. 2007, p. 267.
  579. ^ Donnelly et al. 2017, p. 6221.
  580. ^ Burr et al. 2014, pp. 16–17.
  581. ^ a b "Impacts of 1.5°C of Global Warming on Natural and Human Systems". IPCC. 23 May 2019. p. 197. Retrieved 29 December 2018.
  582. ^ a b Burr et al. 2014, p. 11.
  583. ^ Petoukhov et al. 2003, p. 100.
  584. ^ Petoukhov et al. 2003, p. 114.
  585. ^ Petoukhov et al. 2003, p. 113.
  586. ^ Brooks et al. 2007, p. 268.
  587. ^ Brooks et al. 2007, p. 269.

Sources

External links