Jump to content

Youngest Toba eruption

Coordinates: 2°41′04″N 98°52′32″E / 2.6845°N 98.8756°E / 2.6845; 98.8756
From Wikipedia, the free encyclopedia

The Toba eruption (sometimes called the Toba supereruption or the Youngest Toba eruption) was a supervolcanic eruption that occurred about 74,000 years ago during the Late Pleistocene[1] at the site of present-day Lake Toba in Sumatra, Indonesia. It was the last in a series of at least four caldera-forming eruptions at this location, with the earlier known caldera having formed around 1.2 million years ago.[2] This last eruption had an estimated VEI of 8, making it the largest-known explosive volcanic eruption in the Quaternary, and one of the largest known explosive eruptions in the Earth's history.

Youngest Toba eruption
Artist's impression of early stages of eruption from about 42 km (26 mi) above northern Sumatra
VolcanoToba Caldera Complex
Datec. 74,000 years BP
LocationSumatra, Indonesia
2°41′04″N 98°52′32″E / 2.6845°N 98.8756°E / 2.6845; 98.8756
VEI8
ImpactCovered the Indian subcontinent in 5 cm (2.0 in) of ash,[3] volcanic winter may have caused a severe human population bottleneck
Deaths(Potentially) almost all of humanity, leaving around 3,000–10,000 humans left on the planet
Lake Toba is the resulting crater lake

Eruption[edit]

Location of Lake Toba shown in red on map

Chronology of the Toba eruption[edit]

The exact year of the eruption is unknown, but the pattern of ash deposits suggests that it occurred during the northern summer because only the summer monsoon could have deposited Toba ashfall in the South China Sea.[4] The eruption lasted perhaps 9 to 14 days.[5] The most recent two high-precision argon–argon datings dated the eruption to 73,880 ± 320[6] and 73,700 ± 300 years ago.[7] Five distinct magma bodies were activated within a few centuries before the eruption.[8][9] The eruption commenced with small and limited air-fall and was directly followed by the main phase of ignimbrite flows.[10] The ignimbrite phase is characterized by low eruption fountain,[11] but co-ignimbrite column developed on top of pyroclastic flows reached a height of 32 km (20 mi).[12] Petrological constraints on sulfur emission yielded a wide range from 1×1013 to 1×1015 g, depending on the existence of separate sulfur gas in the Toba magma chamber.[13][14] The lower end of estimate is due to the low solubility of sulfur in the magma.[13] Ice core records estimate the sulfur emission on the order of 1×1014 g.[15]

Effects of the eruption[edit]

Bill Rose and Craig Chesner of Michigan Technological University have estimated that the total amount of material released in the eruption was at least 2,800 km3 (670 cu mi)[16]—about 2,000 km3 (480 cu mi) of ignimbrite that flowed over the ground, and approximately 800 km3 (190 cu mi) that fell as ash mostly to the west. However, as more outcrops become available, the most recent estimate of eruptive volume is 3,800 km3 (910 cu mi) dense-rock equivalent (DRE), of which 1,800 km3 (430 cu mi) was deposited as ash fall and 2,000 km3 (480 cu mi) as ignimbrite, making this eruption the largest during the Quaternary period.[17] Previous volume estimates have ranged from 2,000 km3 (480 cu mi)[5] to 6,000 km3 (1,400 cu mi).[18] Inside the caldera, the maximum thickness of pyroclastic flows is over 600 m (2,000 ft).[19] The outflow sheet originally covered an area of 20,000–30,000 km2 (7,700–11,600 sq mi) with thickness nearly 100 m (330 ft), likely reaching into the Indian Ocean and the Straits of Malacca.[10] The air-fall of this eruption blanketed the Indian subcontinent in a layer of 5 cm (2.0 in) ash,[20] the Arabian Sea in 1 mm (0.039 in),[21] the South China Sea in 3.5 cm (1.4 in),[4] and Central Indian Ocean Basin in 10 cm (3.9 in).[22] Its horizon of ashfall covered an area of more than 38,000,000 km2 (15,000,000 sq mi) in 1 cm (0.39 in) or more thickness.[17] In Sub-Saharan Africa, microscopic glass shards from this eruption are also discovered on the south coast of South Africa,[23] in the lowlands of northwest Ethiopia,[24] in Lake Malawi,[25] and in Lake Chala.[26] In South China, Toba tephras is found in Huguangyan Maar Lake.[27]

The subsequent collapse formed a caldera that filled with water, creating Lake Toba. The island in the center of the lake is formed by a resurgent dome.

Climatic effects[edit]

Climate at time of eruption[edit]

Greenland stadial 20 (GS20) is a millennium-long cold event in the north Atlantic ocean that started around the time of Toba eruption.[28] The timing of the initiation of GS20 is dated to 74.0–74.2 kyr, and the entire event lasted about 1,500 years.[28][29] It is the stadial part of Dansgaard–Oeschger event 20 (DO20), commonly explained by an abrupt reduction in the strength of the Atlantic meridional overturning circulation (AMOC). Weaker AMOC caused warming in Southern Ocean and Antarctica, and this asynchrony is known as bipolar seesaw.[30][31] The start of GS20 cooling event corresponds to the start of Antarctic Isotope Maxima 19 (AIM19) warming event.[32] GS20 was associated with iceberg discharges into the North Atlantic, thus it was also named Heinrich stadial 7a.[33] Heinrich events tend to be longer, colder and with weaker AMOC in the Atlantic ocean than other DO stadials.[30] From 74 to 58 kyr, Earth transitioned from interglacial MIS 5 to glacial MIS 4, experiencing cooling and glacial expansion.[34][35] This transition is a part of Pleistocene interglacial-glacial cycle driven by variations in the earth's orbit.[36] Ocean temperature cooled by 0.9 °C (1.6 °F).[37] Sea level fell 60 m (200 ft).[38] Northern Hemisphere ice sheets embarked on significant expansion and surpassed the extent of Last Glacial Maximum in eastern Europe, Northeast Asia and the North American Cordillera.[39] Southern Hemisphere glaciation grew to its maximum extent during MIS 4.[40] Australasian region, Africa and Europe were characterized by increasingly cold and arid environment.[41][42][43]

Possible climate records of eruption[edit]

