1257 Samalas eruption

Coordinates: 8°24′36″S 116°24′30″E / 8.41000°S 116.40833°E / -8.41000; 116.40833
Source: Wikipedia, the free encyclopedia.

1257 Samalas eruption
View of Mount Samalas along with Mount Rinjani
VolcanoSamalas
Date1257
TypeUltra-Plinian
LocationLombok, Indonesia
8°24′36″S 116°24′30″E / 8.41000°S 116.40833°E / -8.41000; 116.40833
VEI7[1]
The volcano-caldera complex in the north of Lombok

In 1257, a catastrophic eruption occurred at Samalas, a

Volcanic Explosivity Index of 7,[a] making it one of the largest volcanic eruptions during the Holocene epoch. It left behind a large caldera that contains Lake Segara Anak
. Later volcanic activity created more volcanic centres in the caldera, including the Barujari cone, which remains active.

The event created eruption columns reaching tens of kilometres into the atmosphere and pyroclastic flows that buried much of Lombok and crossed the sea to reach the neighbouring island of Sumbawa. The flows destroyed human habitations, including the city of Pamatan, which was the capital of a kingdom on Lombok. Ash from the eruption fell as far as 340 kilometres (210 mi) away in Java; the volcano deposited more than 10 cubic kilometres (2.4 cu mi) of rocks and ash.

The aerosols injected into the atmosphere reduced the solar radiation reaching the Earth's surface, causing a volcanic winter and cooling the atmosphere for several years. This led to famines and crop failures in Europe and elsewhere, although the exact scale of the temperature anomalies and their consequences is still debated. The eruption may have helped trigger the Little Ice Age, a centuries-long cold period during the last thousand years.

Before the site of the eruption was known, an examination of ice cores around the world had detected a large spike in sulfate deposition from around 1257 providing strong evidence of a large volcanic eruption occurring at that time. In 2013, scientists linked the historical records about Mount Samalas to these spikes. These records were written by people who witnessed the event and recorded it on the Babad Lombok, a document written on palm leaves.

Geology

Samalas (also known as Rinjani Tua[4]) was part of what is now the Rinjani volcanic complex, on Lombok, in Indonesia.[5] The remains of the volcano form the Segara Anak caldera, with Mount Rinjani at its eastern edge.[4] Since the destruction of Samalas, two new volcanoes, Rombongan and Barujari, have formed in the caldera. Mount Rinjani has also been volcanically active, forming its own crater, Segara Muncar.[6] Other volcanoes in the region include Agung, Batur, and Bratan, on the island of Bali to the west.[7]

Location of Lombok

Lombok is one of the

Eurasian plate[9] at a rate of 7 centimetres per year (2.8 in/year).[11] The magmas feeding Mount Samalas and Mount Rinjani are likely derived from peridotite rocks beneath Lombok, in the mantle wedge.[9] Before the eruption, Mount Samalas may have been as tall as 4,200 ± 100 metres (13,780 ± 330 ft), based on reconstructions that extrapolate upwards from the surviving lower slopes,[12] and thus taller than Mount Kinabalu which is presently the highest mountain in tropical Asia;[13] Samalas's current height is less than that of the neighbouring Mount Rinjani, which reaches 3,726 metres (12,224 ft).[12]

The oldest geological units on Lombok are from the OligoceneMiocene,[5][10] with old volcanic units cropping out in southern parts of the island.[4][5] Samalas was built up by volcanic activity before 12,000 BP. Rinjani formed between 11,940 ± 40 and 2,550 ± 50 BP,[10] with an eruption 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).[14] The Rinjani Pumice, with a volume of 0.3 cubic kilometres (0.072 cu mi) dense rock equivalent,[15][b] may have been deposited by an eruption from either Rinjani or Samalas;[17] it is dated to 2,550 ± 50 BP,[15] at the end of the time range during which Rinjani formed.[10] The deposits from this eruption reached thicknesses of 6 centimetres (2.4 in) 28 kilometres (17 mi) away.[18] Additional eruptions by either Rinjani or Samalas are dated 11,980 ± 40, 11,940 ± 40, and 6,250 ± 40 BP.[14] Eruptive activity continued until about 500 years before 1257.[19] Most volcanic activity now occurs at the Barujari volcano with eruptions in 1884, 1904, 1906, 1909, 1915, 1966, 1994, 2004, and 2009; Rombongan was active in 1944. Volcanic activity mostly consists of explosive eruptions and ash flows.[20]

