1257 Samalas eruption

The purple surface surrounded by white is the Samalas caldera.

The 1257 Samalas eruption was a major eruption of the Samalas volcano, next to Mount Rinjani on Lombok Island in Indonesia. The eruption left behind a large caldera next to Rinjani, with Lake Segara Anak inside it.[1] This eruption probably had a Volcanic Explosivity Index of 7, making it one of the largest eruptions of the current Holocene epoch.

An examination of ice cores showed a large spike in sulfate deposition around 1257. This was strong evidence of a large eruption having occurred somewhere in the world. In 2013, scientists proved that the eruption occurred at Mount Samalas.

This eruption had four distinct phases, alternately creating eruption columns reaching tens of kilometres into the atmosphere and pyroclastic flows burying large parts of Lombok Island. The flows destroyed human habitations, including the city of Pamatan. Ash from the eruption fell as far away as Java Island. The volcano deposited more than 10 cubic kilometres (2.4 cu mi) of material. The eruption was witnessed by people who recorded it on palm leaves, the Babad Lombok. Later volcanic activity created additional volcanic centres in the caldera, including the Barujari cone that remains active.

The aerosols injected into the atmosphere reduced the solar radiation reaching the Earth's surface, which cooled the atmosphere for several years and led to famines and crop failures in Europe and elsewhere, although the exact scale of the temperature anomalies and their consequences is still debated. It is possible that the eruption helped trigger the Little Ice Age.

Geology

General geology

Samalas and Mount Rinjani are in the Sunda Arc, a subduction zone where the Australian plate subducts beneath the Eurasian plate. The magmas feeding Samalas and Rinjani probably form from peridotites in the mantle wedge beneath Lombok Island.[2] Other volcanoes in the region include Agung and Batur plus Bratan on the island of Bali to the west.[3] Before the eruption, Samalas may have been as high as 4,200 ± 100 metres (13,780 ± 330 ft). This estimate is based on reconstructions such as continuing upwards from its surviving lower slopes.[4]

The oldest geological units on Lombok Island are from the Oligocene-Miocene,[5][6] with old volcanic units cropping out in southern Lombok.[7][5] Before 12,000 BP, volcanic activity built up the Samalas volcano. During a phase between 11,940±40 and 2,550±50 BP the Rinjani volcano formed;[6] this last eruption generated the Rinjani pumice with a volume of 0.3 cubic kilometres (0.072 cu mi) dense rock equivalent[8] and thicknesses of 6 centimetres (2.4 in) at 28 kilometres (17 mi) distance,[9] although later research suggests that this eruption may have occurred on Samalas instead.[10] Another eruption took place between 5,990 ± 50 and 2,550 ± 50 BP forming the Propok Pumice with a dense rock equivalent volume of 0.1 cubic kilometres (0.024 cu mi). More eruptions are dated 11,980±40, 11,940±40, 6250±40 BP,[11] and continued until about 500 years before 1257.[12] Then, a large caldera-forming eruption destroyed Samalas volcano. Later volcanic activity occurred in the Segara Anak caldera, forming the Segara Munac, Rombogan, and Barujari volcanoes.[13] Of these, most volcanic activity occurs in the Barujari volcano with eruptions in 1884, 1904, 1906, 1909, 1915, 1966, 1994, 2004, and 2009. Rombogan was active in 1944. Volcanic activity mostly consists of explosive eruptions and ash flows.[14]

The rocks of the volcano are mostly dacitic, SiO
2
contents range between 62–63 percent by weight.[6] Volcanic rocks in the Banda arc are mostly calc-alkaline ranging from basalt over andesite to dacite.[14] The volcano rests on crust of about 20 kilometres (12 mi) thickness, and the lower extremity of the Wadati–Benioff zone is about 164 kilometres (102 mi) deep.[2]

Eruption

The Segara Anak caldera, which was created by the eruption

The eruption of 1257 occurred probably during September.[15] Based on the deposits of the eruption, it commenced with a first phreatic (steam explosion powered) stage that deposited 3 centimetres (1.2 in) of ash over 400 square kilometres (150 sq mi) of northwest Lombok Island. In the subsequent first magmatic stage, lithic-rich pumice rained down, with the fallout reaching a thickness of 8 centimetres (3.1 in) both upwind on East Lombok and on Bali.[11] Subsequently, various phases of lapilli rock and ash fallout occurred, as well as pyroclastic flows that were partially confined within the valleys on Samalas's western flank. Some ash deposits were eroded by the pyroclastic flows, generating furrow structures in the ash deposits. Pyroclastic flows even crossed the Bali Sea, reaching the Gili Islands to the northwest. This eruption phase was probably phreatomagmatic, as the deposits show evidence of interaction of the lava with water. Three pumice fallout episodes subsequently occurred, which covered the widest extent of all deposits formed by the eruption.[16] These pumices fell as far as Sumbawa in the east, where they were up to 7 centimetres (2.8 in) thick.[17]

The emplacement of these pumices was followed by another stage of pyroclastic flow activity, probably caused by the collapse of the eruption column which generated the flows and the beginning of the formation of the caldera, where the conditions of the eruption changed from column-generating to fountain-generating. These pyroclastic flows were deflected by the topography as they flowed across the island incinerated the island's vegetation. Interaction between these flows and the air triggered the formation of additional eruption clouds and secondary pyroclastic flows. Where the flows entered the sea north and east of Lombok Island, steam explosions generated pumice cones on the beach and additional secondary pyroclastic flows.[17] These pyroclastic flows had an on-land volume of 29 cubic kilometres (7.0 cu mi),[18] and reached thicknesses of 35 metres (115 ft) as far away as 25 kilometres (16 mi) from Samalas.[19] These phases are also known as P1 (phreatic and first magmatic phase), P2 (phreatomagmatic with pyroclastic flows), P3 (plinian) and P4 (pyroclastic flows).[20] The whole eruption lasted between 12 and 15 hours, not accounting for the first emission of pyroclastic flows (P2).[21]

Tephra fragmental material from the eruption fell as far as Java, forming part of the so-called Muntilan Tephras which were found on the slopes of other volcanoes of Java, but could not be linked to eruptions to these volcanic systems.[22][17] These tephras reach thicknesses of 2–3 centimetres (0.79–1.18 in) on Mount Merapi, 15 centimetres (5.9 in) on Mount Bromo, 22 centimetres (8.7 in) at Ijen[23] and 12–17 centimetres (4.7–6.7 in) on Agung volcano on Bali. In Lake Logung on Java it was 3 centimetres (1.2 in) thick. Most of the tephra was deposited west-southwest of Samalas.[24] Considering the thickness of Samalas tephras found at Mount Merapi, the total volume may have reached 32–39 cubic kilometres (7.7–9.4 cu mi).[25] The dispersal index (the surface area covered by an ash or tephra fall) of the eruption reached 7,500 square kilometres (2,900 sq mi) during the first stage and 110,500 square kilometres (42,700 sq mi) during the third stage, implying that it was a Plinian eruption and an ultraplinian eruption respectively.[26]