While Toba eruption occurred in the backdrop of rapid climate transitions of GS20 and MIS 4 triggered by changes in ocean currents and insolation,[44][28] whether the eruption played any role in accelerating these events is much more debated. South China Sea marine records of climate, sampled at every centennial interval, shows 1 °C (1.8 °F) cooling above Toba ash layer for a thousand year but the authors concede that it may just be GS20.[45] Arabian Sea marine records confirm that Toba ash occurred after the onset of GS20 but also that GS20 is not colder than GS21 in the records, from which authors conclude that the eruption did not intensify GS20 cooling.[46] Dense sampling of environmental records, at every 69 year interval, in Lake Malawi, show no cooling-induced change in lake ecology and in grassy woodlands after the deposition of Toba ash,[25][47] but cooling-forced aridity killed high elevation afromontane forests.[48] The Lake Malawi studies concluded that the environmental effects of the eruption were mild and limited to less than a decade in East Africa,[47] but these studies are questioned due to sediment mixing which would have diminished the cooling signal.[49] Environmental records from a Middle Stone Age site in Ethiopia, however, shows that a severe drought occurred concurrently with Toba ash layer which altered early human foraging behaviours.[24]

No Toba ash has been identified in ice core records, but four sulfate events within the ice strata have been proposed to possibly represent the deposition of aerosols from Toba eruption.[50][32][51] One sulfate event at 73.75–74.16 kyr, which has all the characteristics of the Toba eruption, is among the largest sulfate loadings that have ever been identified.[51] In the ice core records, GS20 cooling was already underway by the time of sulfate deposition, nonetheless a 110-year period of accelerated cooling followed the sulfate event, and the authors interpret this acceleration as AMOC weakened by the Toba eruption.[15]

Climate modeling[edit]

The modeled climate effects of the Toba eruption hinges on the mass of sulfurous gases and aerosol microphysical processes. Modeling on an emission of 8.5×1014 g of sulfur, which is 100 times the 1991 Pinatubo sulphur, volcanic winter has a maximum global mean cooling of 3.5 °C (6.3 °F) and returns gradually within the range of natural variability 5 years after the eruption. An initiation of 1,000-year cold period or ice age is not supported by the model.[52][53] Two other emission scenarios, 1×1014 g and 1×1015 g, are investigated using state-of-art simulations provided by the Community Earth System Model. Maximum global mean cooling is 2.3 °C (4.1 °F) for the lower emission and 4.1 °C (7.4 °F) for the higher emission. Strong decrease in precipitation occurs in high emission. Negative temperature anomalies return to less than 1 °C (1.8 °F) within 3 and 6 years for each emission scenario after the eruption.[54] But so far no model can simulate aerosol microphysical processes with sufficient accuracy, empirical constraints from historical eruptions suggest that aerosol size may substantially reduce magnitude of cooling to less than 1.5 °C (2.7 °F) no matter how much sulfur emitted.[55]

Toba catastrophe theory[edit]

The Toba catastrophe theory holds that the eruption caused a severe global volcanic winter of six to ten years and contributed to a 1,000-year-long cooling episode, resulting in a genetic bottleneck in humans.[56][57] However, some physical evidence disputes the association with the millennium-long cold event and genetic bottleneck, and some consider the theory disproven.[58][48][59][60][61]

History[edit]

In 1972, an analysis of human hemoglobins found very few variants, and to account for the low frequency of variation human population must have been as low as a few thousand until very recently.[62] More genetic studies confirmed an effective population on the order of 10,000 for much of human history.[63][64] Subsequent research on the differences in human mitochondrial DNA sequences dated a rapid growth from a small effective population size of 1,000 to 10,000, sometime between 35 and 65 kyr.[65][66][67]

The large magnitude of the Toba eruption has been known since 1939, and various techniques dated the timing of the event to 73,000 to 75,000 years ago.[5] A study published in 1993 suggested that the eruption accelerated climate and environmental transition from the last interglacial period MIS 5 to the last glacial period MIS 4.[68]

In 1993, science journalist Ann Gibbons posited that population growth was suppressed by the cold climate of the last Pleistocene Ice Age, possibly exacerbated by the Toba eruption. The subsequent explosive human expansion was believed to be the result of the end of the ice age.[69] Geologist Michael R. Rampino of New York University and volcanologist Stephen Self of the University of Hawaiʻi at Mānoa supported her theory.[70] In 1998, anthropologist Stanley H. Ambrose of the University of Illinois Urbana-Champaign hypothesized that the Toba eruption caused a human population crash to only a few thousand surviving individuals, and the subsequent recovery was suppressed by the global glacial condition of MIS 4 until the climate eventually transitioned to the warmer condition of MIS 3 about 60,000 years ago, during which rapid human population expansion occurred.[56]

Possible effects on Homo[edit]

At least two other Homo lineages, H. neanderthals, and Denisovans, survived the Toba eruption and subsequent MIS 4 ice age, as their latest presence are dated to ca. 40 kyr,[71] and ca. 55 kyr.[72] Other lineages including H. floresiensis,[73] H. luzonensis,[74] and Penghu 1[75] may have also survived through the eruption. More recently, reconstructions of human demographic history using whole-genome sequencing[76][77][78] and discoveries of archaeological cultures with Toba ash layer[79][23][24] add further light to how humans had fared during the eruption and the following GS20 and MIS 4 ice age.

Human demographic history[edit]

Recent analysis applies Markov model to the complete set of genetic material to infer human population history.[80][81] In non-African populations, studies recover a long-term steep decline in numbers starting 200 kyr and reaching the lowest point around 40–60 kyr.[80][76] During this bottleneck non-African populations experienced 5- to 15-fold reduction,[82] with only 1,000–3,000 remaining individuals at 50 kyr, consistent with the earliest mtDNA studies.[76][77][81] This severe non-African contraction is consistent with founder effect caused by Out-of-Africa dispersal. As a small group with a size of a few thousand people migrated from the African continent into the Near East, the drastic reduction in numbers imprinted on non-African genomic diversity.[76][82][83]

African populations experienced a slightly earlier, milder bottleneck and recovered earlier.[81][84] Luhya and Maasai people attained their lowest numbers around 70–80 kyr, while Yoruba people reached a nadir around 50 kyr.[81] The estimated remaining effective population sizes are around 10,000 individuals, larger than the estimated non-African size during their bottleneck.[76][77][78] Unlike the non-African populations, there is no consensus as to the cause of African bottleneck. Proposed causes include climatic deterioration (from MIS 5, Toba eruption, GS20 and/or MIS 4),[49][83][85] reduction in substructure across African populations, and founder effects from the dispersal within Africa.[83]