The rocks of the Samalas volcano are mostly

calc-alkaline ranging from basalt over andesite to dacite.[20] The crust beneath the volcano is about 20 kilometres (12 mi) thick, and the lower extremity of the Wadati–Benioff zone is about 164 kilometres (102 mi) deep.[9]

Eruption

Segara Anak
caldera, which was created by the eruption

The events of the 1257 eruption have been reconstructed through geological analysis of the deposits it left[14] and by historical records.[21] The eruption probably occurred during the northern summer[22] in September (uncertainty of 2–3 months) that year, in light of the time it would have taken for its traces to reach the polar ice sheets and be recorded in ice cores[23] and the pattern of tephra deposits.[22] 1257 is the most likely year of the eruption, although a date of 1258 is also possible.[24]

Phases

The phases of the eruption are also known as P1 (phreatic and magmatic phase), P2 (phreatomagmatic with pyroclastic flows), P3 (

photolysis at high altitudes.[28]

Event

The eruption began with a

phreatomagmatic. It was followed by three pumice fallout episodes, with deposits over an area wider than was reached by any of the other eruption phases.[29] These pumices fell up to 61 kilometres (38 mi) to the east, against the prevailing wind, in Sumbawa, where they are up to 7 centimetres (2.8 in) thick.[31]

The deposition of these pumices was followed by another stage of pyroclastic flow activity, probably caused by the collapse of the eruption column that generated the flows. At this time the eruption changed from an eruption-column-generating stage to a fountain-like stage and the caldera began to form. These pyroclastic flows were deflected by the topography of Lombok, filling valleys and moving around obstacles such as older volcanoes as they expanded across the island incinerating 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, steam explosions created pumice cones on the beaches and additional secondary pyroclastic flows.[31]

river valleys; a new river network developed on the volcanic deposits after the eruption.[37]

Rock and ash

Volcanic rocks ejected by the eruption covered Bali and Lombok and parts of Sumbawa.

Pumice falls with a fine graining and creamy colour from the Samalas eruption have been used as a tephrochronological[c] marker on Bali.[45] Tephra from the volcano was found in ice cores as far as 13,500 kilometres (8,400 mi) away,[46] and a tephra layer sampled at Dongdao island in the South China Sea has been tentatively linked to Samalas.[47] Ash and aerosols might have impacted humans and corals at large distances from the eruption.[48]

There are several estimates of the volumes expelled during 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 phreatomagmatic phase has been estimated to have had a volume of 0.9–3.5 cubic kilometres (0.22–0.84 cu mi).

basaltic magma by fractional crystallization[50] and had a temperature of about 1,000 °C (1,830 °F).[12] Its eruption may have been triggered either by the entry of new magma into the magma chamber or the effects of gas bubble buoyancy.[51]

Intensity

The eruption had a

Mt. Tambora in 1815.[53] Such large volcanic eruptions can result in catastrophic impacts on humans and widespread loss of life both close to the volcano and at greater distances.[56]

Caldera

The eruption created the 6–7 kilometres (3.7–4.3 mi) wide Segara Anak caldera where the Samalas mountain was formerly located;[6] within its 700–2,800 metres (2,300–9,200 ft) high walls, a 200 metres (660 ft) deep crater lake formed[15] called Lake Segara Anak.[57] The Barujari cone rises 320 metres (1,050 ft) above the water of the lake and has erupted 15 times since 1847.[15] A crater lake may have 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 been supplied by aquifers.[58] Approximately 2.1–2.9 cubic kilometres (0.50–0.70 cu mi) of rock from Rinjani fell into the caldera,[59] a collapse that was witnessed by humans[21] and left a collapse structure that cuts into Rinjani's slopes facing the Samalas caldera.[12]