The eruption column from this eruption reached a height of 39–40 kilometres (24–25 mi) at windspeeds of 10 metres per second (33 ft/s) during the first stage (P1),[27] and 43–38 kilometres (27–24 mi) between (P3) the two stages with emission of pyroclastic flows.[21] This eruption column was high enough that SO
2
in it and its S isotope ratio was influenced by photolysis at high altitudes.[28]

Various volumes have been estimated for the various stages of the Samalas eruption. The first stage reached a volume of 12.6–13.4 cubic kilometres (3.0–3.2 cu mi). The second phreatomagmatic phase has been estimated to have had a volume of 0.9–3.5 cubic kilometres (0.22–0.84 cu mi).[29] The total dense rock equivalent volume of the whole eruption was at least 40 cubic kilometres (9.6 cu mi).[26] The material erupted by the volcano had a temperature of about 1,000 °C (1,830 °F).[4] Compositionally, the material was of trachydacitic composition, containing amphibole, apatite, clinopyroxene, iron sulfide, orthopyroxene, plagioclase, and titanomagnetite. It formed out of basaltic magma by fractional crystallization.[30] Either the entry of new magma into the magma chamber or the effects of gas bubble buoyancy have been proposed as mechanisms that triggered the eruption.[31]

The eruption had a volcanic explosivity index of 7,[32] making it one of the largest eruptions of the current Holocene epoch.[33] It was stronger than the eruptions of Mount Tambora in 1815 and Krakatoa in 1883.[34] Eruptions of comparable intensity include the Kurile lake eruption in the 7th millennium BC, the Mount Mazama eruption in the 6th millennium BC, the Minoan eruption, and the Tierra Blanca Joven eruption of Lake Ilopango in the 6th century.[33] Pumice falls with a fine graining and colour of cream from the Samalas eruption form a useful tephrochronological marker on Bali.[35] Tephra from the volcano was found in ice cores as far as 13,500 kilometres (8,400 mi) away from Samalas,[36] and a tephra layer sampled at Dongdao island in the South China Sea has been tentatively linked to Samalas.[37] Ash and aerosols may have impacted humans and corals at large distances from the eruption.[38]

The eruption left the Segara Anak caldera, which has a diameter of 6–7 kilometres (3.7–4.3 mi), where the Samalas mountain was before;[13] before the eruption the mountain was probably about 4,193 ± 93 metres (13,757 ± 305 ft) high.[5] Within its 700–2,800 metres (2,300–9,200 ft) high walls, a 200 metres (660 ft) deep crater lake formed. The Barujari cone rises 320 metres (1,050 ft) above the water of the lake and has had 15 eruptions since 1847.[8] A crater lake may have already existed on Samalas before the eruption and supplied its phreatomagmatic phase with 0.1–0.3 cubic kilometres (0.024–0.072 cu mi) of water. Alternatively, the water could have come from aquifers.[39] A collapse structure cuts into Rinjani's slopes facing the Samalas caldera.[4]

The eruption that formed the caldera was first recognized in 2003, and in 2004 a volume of 10 cubic kilometres (2.4 cu mi) was attributed to this eruption.[11] Early research considered that the caldera-forming eruption occurred between 1210 and 1300. In 2013, Lavigne suggested that the eruption occurred in May–October 1257, resulting in the climate changes of 1258.[13] Presently, a number of villages on Lombok are constructed on the pyroclastic flow deposits from the 1257 event.[40]

Research history

A major volcanic event in 1257–1258 was identified from data in ice cores and from medieval records in the northern hemisphere.[41] The sulfur deposits in the polar ice caps had already showed that climate disturbances reported in that time were due to a volcanic event, with the global spread indicating a tropical volcano as the cause,[1] although at first a source in a volcano near Greenland had been considered.[42] The spike in sulfate concentrations was first observed in the Crête ice core,[43] associated with a deposition of rhyolitic ash.[44] These ice cores indicated a large sulfate spike around 1257, the largest in 7,000 years and twice the size of the 1815 eruption of Tambora sulfate spike,[45] but later-discovered sulfate spikes around 44 BC and 426 BC rival its size.[46] This discovery was made in the 1980s.[42] In 2003, a dense rock equivalent volume of 200–800 cubic kilometres (48–192 cu mi) was estimated for this eruption,[47] but it was also proposed that the eruption might have been somewhat smaller and enriched in sulfur.[48] The volcano responsible could not be identified at first;[41] Tofua volcano in Tonga was proposed at first but dismissed, as the Tofua eruption was too small to generate the 1257 sulfate spikes.[49] Likewise, a volcanic eruption in 1256 at Harrat al-Rahat near Medina was too small to trigger these events.[50] Other proposals included several simultaneous eruptions.[51] Estimated caldera diameters ranged 10–30 kilometres (6.2–18.6 mi).[52]

The suggestion that Samalas/Rinjani might be the source volcano was first made in 2012, since the other candidate volcanoes – El Chichon and Quilotoa – did not match the chemistry of the sulfur spikes.[53] El Chichon and Quilotoa and Okataina were also inconsistent with the timespan and size of the eruption.[54] The conclusive link between these events and an eruption of Samalas was made in 2013 on the basis of historical records in Indonesia: the Babad Lombok, writings in Old Javanese on palm leaves,[41] written in the 13th century, induced Franck Lavigne, who had already suspected that a volcano on Lombok may be responsible, to conclude that the Samalas volcano was responsible.[42]

All houses were destroyed and swept away, floating on the sea, and many people died

Javanese text, [55]

This event occurred before the end of the 13th century.[4] The role of the Samalas eruption in the global climate events was confirmed by comparing the geochemistry of glass shards found in ice cores to that of the eruption deposits on Lombok.[1]

Search for geological evidence of these events was helped by many exposures and cross-sections of volcanic ash deposits which were caused by men bulk-quarrying them for construction material.[56]

Climate effects

Ice cores in the northern and southern hemisphere display sulfate spikes associated with Samalas, the signal being the strongest in the southern hemisphere for the last 1000 years and being only exceeded by Laki's signal in the northern;[57] one reconstruction even considers it the strongest of the last 2500 years.[58] In addition, ice cores from Illimani in Bolivia contain sulfate spikes from the eruption.[59] For comparison, the 1991 eruption of Pinatubo ejected only about a tenth of Samalas's sulfur.[60] Sulfate deposition from the Samalas eruption has been noted at Svalbard,[61] and the fallout of sulfuric acid from the volcano may have directly affected peatlands in northern Sweden.[62] The amount of SO
2
released by the eruption has been estimated to be 158,000,000 ± 12,000,000 tonnes (156,000,000 ± 12,000,000 long tons; 174,000,000 ± 13,000,000 short tons);[30] the mass release was increased in comparison with the Tambora eruption due to a more effective injection of tephra into the stratosphere and/or higher sulfur contents of the Samalas magma.[63]