Earlier genetic analysis of Alu sequences across the entire human genome has shown that the effective human population size was less than 26,000 at 1.2 million years ago; possible explanations for the low population size of human ancestors may include repeated population crashes or periodic replacement events from competing Homo subspecies.[86] Whole-genome analysis similarly recovers very low African population sizes around 1 million years ago.[77][78][87] This 1 million year old bottleneck is thought to have been caused by severe ice age MIS 22 which marked the mid-Pleistocene climate transition with widespread aridity across Africa.[87][88]

Archaeological studies[edit]

Other research has cast doubt on an association between the Toba Caldera Complex and a genetic bottleneck. For example, ancient stone tools at the Jurreru Valley in southern India were found above and below a thick layer of ash from the Toba eruption and were very similar across these layers, suggesting that the dust clouds from the eruption did not wipe out this local population.[89][90][91] However, another site in India, the Middle Son Valley, exhibits evidence of a major population decline and it has been suggested that the abundant springs of the Jurreru Valley may have offered its inhabitants unique protection.[92] Additional archaeological evidence from southern and northern India also suggests a lack of evidence for effects of the eruption on local populations, causing the authors of the study to conclude, "many forms of life survived the supereruption, contrary to other research which has suggested significant animal extinctions and genetic bottlenecks".[93] However, some researchers have questioned the techniques utilized to date artifacts to the period subsequent to the Toba supervolcano.[94] The Toba Catastrophe also coincides with the disappearance of the Skhul and Qafzeh hominins.[95] Evidence from pollen analysis has suggested prolonged deforestation in South Asia, and some researchers have suggested that the Toba eruption may have forced humans to adopt new adaptive strategies, which may have permitted them to replace Neanderthals and "other archaic human species".[96][97]

Genetic bottlenecks in other mammals[edit]

Some evidence indicates population crashes of other animals after the Toba eruption. The populations of the Eastern African chimpanzee,[98] Bornean orangutan,[99] central Indian macaque,[100] cheetah and tiger,[101] all expanded from very small populations around 70,000–55,000 years ago.

See also[edit]

Citations and notes[edit]