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.

climate changes of 1258.[6] Several villages on Lombok are constructed on the pyroclastic flow deposits from the 1257 event.[60]

Research history

A major volcanic event in 1257–1258 was first discovered from data in ice cores;

stratigraphic markers for ice cores even before the volcano that caused them was known.[72]

The ice cores indicated a large sulfate spike, accompanied by tephra deposition,[73] around 1257–1259,[74][73] the largest[d] in 7,000 years and twice the size of the spike due to the 1815 eruption of Tambora.[74] In 2003, a dense rock equivalent volume of 200–800 cubic kilometres (48–192 cu mi) was estimated for this eruption,[76] but it was also proposed that the eruption might have been somewhat smaller and richer in sulfur.[77][61] The volcano responsible was thought to be located in the Ring of Fire[78] but could not be identified at first;[62] Tofua volcano in Tonga was proposed at first but dismissed, as the Tofua eruption was too small to generate the 1257 sulfate spikes.[79] A volcanic eruption in 1256 at Harrat al-Rahat near Medina was also too small to trigger these events.[80] Other proposals included several simultaneous eruptions.[81] The diameter of the caldera left by the eruption was estimated to be 10–30 kilometres (6.2–18.6 mi),[82] and the location was estimated to be close to the equator and probably north of it.[83]

While at first no clear-cut climate anomaly could be correlated to the 1257 sulfate layers,[84][85] in 2000[84] climate phenomena were identified in medieval records of the northern hemisphere[62][63] that are characteristic for volcanic eruptions.[64] Earlier, climate alterations had been reported from studies of tree rings and climate reconstructions.[84] The deposits showed that climate disturbances reported at that time were due to a volcanic event, the global spread indicating a tropical volcano as the cause.[57]

The suggestion that Samalas/Rinjani might be the source volcano was first raised in 2012, since the other candidate volcanoes—El Chichón and Quilotoa—did not match the chemistry of the sulfur spikes.[86] El Chichon, Quilotoa and Okataina were also inconsistent with the timespan and size of the eruption.[63]

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

Babad Lombok[87]

The conclusive link between these events and an eruption of Samalas was made in 2013 on the basis of

geochemical similarities between tephra found in polar ice cores and eruption products of Samalas reinforced this localization.[90][91]

Climate effects

Aerosol and paleoclimate data

Ice cores in the northern and southern hemisphere display sulfate spikes associated with Samalas. The signal is the strongest in the southern hemisphere over the last 1000 years;

Krakatau.[64] In the northern hemisphere it is only exceeded by the signal of the destructive 1783/1784 Laki eruption.[92] The ice core sulfate spikes have been used as a time marker in chronostratigraphic studies.[94] Ice cores from Illimani in Bolivia contain thallium[95] and sulfate spikes from the eruption.[96] For comparison, the 1991 eruption of Pinatubo ejected only about a tenth of the amount of sulfur erupted by Samalas.[97] Sulfate deposition from the Samalas eruption has been noted at Svalbard,[98] and the fallout of sulfuric acid from the volcano may have directly affected peatlands in northern Sweden.[99]

In addition, the sulfate aerosols may have extracted large amounts of the

stratospheric veils. These reduce the amount of light reaching the surface and cause lower temperatures, which can lead to poor crop yields.[104] Such sulfate aerosols in the case of the Samalas eruption may have remained at high concentrations for about three years according to findings in the Dome C ice core in Antarctica, although a smaller amount may have persisted for an additional time.[105]

Other records of the eruption's impact include decreased tree growth in Mongolia between 1258 and 1262 based on tree ring data,