When large scale volcanic eruptions inject aerosols into the atmosphere, they can form stratospheric veils, which reduce the amount of light reaching the surface. That reduces the temperatures on much of the Earth and can cause problems in agriculture including famine. The social effects of such events, however, are often reduced by the resilience of humans.[64] Not all years with cold summers are linked to volcanic activity.[65] Volcanic eruptions can also deliver bromine and chlorine into the stratosphere, where they contribute to the breakdown of ozone. While most of them would have been removed by scavenging by the eruption column, the quantities that have been modelled for the Samalas halogen release (227,000,000 ± 18,000,000 tonnes (223,000,000 ± 18,000,000 long tons; 250,000,000 ± 20,000,000 short tons) of chlorine and up to 1,300,000 ± 300,000 tonnes (1,280,000 ± 300,000 long tons; 1,430,000 ± 330,000 short tons) of bromine) would have reduced the stratospheric ozone.[30]

Samalas, along with the Kuwae eruption in the 1450s and Tambora in 1815, was one of the strongest cooling events in the last millennium, even more so than at the peak of the Little Ice Age.[66] The effects of the Samalas eruption form the strongest volcanic signal in boreholes of the Urals.[67] After an early warm winter 1257–1258 (winter warming is frequently observed after tropical volcanic eruption), resulting in the early flowering of violets according to reports from France,[68] European summers were colder after the eruption,[69] and winters were long and cold.[70]

According to earlier reconstructions, summer cooling reached 0.69 K (1.24 °F) in the Southern Hemisphere and 0.46 K (0.83 °F) in the Northern Hemisphere.[71] More recent proxy data have indicated that a temperature drop of 0.7 °C (1.3 °F) occurred in 1258 and of 1.2 °C (2.2 °F) in 1259, but with differences between various geographical areas.[72] For comparison, the radiation forcing of Pinatubo's 1991 eruption was about a seventh of that of the Samalas eruption.[73] Sea surface temperatures likewise decreased by 0.3–2.2 °C (0.54–3.96 °F),[74] triggering changes in the ocean circulations and in the formation of deep water. Temperature changes may have lasted for a decade.[75] Precipitation and evaporation both decreased as well, but the decrease of evaporation was stronger.[76]

The Samalas signal, however, is only inconsistently reported from tree ring climate information,[77][78] and the temperature effects were likewise limited, probably because the large sulfate output altered the average size of particles and thus their radiation forcing.[79] Climate modelling indicated that the Samalas eruption may have reduced global temperatures by approximately 2 °C (3.6 °F), a value largely not replicated by proxy data. Better modelling indicated that the principal temperature anomaly occurred in 1258 and continued until 1261.[80] Climate models tend to overestimate the climate impact of a volcanic eruption;[81] one explanation is that climate models tend to assume that aerosol optical depth increases linearly with the quantity of erupted sulfur.[82] The possible occurrence of an El Nino before the eruption may have further reduced the cooling.[83]

The Samalas eruption together with another eruption in the 14th century set off a growth of ice caps and sea ice,[84] and glaciers in Norway advanced. It might also have modified the North Atlantic oscillation, causing it to acquire more negative values in the subsequent decades in co-operation with a beginning decrease in solar activity as part of the Wolf minimum in the solar cycle.[85] The advances of ice after the Samalas eruption may have strengthened and prolonged the climate effects.[62] Later volcanic activity in 1269, 1278, and 1286 and the effects of sea ice on the North Atlantic would have further contributed to ice expansion.[86] The glacier advances triggered by the Samalas eruption are documented on Baffin Island, where vegetation killed by the advancing ice was conserved in it.[87] Likewise, a change from a warm climate phase to a colder one, in Arctic Canada, coincides with the Samalas eruption.[88]

The Samalas eruption came after the Medieval Climate Anomaly[89] and at a time where a period of climate stability was ending, with prior eruptions in 1108, 1171, and 1230 already having upset global climate. Subsequent time periods displayed increased volcanic activity until the early 20th century.[90] The time period 1250–1300 was heavily disturbed by volcanic activity,[86] and is recorded by a moraine from a glacial advance on Disko Island,[91] although the moraine may indicate a pre-Samalas cold spell.[92] These volcanic disturbances along with positive feedback effects from increased ice may have started the Little Ice Age even without the need for changes in solar radiation,[93][94] but this theory is controversial.[95]

The eruption left traces, including decreased tree growth in Mongolia between 1258–1262 based on tree ring data,[96] and a very wet monsoon in Vietnam.[55] Another effect of the eruption-induced climate change may have been a brief decrease of atmospheric CO
2
.[51] Cooling may have lasted for 4–5 years based on simulations and tree ring data.[97]

Other regions such as Alaska were mostly unaffected.[98] There is little evidence that tree growth was affected in Fennoscandia, Quebec, and the Western United States.[99] In the case of Alaska, possibly the climate effect was moderated by the nearby ocean.[100] In 1259 on the other hand western Europe and the west coastal North America displayed mild weather.[72]

Among the effects estimated from analysis are the most negative Southern Annular Mode excursion of the last millennium,[101] onset of El Nino conditions during a climate period where La Nina was more common[93] as well as a weakening of the atlantic meridional overturning circulation which lasted long after the eruption, possibly aiding in the onset of the Little Ice Age as well.[102]

Social and historical consequences

This eruption led to global disaster in 1257–1258.[1]

Lombok Kingdom (Indonesia)

Western and central Indonesia were dominated by many kingdoms in competition with each other, and they often built temple complexes with inscriptions documenting certain kinds of historical events;[103] however very little direct historical evidence of the consequences of the Samalas eruption exists.[104] The Babad Lombok describe how villages on Lombok were destroyed during the middle 13th century by ash and high-speed sweeps of gas and rocks,[41] and is also the source of the name "Samalas".[7] The city of Pamatan, capital of a kingdom on Lombok, was destroyed along with the kingdom by the eruption and disappeared from history, although the royal family survived according to the Javanese text,[105] and there is no clear cut evidence that the kingdom itself was destroyed by the eruption.[104] Thousands of people died during the eruption.[4] Bali and Lombok Island may have been depopulated by the eruption,[106] possibly for generations, allowing Kertanegara to conquer Bali in 1284 with little resistance.[68]

Oceania

Historical events in Oceania are usually poorly dated and thus the timing and role of any specific event are difficult to assess. There is evidence however that between 1250-1300 crisis periods took place in Oceania such as at Easter Island, which may be linked with the beginning of the Little Ice Age and the Samalas eruption.[38]

Europe

The consequences of the Samalas eruption have been analyzed thanks to contemporary chronicles in Europe, which documented anomalous weather conditions in 1258.[107] Reports in 1258 in France and England indicate a dry fog, giving the impression of a persistent cloud cover to contemporary observers.[108] Medieval chronicles say that in 1258, the summer was cold and rainy, causing floods and bad harvests,[54] with cold from February to June.[109] In Europe and the Middle East, changes in atmospheric colours, storms, cold, and severe weather were reported in 1258–1259.[110] In Europe, excess rain and cold and high cloudiness damaged crops and caused famines followed by epidemics,[111][55] although 1258–1259 did not lead to famines as bad as some later ones like the Great Famine of 1315–17.[112] In northwest Europe, the effects included crop failure, famine, and weather changes.[84] Crop failures[1] and a famine in London have been linked to this event.[32] Witnesses reported a death toll of 15,000 to 20,000 in London. A mass burial of famine victims was found in the 1990s in the centre of London.[55] Matthew Paris of St Albans described how until mid-August in 1258, the weather alternated between cold and strong rain, causing high mortality.[113]

Swollen and rotting in groups of five or six, the dead lay abandoned in pigsties, on dunghills, and in the muddy streets.