  1. ^ "Surprisingly, Humanity Survived the Super-volcano 74,000 Years Ago". Haaretz.
  2. ^ Stratigraphy of the Toba Tuffs and the evolution of the Toba Caldera Complex, Sumatra, Indonesia
  3. ^ Petraglia, Michael D.; Ditchfield, Peter; Jones, Sacha; Korisettar, Ravi; Pal, J.N. (2012). "The Toba volcanic super-eruption, environmental change, and hominin occupation history in India over the last 140,000 years". Quaternary International. 258: 119–134. doi:10.1016/j.quaint.2011.07.042. ISSN 1040-6182.
  4. ^ a b Bühring, Christian; Sarnthein, Michael (2000). "Toba ash layers in the South China Sea: Evidence of contrasting wind directions during eruption ca. 74 ka: Comment and Reply". Geology. 28 (11): 1056. doi:10.1130/0091-7613(2000)28<1056:talits>2.0.co;2. ISSN 0091-7613.
  5. ^ a b c Ninkovich, D.; Sparks, R. S. J.; Ledbetter, M. T. (1978-09-01). "The exceptional magnitude and intensity of the Toba eruption, sumatra: An example of the use of deep-sea tephra layers as a geological tool". Bulletin Volcanologique. 41 (3): 286–298. Bibcode:1978BVol...41..286N. doi:10.1007/BF02597228. ISSN 1432-0819. S2CID 128626019.
  6. ^ Storey, Michael; Roberts, Richard G.; Saidin, Mokhtar (2012-11-13). "Astronomically calibrated 40 Ar/ 39 Ar age for the Toba supereruption and global synchronization of late Quaternary records". Proceedings of the National Academy of Sciences. 109 (46): 18684–18688. Bibcode:2012PNAS..10918684S. doi:10.1073/pnas.1208178109. ISSN 0027-8424. PMC 3503200. PMID 23112159.
  7. ^ Channell, J.E.T.; Hodell, D.A. (2017). "High-precision 40Ar/39Ar dating of Pleistocene tuffs and temporal anchoring of the Matuyama-Brunhes boundary". Quaternary Geochronology. 42: 56–59. doi:10.1016/j.quageo.2017.08.002. ISSN 1871-1014.
  8. ^ Pearce, Nicholas J.G.; Westgate, John A.; Gualda, Guilherme A.R.; Gatti, Emma; Muhammad, Ros F. (2019-10-14). "Tephra glass chemistry provides storage and discharge details of five magma reservoirs which fed the 75 ka Youngest Toba Tuff eruption, northern Sumatra". Journal of Quaternary Science. 35 (1–2): 256–271. doi:10.1002/jqs.3149. hdl:2160/dba3b012-8369-4dbb-8a89-1102f11e92c3. ISSN 0267-8179.
  9. ^ Lubbers, Jordan; Kent, Adam J. R.; de Silva, Shanaka (2024-01-18). "Constraining magma storage conditions of the Toba magmatic system: a plagioclase and amphibole perspective". Contributions to Mineralogy and Petrology. 179 (2). doi:10.1007/s00410-023-02089-7. ISSN 0010-7999.
  10. ^ a b Chesner, Craig A. (2012). "The Toba Caldera Complex". Quaternary International. 258: 5–18. doi:10.1016/j.quaint.2011.09.025. ISSN 1040-6182.
  11. ^ CHESNER, C (1998-03-01). "Petrogenesis of the Toba Tuffs, Sumatra, Indonesia". Journal of Petrology. 39 (3): 397–438. doi:10.1093/petrology/39.3.397. ISSN 1460-2415.
  12. ^ Woods, Andrew W.; Wohletz, Kenneth (1991). "Dimensions and dynamics of co-ignimbrite eruption columns". Nature. 350 (6315): 225–227. doi:10.1038/350225a0. ISSN 1476-4687.
  13. ^ a b Chesner, Craig A.; Luhr, James F. (2010-11-30). "A melt inclusion study of the Toba Tuffs, Sumatra, Indonesia". Journal of Volcanology and Geothermal Research. 197 (1–4): 259–278. Bibcode:2010JVGR..197..259C. doi:10.1016/j.jvolgeores.2010.06.001.
  14. ^ Scaillet, Bruno; Luhr, James F.; Carroll, Michael R. (2003), "Petrological and volcanological constraints on volcanic sulfur emissions to the atmosphere", Volcanism and the Earth's Atmosphere, Geophysical Monograph Series, vol. 139, Washington, D. C.: American Geophysical Union, pp. 11–40, doi:10.1029/139gm02, ISBN 0-87590-998-1, retrieved 2024-04-25
  15. ^ a b Lin, Jiamei; Abbott, Peter M.; Sigl, Michael; Steffensen, Jørgen P.; Mulvaney, Robert; Severi, Mirko; Svensson, Anders (2023). "Bipolar ice-core records constrain possible dates and global radiative forcing following the ∼74 ka Toba eruption". Quaternary Science Reviews. 312: 108162. doi:10.1016/j.quascirev.2023.108162. ISSN 0277-3791.
  16. ^ "Supersized eruptions are all the rage!". USGS. 28 April 2005.
  17. ^ a b Kutterolf, S.; Schindlbeck-Belo, J.C.; Müller, F.; Pank, K.; Lee, H.-Y.; Wang, K.-L.; Schmitt, A.K. (2023). "Revisiting the occurrence and distribution of Indian Ocean Tephra: Quaternary marine Toba ash inventory". Journal of Volcanology and Geothermal Research. 441: 107879. doi:10.1016/j.jvolgeores.2023.107879.
  18. ^ Self, S.; Gouramanis, C.; Storey, M. (2019-12-01). "The Young Toba Tuff (73.9 ka) Magma Body – True Size and the most Extensive and Voluminous Ignimbrite Yet Known?". AGU Fall Meeting Abstracts. 2019: V51H–0141. Bibcode:2019AGUFM.V51H0141S.
  19. ^ Chesner, Craig A.; Rose, William I. (1991-06-01). "Stratigraphy of the Toba Tuffs and the evolution of the Toba Caldera Complex, Sumatra, Indonesia". Bulletin of Volcanology. 53 (5): 343–356. doi:10.1007/BF00280226. ISSN 1432-0819.
  20. ^ Petraglia, Michael D.; Ditchfield, Peter; Jones, Sacha; Korisettar, Ravi; Pal, J.N. (2012). "The Toba volcanic super-eruption, environmental change, and hominin occupation history in India over the last 140,000 years". Quaternary International. 258: 119–134. doi:10.1016/j.quaint.2011.07.042. ISSN 1040-6182.
  21. ^ Von Rad, Ulrich; Burgath, Klaus-Peter; Pervaz, Muhammad; Schulz, Hartmut (2002). "Discovery of the Toba Ash ( c. 70 ka) in a high-resolution core recovering millennial monsoonal variability off Pakistan". Geological Society, London, Special Publications. 195 (1): 445–461. doi:10.1144/GSL.SP.2002.195.01.25. ISSN 0305-8719.