Korean Peninsula[110] and in lake sediments of northeastern China,[111] a very wet monsoon in Vietnam,[88] droughts in many places in the Northern Hemisphere[112] as well as in southern Thailand cave records,[e][113] and a decade-long thinning of tree rings in Norway and Sweden.[114] Cooling may have lasted for 4–5 years based on simulations and tree ring data.[115]

Another effect of the eruption-induced climate change may have been a brief decrease in atmospheric carbon dioxide concentrations.[81] A decrease in the growth rate of atmospheric carbon dioxide concentrations was recorded after the 1992 Pinatubo eruption; several mechanisms for volcanically driven decreases in atmospheric CO
2
concentration have been proposed, including colder oceans absorbing extra CO
2
and releasing less of it, decreased respiration rates leading to carbon accumulation in the biosphere,[116] and increased productivity of the biosphere due to increased scattered sunlight and the fertilization of oceans by volcanic ash.[117]

The Samalas signal is only inconsistently reported from

El Niño before the eruption may have further reduced the cooling.[126]

The Samalas eruption, together with 14th century cooling, is thought to have set off a growth of ice caps and

Arctic Canada from a warm climate phase to a colder one coincides with the Samalas eruption.[132]

Simulated effects

According to 2003 reconstructions, summer cooling reached 0.69 °C (1.24 °F) in the southern hemisphere and 0.46 °C (0.83 °F) in the northern hemisphere.[84] More recent proxy data indicate 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.[133] For comparison, the radiative forcing of Pinatubo's 1991 eruption was about a seventh of that of the Samalas eruption.[134] Sea surface temperatures too decreased by 0.3–2.2 °C (0.54–3.96 °F),[135] triggering changes in the ocean circulations. Ocean temperature and salinity changes may have lasted for a decade.[136] Precipitation and evaporation both decreased, evaporation reduced more than precipitation.[137]

Volcanic eruptions can also deliver bromine and chlorine into the stratosphere, where they contribute to the breakdown of

ultraviolet radiation on the surface of Earth may have led to widespread immunosuppression in human populations, explaining the onset of epidemics in the years following the eruption.[139]

Climate effects in various areas

Samalas, along with the 1452/1453 mystery eruption and the 1815 eruption of Mount Tambora, was one of the strongest cooling events in the last millennium, even more so than at the peak of the Little Ice Age.[140] After an early warm winter 1257–1258[f][141] resulting in the early flowering of violets according to reports from the Kingdom of France,[142] European summers were colder after the eruption,[144] and winters were long and cold.[145]

The Samalas eruption came after the

Medieval Climate Anomaly,[146] a period early in the last millennium with unusually warm temperatures,[147] and at a time when a period of climate stability was ending, with earlier eruptions in 1108, 1171, and 1230 already having upset global climate. Subsequent time periods displayed increased volcanic activity until the early 20th century.[148] The time period 1250–1300 was heavily disturbed by volcanic activity[130] from four eruptions in 1230, 1257, 1276 and 1286,[149] and is recorded by a moraine from a glacial advance on Disko Island,[150] although the moraine may indicate a pre-Samalas cold spell.[151] These volcanic disturbances along with positive feedback effects from increased ice may have started the Little Ice Age[g] even without the need for changes in solar radiation,[153][154] though this theory is not without disagreement.[155] The Samalas eruption in Europe is sometimes used as a chronological marker for the beginning of the Little Ice Age.[156]

Other inferred effects of the eruption are:

Other regions such as Alaska were mostly unaffected.[183] There is little evidence that tree growth was influenced by cold in what is now the Western United States,[184] where the eruption may have interrupted a prolonged drought period.[185] The climate effect in Alaska may have been moderated by the nearby ocean.[186] In 1259, Western Europe and the west coastal North America had mild weather[133] and there is no evidence for summer precipitation changes in Central Europe.[187] Tree rings do not show much evidence of precipitation changes.[188]