Matthew Paris, chronicler of St. Albans, [113]

The resulting famine was severe enough that grain was imported from Germany and Holland.[114] The price for cereal increased in Britain,[110] France, and Italy. Outbreaks of disease occurred during this time in the Middle East and England.[115] With and after the winter of 1258–9, exceptional weathers were reported less commonly, but the winter of 1260–1 was very severe in Iceland, Italy, and elsewhere.[116]

Byzantine Empire

Potential long term consequences of the eruption were the Byzantine Empire losing control over western Anatolia, resulting from a shift in the political power from Byzantine farmers to mostly Turkoman pastoralists in the area.[117] The origins of the Flagellante movement may also be the social distress triggered by the eruption, but warfare and other plights may have contributed.[118]

Northeast Asia

Problems were also recorded in China, Japan, and Korea;[55] in Japan, the Mirror of the East chronicle from Azuma Kagami mentions that rice paddies and gardens were destroyed by the cold and wet weather,[119] and the Shoga famine may have been aggravated by bad weather in 1258 and 1259.[112] Other effects of the eruption include a total darkening of the Moon in May 1258 during a lunar eclipse.[120]

See also

References

  1. 1 2 3 4 5 Reid, Anthony (10 July 2016). "Revisiting Southeast Asian History with Geology: Some Demographic Consequences of a Dangerous Environment". In Bankoff, Greg; Christensen, Joseph. Natural Hazards and Peoples in the Indian Ocean World. Palgrave Macmillan US. p. 33. doi:10.1057/978-1-349-94857-4_2. ISBN 978-1-349-94857-4. Retrieved 1 October 2016.
  2. 1 2 Rachmat et al. 2016, p. 107.
  3. Fontijn et al. 2015, p. 2.
  4. 1 2 3 4 5 Lavigne et al. 2013, p. 16743.
  5. 1 2 3 Métrich et al. 2018, p. 4.
  6. 1 2 3 Rachmat et al. 2016, p. 108.
  7. 1 2 "Rinjani Dari Evolusi Kaldera hingga Geopark". Geomagz (in Indonesian). 4 April 2016. Retrieved 3 March 2018.
  8. 1 2 Vidal et al. 2015, p. 2.
  9. Métrich et al. 2018, p. 12.
  10. Métrich et al. 2018, p. 6.
  11. 1 2 3 Vidal et al. 2015, p. 3.
  12. Métrich et al. 2018, p. 10.
  13. 1 2 3 Rachmat et al. 2016, p. 109.
  14. 1 2 Rachmat et al. 2016, p. 110.
  15. Crowley, T. J.; Unterman, M. B. (23 May 2013). "Technical details concerning development of a 1200 yr proxy index for global volcanism". Earth System Science Data. 5 (1): 193. Bibcode:2013ESSD....5..187C. doi:10.5194/essd-5-187-2013.
  16. Vidal et al. 2015, p. 5.
  17. 1 2 3 Vidal et al. 2015, p. 7.
  18. Vidal et al. 2015, p. 17.
  19. Lavigne et al. 2013, p. 16744.
  20. Vidal et al. 2015, p. 21-22.
  21. 1 2 Vidal et al. 2015, p. 18.
  22. Alloway et al. 2017, p. 87.
  23. Alloway et al. 2017, p. 90.
  24. Vidal et al. 2015, p. 12.
  25. Vidal et al. 2015, p. 16.
  26. 1 2 Vidal et al. 2015, p. 19.
  27. Vidal et al. 2015, p. 17-18.
  28. Whitehill, A. R.; Jiang, B.; Guo, H.; Ono, S. (20 February 2015). "SO2 photolysis as a source for sulfur mass-independent isotope signatures in stratospehric aerosols". Atmospheric Chemistry and Physics. 15 (4): 1861. Bibcode:2015ACP....15.1843W. doi:10.5194/acp-15-1843-2015.
  29. Vidal et al. 2015, p. 14.
  30. 1 2 3 Vidal, Céline M.; Métrich, Nicole; Komorowski, Jean-Christophe; Pratomo, Indyo; Michel, Agnès; Kartadinata, Nugraha; Robert, Vincent; Lavigne, Franck (10 October 2016). "The 1257 Samalas eruption (Lombok, Indonesia): the single greatest stratospheric gas release of the Common Era". Scientific Reports. 6: 34868. Bibcode:2016NatSR...634868V. doi:10.1038/srep34868. PMC 5056521. PMID 27721477.
  31. Métrich et al. 2018, p. 29.
  32. 1 2 Whelley, Patrick L.; Newhall, Christopher G.; Bradley, Kyle E. (22 January 2015). "The frequency of explosive volcanic eruptions in Southeast Asia". Bulletin of Volcanology. 77 (1): 3. Bibcode:2015BVol...77....1W. doi:10.1007/s00445-014-0893-8.
  33. 1 2 Lavigne et al. 2013, p. 16745.
  34. Brata, Aloysius Gunadi; Rietveld, Piet; de Groot, Henri L.F.; Zant, Wouter (December 2013). "The Krakatau Eruption in 1883: Its Implications for the Spatial Distribution of Population in Java". Economic History of Developing Regions. 28 (2): 27–55. doi:10.1080/20780389.2013.866381.
  35. Fontijn et al. 2015, p. 8.
  36. Stevenson, J. A.; Millington, S. C.; Beckett, F. M.; Swindles, G. T.; Thordarson, T. (19 May 2015). "Big grains go far: understanding the discrepancy between tephrochronology and satellite infrared measurements of volcanic ash". Atmospheric Measurement Techniques. 8 (5): 2075. Bibcode:2015AMT.....8.2069S. doi:10.5194/amt-8-2069-2015.