  22. ^ Pattan, J. N; Shane, Phil; Banakar, V. K (1999-03-01). "New occurrence of Youngest Toba Tuff in abyssal sediments of the Central Indian Basin". Marine Geology. 155 (3): 243–248. doi:10.1016/S0025-3227(98)00160-1. ISSN 0025-3227.
  23. ^ a b Smith, Eugene I.; Jacobs, Zenobia; Johnsen, Racheal; Ren, Minghua; Fisher, Erich C.; Oestmo, Simen; Wilkins, Jayne; Harris, Jacob A.; Karkanas, Panagiotis; Fitch, Shelby; Ciravolo, Amber; Keenan, Deborah; Cleghorn, Naomi; Lane, Christine S.; Matthews, Thalassa (2018). "Humans thrived in South Africa through the Toba eruption about 74,000 years ago". Nature. 555 (7697): 511–515. doi:10.1038/nature25967. ISSN 1476-4687.
  24. ^ a b c Kappelman, John; Todd, Lawrence C.; Davis, Christopher A.; Cerling, Thure E.; Feseha, Mulugeta; Getahun, Abebe; Johnsen, Racheal; Kay, Marvin; Kocurek, Gary A.; Nachman, Brett A.; Negash, Agazi; Negash, Tewabe; O'Brien, Kaedan; Pante, Michael; Ren, Minghua (2024). "Adaptive foraging behaviours in the Horn of Africa during Toba supereruption". Nature. 628 (8007): 365–372. doi:10.1038/s41586-024-07208-3. ISSN 1476-4687. PMID 38509364.
  25. ^ a b Lane, C. S.; Chorn, B. T.; Johnson, T. C. (2013). "Ash from the Toba supereruption in Lake Malawi shows no volcanic winter in East Africa at 75 ka". Proceedings of the National Academy of Sciences. 110 (20): 8025–8029. Bibcode:2013PNAS..110.8025L. doi:10.1073/pnas.1301474110. PMC 3657767. PMID 23630269.
  26. ^ Baxter, A. J.; Verschuren, D.; Peterse, F.; Miralles, D. G.; Martin-Jones, C. M.; Maitituerdi, A.; Van der Meeren, T.; Van Daele, M.; Lane, C. S.; Haug, G. H.; Olago, D. O.; Sinninghe Damsté, J. S. (2023). "Reversed Holocene temperature–moisture relationship in the Horn of Africa". Nature. 620 (7973): 336–343. doi:10.1038/s41586-023-06272-5. hdl:1854/LU-01HF6GN7WZQ65R3C82NK0HC57E. ISSN 1476-4687. PMID 37558848.
  27. ^ Guo, Z., Liu, J., Chu, G., & JFW, N. (2002). Composition and origin of tephra of the Huguangyan Maar Lake. Quaternary Sciences, 22(3), 266-272.
  28. ^ a b c Polyak, Victor J.; Asmerom, Yemane; Lachniet, Matthew S. (2017-09-01). "Rapid speleothem δ13C change in southwestern North America coincident with Greenland stadial 20 and the Toba (Indonesia) supereruption". Geology. 45 (9): 843–846. doi:10.1130/G39149.1. ISSN 0091-7613.
  29. ^ Du, Wenjing; Cheng, Hai; Xu, Yao; Yang, Xunlin; Zhang, Pingzhong; Sha, Lijuan; Li, Hanying; Zhu, Xiaoyan; Zhang, Meiliang; Stríkis, Nicolás M.; Cruz, Francisco W.; Edwards, R. Lawrence; Zhang, Haiwei; Ning, Youfeng (2019). "Timing and structure of the weak Asian Monsoon event about 73,000 years ago". Quaternary Geochronology. 53: 101003. doi:10.1016/j.quageo.2019.05.002. ISSN 1871-1014.
  30. ^ a b Menviel, Laurie C.; Skinner, Luke C.; Tarasov, Lev; Tzedakis, Polychronis C. (2020). "An ice–climate oscillatory framework for Dansgaard–Oeschger cycles". Nature Reviews Earth & Environment. 1 (12): 677–693. doi:10.1038/s43017-020-00106-y. ISSN 2662-138X.
  31. ^ Anderson, H. J.; Pedro, J. B.; Bostock, H. C.; Chase, Z.; Noble, T. L. (2021-03-01). "Compiled Southern Ocean sea surface temperatures correlate with Antarctic Isotope Maxima". Quaternary Science Reviews. 255: 106821. doi:10.1016/j.quascirev.2021.106821. ISSN 0277-3791.
  32. ^ a b Svensson, A.; Bigler, M.; Blunier, T.; Clausen, H. B.; Dahl-Jensen, D.; Fischer, H.; Fujita, S.; Goto-Azuma, K.; Johnsen, S. J.; Kawamura, K.; Kipfstuhl, S.; Kohno, M.; Parrenin, F.; Popp, T.; Rasmussen, S. O. (2013-03-19). "Direct linking of Greenland and Antarctic ice cores at the Toba eruption (74 ka BP)". Climate of the Past. 9 (2): 749–766. Bibcode:2013CliPa...9..749S. doi:10.5194/cp-9-749-2013. hdl:2158/774798. ISSN 1814-9324. S2CID 17741316.
  33. ^ Davtian, Nina; Bard, Edouard (2023-03-13). "A new view on abrupt climate changes and the bipolar seesaw based on paleotemperatures from Iberian Margin sediments". Proceedings of the National Academy of Sciences. 120 (12). doi:10.1073/pnas.2209558120. ISSN 0027-8424.
  34. ^ Menking, James A.; Shackleton, Sarah A.; Bauska, Thomas K.; Buffen, Aron M.; Brook, Edward J.; Barker, Stephen; Severinghaus, Jeffrey P.; Dyonisius, Michael N.; Petrenko, Vasilii V. (2022-09-16). "Multiple carbon cycle mechanisms associated with the glaciation of Marine Isotope Stage 4". Nature Communications. 13 (1): 5443. doi:10.1038/s41467-022-33166-3. ISSN 2041-1723. PMC 9481522. PMID 36114188.
  35. ^ Doughty, Alice M.; Kaplan, Michael R.; Peltier, Carly; Barker, Stephen (2021). "A maximum in global glacier extent during MIS 4". Quaternary Science Reviews. 261: 106948. doi:10.1016/j.quascirev.2021.106948. ISSN 0277-3791.
  36. ^ Hays, J. D.; Imbrie, John; Shackleton, N. J. (1976-12-10). "Variations in the Earth's Orbit: Pacemaker of the Ice Ages: For 500,000 years, major climatic changes have followed variations in obliquity and precession". Science. 194 (4270): 1121–1132. doi:10.1126/science.194.4270.1121. ISSN 0036-8075. PMID 17790893.
  37. ^ Shackleton, Sarah; Menking, James A.; Brook, Edward; Buizert, Christo; Dyonisius, Michael N.; Petrenko, Vasilii V.; Baggenstos, Daniel; Severinghaus, Jeffrey P. (2021-10-27). "Evolution of mean ocean temperature in Marine Isotope Stage 4". Climate of the Past. 17 (5): 2273–2289. doi:10.5194/cp-17-2273-2021. ISSN 1814-9324.