Social and historical consequences

The eruption led to global disaster in 1257–1258.[57] Very large volcanic eruptions can cause significant human hardship, including famine, away from the volcano due to their effect on climate. The social effects are often reduced by the resilience of humans; thus there is often uncertainty about causal links between volcano-induced climate variations and societal changes at the same time.[104]

Lombok Kingdom and Bali (Indonesia)

Western and central Indonesia at the time were divided into competing kingdoms that often built temple complexes with inscriptions documenting historical events.[56] However, little direct historical evidence of the consequences of the Samalas eruption exists.[189] The Babad Lombok describe how villages on Lombok were destroyed during the mid-13th century by ash, gas and lava flows,[62] and two additional documents known as the Babad Sembalun and Babad Suwung may also reference the eruption.[190][i] They are also—together with other texts—the source of the name "Samalas"[4] while the name "Suwung"—"quiet and without life"—may, in turn, be a reference to the aftermath of the eruption.[191]

Mount Rinjani avalanched and Mount Samalas collapsed, followed by large flows of debris accompanied by the noise coming from boulders. These flows destroyed Pamatan. All houses were destroyed and swept away, floating on the sea, and many people died. During seven days, big earthquakes shook the Earth, stranded in Leneng, dragged by the boulder flows, People escaped and some of them climbed the hills.

— Babad Lombok[192]

The city of Pamatan, capital of a kingdom on Lombok, was destroyed, and both disappeared from the historical record. The royal family survived the disaster according to the Javanese text,

Kertanegara of Singhasari on Java to conquer Bali in 1284 with little resistance.[142][197] It might have taken about a century for Lombok to recover from the eruption.[199] The western coast of Sumbawa was depopulated and remains so to this day; presumably the local populace viewed the area devastated by the eruption as "forbidden" and this memory persisted until recent times.[200]

Oceania and New Zealand

Historical events in Oceania are usually poorly dated, making it difficult to assess the timing and role of specific events, but there is evidence that between 1250 and 1300 there were crises in Oceania, for example at Easter Island, which may be linked with the beginning of the Little Ice Age and the Samalas eruption.[48] Around 1300, settlements in many places of the Pacific relocated, perhaps because of a sea level drop that occurred after 1250, and the 1991 eruption of Pinatubo has been linked to small drops in sea level.[169]

Climate change triggered by the Samalas eruption and the beginning of the Little Ice Age may have led to people in

AD and the arrival of people there and on other islands in the region may reflect such a climate-induced migration.[201]

Europe, Near East and Middle East

Contemporary chronicles in Europe mention unusual weather conditions in 1258.

The price for cereal increased in Britain,

England and Italy

A famine in London has been linked to this event;[52] this food crisis was not extraordinary[216] and there were issues with harvests already before the eruption.[217][218] The famine occurred at a time of political crisis between King Henry III of England and the English magnates.[219] 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.[88] Matthew Paris of St Albans described how until mid-August 1258, the weather alternated between cold and strong rain, causing high mortality.[215] The resulting famine was severe enough that grain was imported from Germany and Holland.[220]

In Italy, bad weather including intense rains in 1258 caused crop failures throughout the peninsula, as documented by numerous chronicles,

podestá (lord) of Parma Giberto da Gente [it] in 1259 was facilitated by the crisis, which induced his supporters to remain passive.[229] In Pavia, where a political crisis was already underway in 1257,[230] various economical and police measures were taken during the following two years to secure food supplies.[231] The city of Como in northern Italy repaired river banks that had been damaged by flooding,[232] and acquired grain for its consumption.[233] In Perugia, there were three years of food crisis between 1257 and 1260,[234] and the question of food supply played a major role in city politics and drove increased social control.[235] Perugia is also where the Flagellant movement arose;[236] it may have originated in the social distress caused by the effects of the eruption, though warfare and other causes probably played a more important role than natural events.[237]