  37. Yang, Zhongkang; Long, Nanye; Wang, Yuhong; Zhou, Xin; Liu, Yi; Sun, Liguang (1 February 2017). "A great volcanic eruption around AD 1300 recorded in lacustrine sediment from Dongdao Island, South China Sea". Journal of Earth System Science. 126 (1): 5. doi:10.1007/s12040-016-0790-y. ISSN 0253-4126.
  38. 1 2 Margalef et al. 2018, p. 5.
  39. Vidal et al. 2015, p. 14-15.
  40. Lavigne, Franck; Morin, Julie; Mei, Estuning Tyas Wulan; Calder, Eliza S.; Usamah, Muhi; Nugroho, Ute (2017). "Mapping Hazard Zones, Rapid Warning Communication and Understanding Communities: Primary Ways to Mitigate Pyroclastic Flow Hazard". SpringerLink. Advances in Volcanology. Springer, Berlin, Heidelberg: 4. doi:10.1007/11157_2016_34.
  41. 1 2 3 4 "ID'ed: Culprit Behind Medieval Eruption". Science. 342 (6154): 21–21. 3 October 2013. doi:10.1126/science.342.6154.21-b.
  42. 1 2 3 Hamilton 2013, p. 39.
  43. Oppenheimer 2003, p. 417.
  44. Oppenheimer 2003, p. 418.
  45. Auchmann, Renate; Brönnimann, Stefan; Arfeuille, Florian (March 2015). "Tambora: das Jahr ohne Sommer". Physik in unserer Zeit (in German). 46 (2): 67. Bibcode:2015PhuZ...46...64A. doi:10.1002/piuz.201401390.
  46. Sigl, M.; Winstrup, M.; McConnell, J. R.; Welten, K. C.; Plunkett, G.; Ludlow, F.; Büntgen, U.; Caffee, M.; Chellman, N.; Dahl-Jensen, D.; Fischer, H.; Kipfstuhl, S.; Kostick, C.; Maselli, O. J.; Mekhaldi, F.; Mulvaney, R.; Muscheler, R.; Pasteris, D. R.; Pilcher, J. R.; Salzer, M.; Schüpbach, S.; Steffensen, J. P.; Vinther, B. M.; Woodruff, T. E. (8 July 2015). "Timing and climate forcing of volcanic eruptions for the past 2,500 years". Nature. 523 (7562): 546. Bibcode:2015Natur.523..543S. doi:10.1038/nature14565. PMID 26153860.
  47. Oppenheimer 2003, p. 419.
  48. Oppenheimer 2003, p. 420.
  49. Caulfield, J. T.; Cronin, S. J.; Turner, S. P.; Cooper, L. B. (27 April 2011). "Mafic Plinian volcanism and ignimbrite emplacement at Tofua volcano, Tonga". Bulletin of Volcanology. 73 (9): 1274. Bibcode:2011BVol...73.1259C. doi:10.1007/s00445-011-0477-9.
  50. Stothers 2000, p. 361.
  51. 1 2 Brovkin et al. 2010, p. 675.
  52. Oppenheimer 2003, p. 424.
  53. Witze, Alexandra (14 July 2012). "Earth: Volcanic bromine destroyed ozone: Blasts emitted gas that erodes protective atmospheric layer". Science News. 182 (1): 12. doi:10.1002/scin.5591820114.
  54. 1 2 Lavigne et al. 2013, p. 16742.
  55. 1 2 3 4 5 Hamilton 2013, p. 40.
  56. More 4 television channel, 9–10 pm (British time), Volatile Earth: Killer Volcanoes, episode 1 of 3.
  57. Kokfelt et al. 2016, p. 2.
  58. Swingedouw et al. 2017, p. 28.
  59. Knüsel, S. (2003). "Dating of two nearby ice cores from the Illimani, Bolivia". Journal of Geophysical Research. 108 (D6): 4181. Bibcode:2003JGRD..108.4181K. doi:10.1029/2001JD002028.
  60. Fu et al. 2016, p. 2862.
  61. Wendl, I. A.; Eichler, A.; Isaksson, E.; Martma, T.; Schwikowski, M. (7 July 2015). "800-year ice-core record of nitrogen deposition in Svalbard linked to ocean productivity and biogenic emissions". Atmospheric Chemistry and Physics. 15 (13): 7290. Bibcode:2015ACP....15.7287W. doi:10.5194/acp-15-7287-2015.
  62. 1 2 Kokfelt et al. 2016, p. 6.
  63. Vidal et al. 2015, p. 21.
  64. Stothers 2000, p. 362.
  65. D'Arrigo, Rosanne; Wilson, Rob; Anchukaitis, Kevin J. (27 August 2013). "Volcanic cooling signal in tree ring temperature records for the past millennium". Journal of Geophysical Research: Atmospheres. 118 (16): 900–9002. Bibcode:2013JGRD..118.9000D. doi:10.1002/jgrd.50692.
  66. Neukom, Raphael; Gergis, Joëlle; Karoly, David J.; Wanner, Heinz; Curran, Mark; Elbert, Julie; González-Rouco, Fidel; Linsley, Braddock K.; Moy, Andrew D.; Mundo, Ignacio; Raible, Christoph C.; Steig, Eric J.; van Ommen, Tas; Vance, Tessa; Villalba, Ricardo; Zinke, Jens; Frank, David (30 March 2014). "Inter-hemispheric temperature variability over the past millennium". Nature Climate Change. 4 (5): 364. Bibcode:2014NatCC...4..362N. doi:10.1038/nclimate2174.
  67. Demezhko, D. Yu.; Gornostaeva, A. A. (24 December 2015). "Reconstructions of ground surface heat flux variations in the urals from geothermal and meteorological data". Izvestiya, Atmospheric and Oceanic Physics. 51 (7): 728. Bibcode:2015IzAOP..51..723D. doi:10.1134/S0001433815070026.
  68. 1 2 Lavigne et al. 2013, p. 16746.
  69. Luterbacher, J; Werner, J P; Smerdon, J E; Fernández-Donado, L; González-Rouco, F J; Barriopedro, D; Ljungqvist, F C; Büntgen, U; Zorita, E; Wagner, S; Esper, J; McCarroll, D; Toreti, A; Frank, D; Jungclaus, J H; Barriendos, M; Bertolin, C; Bothe, O; Brázdil, R; Camuffo, D; Dobrovolný, P; Gagen, M; García-Bustamante, E; Ge, Q; Gómez-Navarro, J J; Guiot, J; Hao, Z; Hegerl, G C; Holmgren, K; Klimenko, V V; Martín-Chivelet, J; Pfister, C; Roberts, N; Schindler, A; Schurer, A; Solomina, O; von Gunten, L; Wahl, E; Wanner, H; Wetter, O; Xoplaki, E; Yuan, N; Zanchettin, D; Zhang, H; Zerefos, C (1 February 2016). "European summer temperatures since Roman times". Environmental Research Letters. 11 (2): 8. doi:10.1088/1748-9326/11/2/024001.