  38. ^ Cutler, K.B; Edwards, R.L; Taylor, F.W; Cheng, H; Adkins, J; Gallup, C.D; Cutler, P.M; Burr, G.S; Bloom, A.L (2003). "Rapid sea-level fall and deep-ocean temperature change since the last interglacial period". Earth and Planetary Science Letters. 206 (3–4): 253–271. doi:10.1016/s0012-821x(02)01107-x. ISSN 0012-821X.
  39. ^ Batchelor, Christine L.; Margold, Martin; Krapp, Mario; Murton, Della K.; Dalton, April S.; Gibbard, Philip L.; Stokes, Chris R.; Murton, Julian B.; Manica, Andrea (2019-08-16). "The configuration of Northern Hemisphere ice sheets through the Quaternary". Nature Communications. 10 (1): 3713. doi:10.1038/s41467-019-11601-2. ISSN 2041-1723. PMC 6697730. PMID 31420542.
  40. ^ Schaefer, Joerg M.; Putnam, Aaron E.; Denton, George H.; Kaplan, Michael R.; Birkel, Sean; Doughty, Alice M.; Kelley, Sam; Barrell, David J.A.; Finkel, Robert C.; Winckler, Gisela; Anderson, Robert F.; Ninneman, Ulysses S.; Barker, Stephen; Schwartz, Roseanne; Andersen, Bjorn G. (2015). "The Southern Glacial Maximum 65,000 years ago and its Unfinished Termination". Quaternary Science Reviews. 114: 52–60. doi:10.1016/j.quascirev.2015.02.009.
  41. ^ Stewart, John R.; Fenberg, Phillip B. (2018-05-01). "A climatic context for the out-of-Africa migration: COMMENT". Geology. 46 (5): e442. doi:10.1130/g40057c.1. ISSN 0091-7613.
  42. ^ Helmens, Karin F. (2014). "The Last Interglacial–Glacial cycle (MIS 5–2) re-examined based on long proxy records from central and northern Europe". Quaternary Science Reviews. 86: 115–143. doi:10.1016/j.quascirev.2013.12.012. ISSN 0277-3791.
  43. ^ De Deckker, Patrick; Arnold, Lee J.; van der Kaars, Sander; Bayon, Germain; Stuut, Jan-Berend W.; Perner, Kerstin; Lopes dos Santos, Raquel; Uemura, Ryu; Demuro, Martina (2019). "Marine Isotope Stage 4 in Australasia: A full glacial culminating 65,000 years ago – Global connections and implications for human dispersal". Quaternary Science Reviews. 204: 187–207. doi:10.1016/j.quascirev.2018.11.017. ISSN 0277-3791.
  44. ^ Rampino, Michael R.; Self, Stephen (1992). "Volcanic winter and accelerated glaciation following the Toba super-eruption". Nature. 359 (6390): 50–52. doi:10.1038/359050a0. ISSN 1476-4687.
  45. ^ Huang, Chi-Yue; Zhao, Meixun; Wang, Chia-Chun; Wei, Ganjian (2001-10-15). "Cooling of the South China Sea by the Toba Eruption and correlation with other climate proxies ~71,000 years ago". Geophysical Research Letters. 28 (20): 3915–3918. Bibcode:2001GeoRL..28.3915H. doi:10.1029/2000GL006113. S2CID 128903263.
  46. ^ Schulz, Hartmut; Emeis, Kay-Christian; Erlenkeuser, Helmut; Rad, Ulrich von; Rolf, Christian (2002). "The Toba Volcanic Event and Interstadial/Stadial Climates at the Marine Isotopic Stage 5 to 4 Transition in the Northern Indian Ocean". Quaternary Research. 57 (1): 22–31. Bibcode:2002QuRes..57...22S. doi:10.1006/qres.2001.2291. ISSN 0033-5894. S2CID 129838182.
  47. ^ a b Jackson, Lily J.; Stone, Jeffery R.; Cohen, Andrew S.; Yost, Chad L. (2015-09-01). "High-resolution paleoecological records from Lake Malawi show no significant cooling associated with the Mount Toba supereruption at ca. 75 ka". Geology. 43 (9): 823–826. Bibcode:2015Geo....43..823J. doi:10.1130/G36917.1. ISSN 0091-7613.
  48. ^ a b Yost, Chad; et al. (March 2018). "Subdecadal phytolith and charcoal records from Lake Malawi, East Africa imply minimal effects on human evolution from the ~74 ka Toba supereruption". Journal of Human Evolution. 116. Elsevier: 75–94. doi:10.1016/j.jhevol.2017.11.005. PMID 29477183.
  49. ^ a b Ambrose, S. H. (2019), "Chapter 6 chronological calibration of Late Pleistocene Modern Human dispersals, climate change and Archaeology with Geochemical Isochrons", in Sahle, Yonatan; Reyes-Centeno, Hugo; Bentz, Christian (eds.), Modern Human Origins and Dispersal, Kerns Verlag, pp. 171–213
  50. ^ Zielinski, G. A.; Mayewski, P. A.; Meeker, L. D.; Whitlow, S.; Twickler, M. S.; Taylor, K. (1996-04-15). "Potential atmospheric impact of the Toba Mega-Eruption ~71,000 years ago". Geophysical Research Letters. 23 (8): 837–840. Bibcode:1996GeoRL..23..837Z. doi:10.1029/96GL00706.
  51. ^ a b Crick, Laura; Burke, Andrea; Hutchison, William; Kohno, Mika; Moore, Kathryn A.; Savarino, Joel; Doyle, Emily A.; Mahony, Sue; Kipfstuhl, Sepp; Rae, James W. B.; Steele, Robert C. J.; Sparks, R. Stephen J.; Wolff, Eric W. (2021-10-18). "New insights into the ~ 74 ka Toba eruption from sulfur isotopes of polar ice cores". Climate of the Past. 17 (5): 2119–2137. Bibcode:2021CliPa..17.2119C. doi:10.5194/cp-17-2119-2021. hdl:10023/24161. ISSN 1814-9324. S2CID 239203480.
  52. ^ Timmreck, Claudia; Graf, Hans-F.; Zanchettin, Davide; Hagemann, Stefan; Kleinen, Thomas; Krüger, Kirstin (2012-05-01). "Climate response to the Toba super-eruption: Regional changes". Quaternary International. 258: 30–44. Bibcode:2012QuInt.258...30T. doi:10.1016/j.quaint.2011.10.008.
  53. ^ Timmreck, Claudia; Graf, Hans-F.; Lorenz, Stephan J.; Niemeier, Ulrike; Zanchettin, Davide; Matei, Daniela; Jungclaus, Johann H.; Crowley, Thomas J. (2010-12-22). "Aerosol size confines climate response to volcanic super-eruptions". Geophysical Research Letters. 37 (24): n/a. doi:10.1029/2010GL045464. hdl:11858/00-001M-0000-0011-F70C-7. S2CID 12790660.