  70. Hernández-Almeida, I.; Grosjean, M.; Przybylak, R.; Tylmann, W. (August 2015). "A chrysophyte-based quantitative reconstruction of winter severity from varved lake sediments in NE Poland during the past millennium and its relationship to natural climate variability". Quaternary Science Reviews. 122: 74–88. Bibcode:2015QSRv..122...74H. doi:10.1016/j.quascirev.2015.05.029.
  71. Oppenheimer 2003, p. 422.
  72. 1 2 Guillet et al. 2017, p. 126.
  73. Lim, Hyung-Gyu; Yeh, Sang-Wook; Kug, Jong-Seong; Park, Young-Gyu; Park, Jae-Hun; Park, Rokjin; Song, Chang-Keun (29 August 2015). "Threshold of the volcanic forcing that leads the El Niño-like warming in the last millennium: results from the ERIK simulation". Climate Dynamics. 46 (11–12): 3727. Bibcode:2016ClDy...46.3725L. doi:10.1007/s00382-015-2799-3.
  74. Chikamoto, Megumi O.; Timmermann, Axel; Yoshimori, Masakazu; Lehner, Flavio; Laurian, Audine; Abe-Ouchi, Ayako; Mouchet, Anne; Joos, Fortunat; Raible, Christoph C.; Cobb, Kim M. (16 February 2016). "Intensification of tropical Pacific biological productivity due to volcanic eruptions". Geophysical Research Letters. 43 (3): 1185. Bibcode:2016GeoRL..43.1184C. doi:10.1002/2015GL067359.
  75. Kim, Seong-Joong; Kim, Baek-Min (30 September 2012). "Ocean Response to the Pinatubo and 1259 Volcanic Eruptions". Ocean and Polar Research. 34 (3): 321. doi:10.4217/OPR.2012.34.3.305.
  76. Fu et al. 2016, p. 2859.
  77. Guillet et al. 2017, p. 123.
  78. Baillie, M. G. L.; McAneney, J. (16 January 2015). "Tree ring effects and ice core acidities clarify the volcanic record of the first millennium". Climate of the Past. 11 (1): 106. Bibcode:2015CliPa..11..105B. doi:10.5194/cp-11-105-2015.
  79. Boucher, Olivier (19 May 2015). "Stratospheric Aerosols". Atmospheric Aerosols. Springer Netherlands. p. 279. doi:10.1007/978-94-017-9649-1_12. ISBN 978-94-017-9649-1. Retrieved 2 October 2016.
  80. Guillet, Sebastien; Corona, Christophe; Stoffel, Markus; Khodri, Myriam; Poulain, Virginie; Guiot, Joel; Luckman, Brian; Churakova, Olga; Beniston, Martin; Franck, Lavigne; Masson-Delmotte, Valerie; Oppenheimer, Clive (2015). "Toward a more realistic assessment of the climatic impacts of the 1257 eruption". EGU General Assembly 2015. EGU General Assembly 2015. 17: 1268. Bibcode:2015EGUGA..17.1268G.
  81. Swingedouw et al. 2017, p. 30.
  82. Stoffel et al. 2015, p. 785.
  83. Timmreck et al. 2009, p. 3.
  84. 1 2 Brewington, Seth D. (May 2016). "6 The Social Costs of Resilience: An Example from the Faroe Islands". Archeological Papers of the American Anthropological Association. 27 (1): 99. doi:10.1111/apaa.12076.
  85. Faust, Johan C.; Fabian, Karl; Milzer, Gesa; Giraudeau, Jacques; Knies, Jochen (February 2016). "Norwegian fjord sediments reveal NAO related winter temperature and precipitation changes of the past 2800 years". Earth and Planetary Science Letters. 435: 91. Bibcode:2016E&PSL.435...84F. doi:10.1016/j.epsl.2015.12.003.
  86. 1 2 Zhong, Y.; Miller, G. H.; Otto-Bliesner, B. L.; Holland, M. M.; Bailey, D. A.; Schneider, D. P.; Geirsdottir, A. (31 December 2010). "Centennial-scale climate change from decadally-paced explosive volcanism: a coupled sea ice-ocean mechanism". Climate Dynamics. 37 (11–12): 2374–2375. Bibcode:2011ClDy...37.2373Z. doi:10.1007/s00382-010-0967-z.
  87. Robock, Alan (27 August 2013). "The Latest on Volcanic Eruptions and Climate". Eos, Transactions American Geophysical Union. 94 (35): 305–306. Bibcode:2013EOSTr..94..305R. doi:10.1002/2013EO350001.
  88. Gennaretti, F.; Arseneault, D.; Nicault, A.; Perreault, L.; Begin, Y. (30 June 2014). "Volcano-induced regime shifts in millennial tree-ring chronologies from northeastern North America". Proceedings of the National Academy of Sciences. 111 (28): 10077–10082. Bibcode:2014PNAS..11110077G. doi:10.1073/pnas.1324220111. PMC 4104845. PMID 24982132.
  89. Andres, Heather J.; Peltier, W. R. (August 2016). "Regional Influences of Natural External Forcings on the Transition from the Medieval Climate Anomaly to the Little Ice Age". Journal of Climate. 29 (16): 5783. Bibcode:2016JCli...29.5779A. doi:10.1175/JCLI-D-15-0599.1.
  90. Bradley, R. S.; Wanner, H.; Diaz, H. F. (22 January 2016). "The Medieval Quiet Period". The Holocene. 26 (6): 992. doi:10.1177/0959683615622552.
  91. Jomelli et al. 2016, p. 3.
  92. Jomelli et al. 2016, p. 5.
  93. 1 2 Margalef et al. 2018, p. 4.