  54. ^ Black, Benjamin A.; Lamarque, Jean-François; Marsh, Daniel R.; Schmidt, Anja; Bardeen, Charles G. (2021-07-20). "Global climate disruption and regional climate shelters after the Toba supereruption". Proceedings of the National Academy of Sciences. 118 (29): e2013046118. Bibcode:2021PNAS..11813046B. doi:10.1073/pnas.2013046118. ISSN 0027-8424. PMC 8307270. PMID 34230096.
  55. ^ McGraw, Zachary; DallaSanta, Kevin; Polvani, Lorenzo M.; Tsigaridis, Kostas; Orbe, Clara; Bauer, Susanne E. (2024-02-15). "Severe Global Cooling After Volcanic Super-Eruptions? The Answer Hinges on Unknown Aerosol Size". Journal of Climate. 37 (4): 1449–1464. doi:10.1175/jcli-d-23-0116.1. ISSN 0894-8755.
  56. ^ a b Ambrose 1998.
  57. ^ Michael R. Rampino, Stanley H. Ambrose, 2000. "Volcanic winter in the Garden of Eden: The Toba supereruption and the late Pleistocene human population crash", Volcanic Hazards and Disasters in Human Antiquity, Floyd W. McCoy, Grant Heiken
  58. ^ "Toba super-volcano catastrophe idea 'dismissed'". BBC News. 30 April 2013. Retrieved 2017-01-08.
  59. ^ Ge, Yong; Gao, Xing (2020-09-10). "Understanding the overestimated impact of the Toba volcanic super-eruption on global environments and ancient hominins". Quaternary International. Current Research on Prehistoric Central Asia. 559: 24–33. Bibcode:2020QuInt.559...24G. doi:10.1016/j.quaint.2020.06.021. ISSN 1040-6182. S2CID 225418492.
  60. ^ Hawks, John (9 February 2018). "The so-called Toba bottleneck didn't happen". john hawks weblog.
  61. ^ Singh, Ajab; Srivastava, Ashok K. (2022-06-01). "Had Youngest Toba Tuff (YTT, ca. 75 ka) eruption really destroyed living media explicitly in entire Southeast Asia or just a theoretical debate? An extensive review of its catastrophic event". Journal of Asian Earth Sciences: X. 7: 100083. Bibcode:2022JAESX...700083S. doi:10.1016/j.jaesx.2022.100083. ISSN 2590-0560. S2CID 246416256.
  62. ^ Haigh, John; Smith, John Maynard (1972). "Population size and protein variation in man". Genetics Research. 19 (1): 73–89. doi:10.1017/S0016672300014282. ISSN 1469-5073.
  63. ^ Takahata, N. (1993). "Allelic genealogy and human evolution". Molecular Biology and Evolution. 10 (1): 2–22. doi:10.1093/oxfordjournals.molbev.a039995. ISSN 1537-1719. PMID 8450756.
  64. ^ Garesse, R (1988-04-01). "Drosophila melanogaster mitochondrial DNA: gene organization and evolutionary considerations". Genetics. 118 (4): 649–663. doi:10.1093/genetics/118.4.649. ISSN 1943-2631. PMC 1203320. PMID 3130291.
  65. ^ Harpending, Henry C.; Sherry, Stephen T.; Rogers, Alan R.; Stoneking, Mark (1993). "The Genetic Structure of Ancient Human Populations". Current Anthropology. 34 (4): 483–496. doi:10.1086/204195. ISSN 0011-3204.
  66. ^ Rogers, Alan R. (1995). "Genetic Evidence for a Pleistocene Population Explosion". Evolution. 49 (4): 608–615. doi:10.1111/j.1558-5646.1995.tb02297.x. PMID 28565146. S2CID 29309837.
  67. ^ Sherry, Stephen T.; Rogers, Alan R.; Harpending, Henry; Soodyall, Himla; Jenkins, Trefor; Stoneking, Mark (1994). "Mismatch Distributions of mtDNA Reveal Recent Human Population Expansions". Human Biology. 66 (5): 761–775. ISSN 0018-7143. JSTOR 41465014.
  68. ^ Rampino, Michael R.; Self, Stephen (1992-09-03). "Volcanic winter and accelerated glaciation following the Toba super-eruption". Nature. 359 (6390): 50–52. Bibcode:1992Natur.359...50R. doi:10.1038/359050a0. ISSN 1476-4687. S2CID 4322781.
  69. ^ Gibbons 1993.
  70. ^ Rampino, Michael R.; Self, Stephen (1993-12-24). "Bottleneck in Human Evolution and the Toba Eruption". Science. 262 (5142): 1955. Bibcode:1993Sci...262.1955R. doi:10.1126/science.8266085. ISSN 0036-8075. PMID 8266085.
  71. ^ Higham, Tom; Douka, Katerina; Wood, Rachel; Ramsey, Christopher Bronk; Brock, Fiona; Basell, Laura; Camps, Marta; Arrizabalaga, Alvaro; Baena, Javier; Barroso-Ruíz, Cecillio; Bergman, Christopher; Boitard, Coralie; Boscato, Paolo; Caparrós, Miguel; Conard, Nicholas J. (2014). "The timing and spatiotemporal patterning of Neanderthal disappearance". Nature. 512 (7514): 306–309. doi:10.1038/nature13621. ISSN 1476-4687.
  72. ^ Jacobs, Zenobia; Li, Bo; Shunkov, Michael V.; Kozlikin, Maxim B.; Bolikhovskaya, Nataliya S.; Agadjanian, Alexander K.; Uliyanov, Vladimir A.; Vasiliev, Sergei K.; O'Gorman, Kieran; Derevianko, Anatoly P.; Roberts, Richard G. (2019). "Timing of archaic hominin occupation of Denisova Cave in southern Siberia". Nature. 565 (7741): 594–599. doi:10.1038/s41586-018-0843-2. ISSN 1476-4687.
  73. ^ Sutikna, Thomas; Tocheri, Matthew W.; Morwood, Michael J.; Saptomo, E. Wahyu; Jatmiko; Awe, Rokus Due; Wasisto, Sri; Westaway, Kira E.; Aubert, Maxime; Li, Bo; Zhao, Jian-xin; Storey, Michael; Alloway, Brent V.; Morley, Mike W.; Meijer, Hanneke J. M. (2016). "Revised stratigraphy and chronology for Homo floresiensis at Liang Bua in Indonesia". Nature. 532 (7599): 366–369. doi:10.1038/nature17179. ISSN 1476-4687.
  74. ^ Détroit, Florent; Mijares, Armand Salvador; Corny, Julien; Daver, Guillaume; Zanolli, Clément; Dizon, Eusebio; Robles, Emil; Grün, Rainer; Piper, Philip J. (2019). "A new species of Homo from the Late Pleistocene of the Philippines". Nature. 568 (7751): 181–186. doi:10.1038/s41586-019-1067-9. ISSN 1476-4687.
  75. ^ Chang, Chun-Hsiang; Kaifu, Yousuke; Takai, Masanaru; Kono, Reiko T.; Grün, Rainer; Matsu'ura, Shuji; Kinsley, Les; Lin, Liang-Kong (2015-01-27). "The first archaic Homo from Taiwan". Nature Communications. 6 (1): 6037. doi:10.1038/ncomms7037. hdl:1885/12938. ISSN 2041-1723.