  94. Miller, Gifford H.; Geirsdóttir, Áslaug; Zhong, Yafang; Larsen, Darren J.; Otto-Bliesner, Bette L.; Holland, Marika M.; Bailey, David A.; Refsnider, Kurt A.; Lehman, Scott J.; Southon, John R.; Anderson, Chance; Björnsson, Helgi; Thordarson, Thorvaldur (January 2012). "Abrupt onset of the Little Ice Age triggered by volcanism and sustained by sea-ice/ocean feedbacks". Geophysical Research Letters. 39 (2): n/a. Bibcode:2012GeoRL..39.2708M. doi:10.1029/2011GL050168.
  95. Naulier, M.; Savard, M. M.; Bégin, C.; Gennaretti, F.; Arseneault, D.; Marion, J.; Nicault, A.; Bégin, Y. (17 September 2015). "A millennial summer temperature reconstruction for northeastern Canada using oxygen isotopes in subfossil trees". Climate of the Past. 11 (9): 1160. Bibcode:2015CliPa..11.1153N. doi:10.5194/cp-11-1153-2015.
  96. Davi, N.K.; D'Arrigo, R.; Jacoby, G.C.; Cook, E.R.; Anchukaitis, K.J.; Nachin, B.; Rao, M.P.; Leland, C. (August 2015). "A long-term context (931–2005 C.E.) for rapid warming over Central Asia". Quaternary Science Reviews. 121: 95. Bibcode:2015QSRv..121...89D. doi:10.1016/j.quascirev.2015.05.020.
  97. Stoffel et al. 2015, p. 787.
  98. Guillet, Sebastien; Corona, Christophe; Stoffel, Markus; Khodri, Myriam; Poulain, Virginie; Lavigne, Franck; Churakova, Olga; Ortega, Pablo; Daux, Valerie; Luckman, Brian; Guiot, Joel; Oppenheimer, Clive; Masson-Delmotte, Valérie; Edouard, Jean-Louis (2016). "Reassessing the climatic impacts of the AD 1257 Samalas eruption in Europe and in the Northern Hemisphere using historical archives and tree-rings". EGU General Assembly 2016 – SAO/NASA ADS Physics Abstract Service. EGU General Assembly 2016. 18: 15250. Bibcode:2016EGUGA..1815250G.
  99. D'Arrigo, Rosanne; Frank, David; Jacoby, Gordon; Pederson, Neil (2001). "Spatial Response to Major Volcanic Events in or about AD 536, 934 and 1258: Frost Rings and Other Dendrochronological Evidence from Mongolia and Northern Siberia: Comment on R. B. Stothers, 'Volcanic Dry Fogs, Climate Cooling, and Plague Pandemics in Europe and the Middle East' (Climatic Change, 42, 1999)". Climatic Change. 49 (1/2): 243. doi:10.1023/A:1010727122905.
  100. Schneider, David P.; Ammann, Caspar M.; Otto-Bliesner, Bette L.; Kaufman, Darrell S. (1 August 2009). "Climate response to large, high-latitude and low-latitude volcanic eruptions in the Community Climate System Model". Journal of Geophysical Research. 114 (D15): 19. Bibcode:2009JGRD..11415101S. doi:10.1029/2008JD011222.
  101. Dätwyler, Christoph; Neukom, Raphael; Abram, Nerilie J.; Gallant, Ailie J. E.; Grosjean, Martin; Jacques-Coper, Martín; Karoly, David J.; Villalba, Ricardo (30 November 2017). "Teleconnection stationarity, variability and trends of the Southern Annular Mode (SAM) during the last millennium". Climate Dynamics: 1–19. doi:10.1007/s00382-017-4015-0. ISSN 0930-7575.
  102. Swingedouw et al. 2017, p. 41.
  103. Alloway et al. 2017, p. 86.
  104. 1 2 Alloway et al. 2017, p. 98.
  105. Hamilton 2013, p. 41.
  106. Reid, Anthony (2016). "Building Cities in a Subduction Zone: Some Indonesian Dangers". In Miller, Michelle Ann; Douglass, Mike. Disaster Governance in Urbanising Asia. Springer Singapore. p. 51. doi:10.1007/978-981-287-649-2_3. ISBN 978-981-287-649-2. Retrieved 1 October 2016.
  107. Ludlow, Francis (2017). "Volcanology: Chronicling a medieval eruption". Nature Geoscience. 10 (2): 78–79. doi:10.1038/ngeo2881. ISSN 1752-0908.
  108. Stothers 2000, p. 363.
  109. D'Arrigo, Rosanne; Jacoby, Gordon; Frank, David (2003). "Dendroclimatological evidence for major volcanic events of the past two millennia". Volcanism and the Earth's Atmosphere: Dendroclimatological evidence for major volcanic events of the past two millennia. Geophysical Monograph Series. 139. p. 259. Bibcode:2003GMS...139..255D. doi:10.1029/139GM16. ISBN 0-87590-998-1.
  110. 1 2 Dodds & Liddy 2011, p. 54.
  111. Guillet et al. 2017, p. 124.
  112. 1 2 Guillet et al. 2017, p. 127.
  113. 1 2 John Gillingham (30 October 2014). Conquests, Catastrophe and Recovery: Britain and Ireland 1066–1485. Random House. p. 26. ISBN 978-1-4735-2233-6.
  114. Speed, Robert; Tickner, David; Lei, Gang; Sayers, Paul; Wei, Yu; Li, Yuanyuan; Moncrieff, Catherine; Pegram, Guy (19 September 2016). Drought risk management: a strategic approach. UNESCO Publishing. p. 44. ISBN 978-92-3-100094-2.
  115. Stothers 2000, p. 366.
  116. Stothers 2000, p. 364.
  117. Xoplaki, Elena; Fleitmann, Dominik; Luterbacher, Juerg; Wagner, Sebastian; Haldon, John F.; Zorita, Eduardo; Telelis, Ioannis; Toreti, Andrea; Izdebski, Adam (March 2016). "The Medieval Climate Anomaly and Byzantium: A review of the evidence on climatic fluctuations, economic performance and societal change". Quaternary Science Reviews. 136: 229–252. Bibcode:2016QSRv..136..229X. doi:10.1016/j.quascirev.2015.10.004.
  118. Stothers 2000, pp. 367–368.
  119. Guillet et al. 2017, p. 125.
  120. Timmreck et al. 2009, p. 1.

Sources

  • Alloway, Brent V.; Andreastuti, Supriyati; Setiawan, Ruly; Miksic, John; Hua, Quan (January 2017). "Archaeological implications of a widespread 13th Century tephra marker across the central Indonesian Archipelago". Quaternary Science Reviews. 155: 86–99. doi:10.1016/j.quascirev.2016.11.020. ISSN 0277-3791.
  • Brovkin, Victor; Lorenz, Stephan J.; Jungclaus, Johann; Raddatz, Thomas; Timmreck, Claudia; Reick, Christian H.; Segschneider, Joachim; Sis, Katharina (November 2010). "Sensitivity of a coupled climate-carbon cycle model to large volcanic eruptions during the last millennium". Tellus B. 62 (5): 674–681. Bibcode:2010TellB..62..674B. doi:10.1111/j.1600-0889.2010.00471.x.
  • Dodds, Ben; Liddy, Christian D. (2011). Commercial Activity, Markets and Entrepreneurs in the Middle Ages: Essays in Honour of Richard Britnell. Boydell & Brewer Ltd. ISBN 978-1-84383-684-1.
  • Fontijn, Karen; Costa, Fidel; Sutawidjaja, Igan; Newhall, Christopher G.; Herrin, Jason S. (10 June 2015). "A 5000-year record of multiple highly explosive mafic eruptions from Gunung Agung (Bali, Indonesia): implications for eruption frequency and volcanic hazards". Bulletin of Volcanology. 77 (7): 59. Bibcode:2015BVol...77...59F. doi:10.1007/s00445-015-0943-x.
  • Fu, Qiang; Lin, Lei; Huang, Jianping; Feng, Song; Gettelman, Andrew (27 March 2016). "Changes in terrestrial aridity for the period 850–2080 from the Community Earth System Model". Journal of Geophysical Research: Atmospheres. 121 (6): 2857–2873. Bibcode:2016JGRD..121.2857F. doi:10.1002/2015JD024075.
  • Guillet, Sébastien; Corona, Christophe; Stoffel, Markus; Khodri, Myriam; Lavigne, Franck; Ortega, Pablo; Eckert, Nicolas; Sielenou, Pascal Dkengne; Daux, Valérie; (Sidorova), Olga V. Churakova; Davi, Nicole; Edouard, Jean-Louis; Zhang, Yong; Luckman, Brian H.; Myglan, Vladimir S.; Guiot, Joël; Beniston, Martin; Masson-Delmotte, Valérie; Oppenheimer, Clive (2017). "Climate response to the Samalas volcanic eruption in 1257 revealed by proxy records". Nature Geoscience. 10 (2): 123–128. doi:10.1038/ngeo2875. ISSN 1752-0908.
  • Hamilton, Garry (October 2013). "Mystery blast: The lost volcano that changed the world". New Scientist. 220 (2939): 38–41. doi:10.1016/S0262-4079(13)62487-2.
  • Jomelli, Vincent; Lane, Timothy; Favier, Vincent; Masson-Delmotte, Valerie; Swingedouw, Didier; Rinterknecht, Vincent; Schimmelpfennig, Irene; Brunstein, Daniel; Verfaillie, Deborah; Adamson, Kathryn; Leanni, Laëtitia; Mokadem, Fatima; Aumaître, Georges; Bourlès, Didier L.; Keddadouche, Karim (9 September 2016). "Paradoxical cold conditions during the medieval climate anomaly in the Western Arctic". Scientific Reports. 6: 32984. Bibcode:2016NatSR...632984.. doi:10.1038/srep32984. PMC 5016737. PMID 27609585.
  • Kokfelt, U.; Muscheler, R.; Mellström, A.; Struyf, E.; Rundgren, M.; Wastegård, S.; Hammarlund, D. (September 2016). "Diatom blooms and associated vegetation shifts in a subarctic peatland: responses to distant volcanic eruptions?". Journal of Quaternary Science. 31 (7): 723–730. doi:10.1002/jqs.2898.
  • Lavigne, F.; Degeai, J.-P.; Komorowski, J.-C.; Guillet, S.; Robert, V.; Lahitte, P.; Oppenheimer, C.; Stoffel, M.; Vidal, C. M.; Surono; Pratomo, I.; Wassmer, P.; Hajdas, I.; Hadmoko, D. S.; de Belizal, E. (30 September 2013). "Source of the great A.D. 1257 mystery eruption unveiled, Samalas volcano, Rinjani Volcanic Complex, Indonesia". Proceedings of the National Academy of Sciences. 110 (42): 16742–16747. Bibcode:2013PNAS..11016742L. doi:10.1073/pnas.1307520110. PMC 3801080. PMID 24082132.
  • Margalef, Olga; Álvarez-Gómez, José A.; Pla-Rabes, Sergi; Cañellas-Boltà, Núria; Rull, Valentí; Sáez, Alberto; Geyer, Adelina; Peñuelas, Josep; Sardans, Jordi; Giralt, Santiago (2 May 2018). "Revisiting the role of high-energy Pacific events in the environmental and cultural history of Easter Island (Rapa Nui)". The Geographical Journal. doi:10.1111/geoj.12253. ISSN 0016-7398.
  • Métrich, Nicole; Vidal, Céline M.; Komorowski, Jean-Christophe; Pratomo, Indyo; Michel, Agnès; Kartadinata, Nugraha; Prambada, Oktory; Rachmat, Heryadi; Surono, Surono (3 February 2018). "New Insights into magma differentiation and storage in Holocene crustal reservoirs of the Lesser Sunda Arc: the Rinjani-Samalas volcanic complex (Lombok, Indonesia)". Journal of Petrology. doi:10.1093/petrology/egy006.
  • Oppenheimer, Clive (30 March 2003). "Ice core and palaeoclimatic evidence for the timing and nature of the great mid-13th century volcanic eruption". International Journal of Climatology. 23 (4): 417–426. Bibcode:2003IJCli..23..417O. doi:10.1002/joc.891.
  • Rachmat, Heryadi; Rosana, Mega Fatimah; Wirakusumah, Ade Djumarma; Jabbar, Gamma Abdul (2 August 2016). "Petrogenesis of Rinjani Post-1257-Caldera-Forming-Eruption Lava Flows". Indonesian Journal on Geoscience. 3 (2). doi:10.17014/ijog.3.2.107-126.
  • Stoffel, Markus; Khodri, Myriam; Corona, Christophe; Guillet, Sébastien; Poulain, Virginie; Bekki, Slimane; Guiot, Joël; Luckman, Brian H.; Oppenheimer, Clive; Lebas, Nicolas; Beniston, Martin; Masson-Delmotte, Valérie (31 August 2015). "Estimates of volcanic-induced cooling in the Northern Hemisphere over the past 1,500 years". Nature Geoscience. 8 (10): 784–788. Bibcode:2015NatGe...8..784S. doi:10.1038/ngeo2526.
  • Stothers, Richard B. (2000). "Climatic and Demographic Consequences of the Massive Volcanic Eruption of 1258". Climatic Change. 45 (2): 361–374. doi:10.1023/A:1005523330643.
  • Swingedouw, Didier; Mignot, Juliette; Ortega, Pablo; Khodri, Myriam; Menegoz, Martin; Cassou, Christophe; Hanquiez, Vincent (March 2017). "Impact of explosive volcanic eruptions on the main climate variability modes". Global and Planetary Change. 150: 24–45. doi:10.1016/j.gloplacha.2017.01.006. ISSN 0921-8181.
  • Timmreck, Claudia; Lorenz, Stephan J.; Crowley, Thomas J.; Kinne, Stefan; Raddatz, Thomas J.; Thomas, Manu A.; Jungclaus, Johann H. (6 November 2009). "Limited temperature response to the very large AD 1258 volcanic eruption". Geophysical Research Letters. 36 (21): L21708. Bibcode:2009GeoRL..3621708T. doi:10.1029/2009GL040083.
  • Vidal, Céline M.; Komorowski, Jean-Christophe; Métrich, Nicole; Pratomo, Indyo; Kartadinata, Nugraha; Prambada, Oktory; Michel, Agnès; Carazzo, Guillaume; Lavigne, Franck; Rodysill, Jessica; Fontijn, Karen; Surono (8 August 2015). "Dynamics of the major plinian eruption of Samalas in 1257 A.D. (Lombok, Indonesia)". Bulletin of Volcanology. 77 (9): 1–24. Bibcode:2015BVol...77...73V. doi:10.1007/s00445-015-0960-9.

Coordinates: 8°24′36″S 116°24′30″E / 8.41000°S 116.40833°E / -8.41000; 116.40833

This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.