  76. ^ a b c d e Mallick, Swapan; Li, Heng; Lipson, Mark; Mathieson, Iain; Gymrek, Melissa; Racimo, Fernando; Zhao, Mengyao; Chennagiri, Niru; Nordenfelt, Susanne; Tandon, Arti; Skoglund, Pontus; Lazaridis, Iosif; Sankararaman, Sriram; Fu, Qiaomei; Rohland, Nadin (2016). "The Simons Genome Diversity Project: 300 genomes from 142 diverse populations". Nature. 538 (7624): 201–206. doi:10.1038/nature18964. hdl:11336/125570. ISSN 1476-4687.
  77. ^ a b c d A, Bergstrom; SA, McCarthy; R, Hui; MA, Almarri; Q, Ayub; P, Danecek; Y, Chen; S, Felkel; P, Hallast; J, Kamm; H, Blanche; JF, Deleuze; H, Cann; S, Mallick; D, Reich (2020-10-23). "Insights into human genetic variation and population history from 929 diverse genomes". Yearbook of Paediatric Endocrinology. doi:10.1530/ey.17.14.4. ISSN 1662-4009.
  78. ^ a b c Fan, Shaohua; Spence, Jeffrey P.; Feng, Yuanqing; Hansen, Matthew E.B.; Terhorst, Jonathan; Beltrame, Marcia H.; Ranciaro, Alessia; Hirbo, Jibril; Beggs, William; Thomas, Neil; Nyambo, Thomas; Mpoloka, Sununguko Wata; Mokone, Gaonyadiwe George; Njamnshi, Alfred K.; Fokunang, Charles (2023). "Whole-genome sequencing reveals a complex African population demographic history and signatures of local adaptation". Cell. 186 (5): 923–939.e14. doi:10.1016/j.cell.2023.01.042. ISSN 0092-8674. PMC 10568978.
  79. ^ Petraglia, Michael; Korisettar, Ravi; Boivin, Nicole; Clarkson, Christopher; Ditchfield, Peter; Jones, Sacha; Koshy, Jinu; Lahr, Marta Mirazón; Oppenheimer, Clive; Pyle, David; Roberts, Richard; Schwenninger, Jean-Luc; Arnold, Lee; White, Kevin (2007-07-06). "Middle Paleolithic Assemblages from the Indian Subcontinent Before and After the Toba Super-Eruption". Science. 317 (5834): 114–116. doi:10.1126/science.1141564. ISSN 0036-8075.
  80. ^ a b Schiffels, Stephan; Durbin, Richard (2014). "Inferring human population size and separation history from multiple genome sequences". Nature Genetics. 46 (8): 919–925. doi:10.1038/ng.3015. ISSN 1546-1718.
  81. ^ a b c d Terhorst, Jonathan; Kamm, John A.; Song, Yun S. (2017). "Robust and scalable inference of population history from hundreds of unphased whole genomes". Nature Genetics. 49 (2): 303–309. doi:10.1038/ng.3748. ISSN 1546-1718.
  82. ^ a b Henn, Brenna M.; Cavalli-Sforza, L. L.; Feldman, Marcus W. (2012-10-30). "The great human expansion". Proceedings of the National Academy of Sciences. 109 (44): 17758–17764. doi:10.1073/pnas.1212380109. ISSN 0027-8424. PMC 3497766. PMID 23077256.{{cite journal}}: CS1 maint: PMC format (link)
  83. ^ a b c Henn, Brenna M.; Botigué, Laura R.; Bustamante, Carlos D.; Clark, Andrew G.; Gravel, Simon (2015). "Estimating the mutation load in human genomes". Nature Reviews Genetics. 16 (6): 333–343. doi:10.1038/nrg3931. ISSN 1471-0064.
  84. ^ Li, Heng; Durbin, Richard (2011). "Inference of human population history from individual whole-genome sequences". Nature. 475 (7357): 493–496. doi:10.1038/nature10231. ISSN 1476-4687.
  85. ^ Powell, Adam; Shennan, Stephen; Thomas, Mark G. (2009-06-05). "Late Pleistocene Demography and the Appearance of Modern Human Behavior". Science. 324 (5932): 1298–1301. doi:10.1126/science.1170165. ISSN 0036-8075.
  86. ^ See Huff & others 2010, p.6; Gibbons 2010.
  87. ^ a b Hu, Wangjie; Hao, Ziqian; Du, Pengyuan; Di Vincenzo, Fabio; Manzi, Giorgio; Cui, Jialong; Fu, Yun-Xin; Pan, Yi-Hsuan; Li, Haipeng (2023). "Genomic inference of a severe human bottleneck during the Early to Middle Pleistocene transition". Science. 381 (6661): 979–984. doi:10.1126/science.abq7487. ISSN 0036-8075.
  88. ^ Muttoni, Giovanni; Kent, Dennis V. (2024-03-26). "Hominin population bottleneck coincided with migration from Africa during the Early Pleistocene ice age transition". Proceedings of the National Academy of Sciences. 121 (13). doi:10.1073/pnas.2318903121. ISSN 0027-8424. PMC 10990135. PMID 38466876.{{cite journal}}: CS1 maint: PMC format (link)
  89. ^ "Mount Toba Eruption – Ancient Humans Unscathed, Study Claims". Anthropology.net. 6 July 2007. Archived from the original on 2008-01-11. Retrieved 2008-04-20.
  90. ^ Sanderson, Katherine (July 2007). "Super-eruption: no problem?". Nature: news070702–15. doi:10.1038/news070702-15. S2CID 177216526. Archived from the original on December 7, 2008.
  91. ^ John Hawks (5 July 2007). "At last, the death of the Toba bottleneck". john hawks weblog.
  92. ^ Jones, Sacha. (2012). Local- and Regional-scale Impacts of the ~74 ka Toba Supervolcanic Eruption on Hominin Population and Habitats in India. Quaternary International 258: 100-118.
  93. ^ See also "Newly Discovered Archaeological Sites in India Reveals Ancient Life before Toba". Anthropology.net. 25 February 2010. Archived from the original on 22 July 2011. Retrieved 28 February 2010.
  94. ^ National Geographic- Did early humans in India survive a supervolcano?
  95. ^ Shea, John. (2008). Transitions or Turnovers? Climatically-forced Extinctions of Homo sapiens and Neanderthals in the East Mediterranean Levant. Quaternary Science Reviews 27: 2253-2270.
  96. ^ "Supervolcano Eruption In Sumatra Deforested India 73,000 Years ago". ScienceDaily. 24 November 2009.
  97. ^ Williams & others 2009.
  98. ^ Goldberg 1996
  99. ^ Steiper 2006
  100. ^ Hernandez & others 2007
  101. ^ Luo & others 2004

References[edit]

Further reading[edit]

External links[edit]