Long-term consequences in Europe and the Near East

Over the long term, the cooling of the North Atlantic and sea ice expansion therein may have impacted the societies of Greenland and Iceland[238] by restraining navigation and agriculture, perhaps allowing further climate shocks around 1425 to end the existence of the Norse settlement in Greenland.[239] Another possible longer-term consequence of the eruption was the Byzantine Empire's loss of control over western Anatolia, because of a shift in political power from Byzantine farmers to mostly Turkoman pastoralists in the area. Colder winters caused by the eruption would have impacted agriculture more severely than pastoralism.[240]

Four Corners region, North America

The 1257 Samalas eruption took place during the

San Juan River was the site of the so-called cliff dwellings. Several sites were abandoned after the eruption.[241] The eruption took place during a time of decreased precipitation and lower temperatures and when population was declining.[242] The Samalas eruption[243] was one among several eruptions during this period which may have triggered climate stresses[244] such as a colder climate,[241] which in turn caused strife within the society of the Ancestral Puebloans; possibly they left the northern Colorado Plateau as a consequence.[244]

Altiplano, South America

In the

Salar de Coipasa despite the climatic change, implying that the local population effectively coped with the effects of the eruption.[245]

East Asia

Problems were also recorded in China, Japan, and Korea.

Ch'oe Ui.[248] Monsoon anomalies triggered by the Samalas eruption may have also impacted Angkor Wat in present-day Cambodia, which suffered a population decline at that time.[249] Other effects of the eruption may have[250] included a total darkening of the Moon in May 1258 during a lunar eclipse,[251] a phenomenon also recorded from Europe; volcanic aerosols reduced the amount of sunlight scattered into Earth's shadow and thus the brightness of the eclipsed Moon.[252]

Mongol Empire

Increased precipitation triggered by the eruption may have facilitated the Mongol invasions of the Levant[253] but later the return of the pre-Samalas climate would have reduced the livestock capacity of the region, thus reducing their military effectiveness[254] and paving the way to their military defeat in the Battle of Ain Jalut.[255] The effects of the eruption, such as famines, droughts and epidemics[256] may also have hastened the decline of the Mongol Empire, although the volcanic event is unlikely to have been the sole cause.[169] It may have altered the outcome of the Toluid Civil War[256] and shifted its centre of power towards the Chinese part dominated by Kublai Khan which was more adapted to cold winter conditions.[257]

Central Asia and the Black Death

The eruption of Samalas and other volcanoes caused climate disturbances in Central Asia, including a cooling[258] which was followed by a warming. This warming may have provided the environmental conditions for the spread and diversification of Yersinia pestis, the causative agent of the plague,[259] which about 1268 began diversifying and eventually yielded the strain that caused the Black Death.[260] Human populations may have been weakened by volcanic cooling-induced food crises and political/military unrest, facilitating the establishment of the outbreak.[261]

See also

Notes

  1. ^ The Volcanic Explosivity Index is a scale that measures the intensity of an explosive eruption;[2] a magnitude of 7 indicates an eruption that produces at least 100 cubic kilometres (24 cu mi) of volcanic deposits. Such eruptions occur once or twice per millennium, although their frequency might be underestimated due to incomplete geological and historical records.[3]
  2. ^ The dense rock equivalent is a measure of how voluminous the magma that the pyroclastic material originated from was.[16]
  3. ^ Tephrochronology is a technique that uses dated layers of tephra to correlate and synchronize events.[44]
  4. ^ Sulfate spikes around 44 BC and 426 BC, discovered later, rival its size.[75]
  5. ^ Although the Thailand droughts appear to continue past the point where the effects of the Samalas aerosols should have ceased.[113]
  6. ^ Winter warming is frequently observed after tropical volcanic eruptions,[141] due to dynamic effects triggered by the sulfate aerosols.[142][143]
  7. ^ The Little Ice Age was a period of several centuries during the last millennium during which global temperatures were depressed;[147] the cooling was associated with volcanic eruptions.[152]
  8. ^ δ18O is the ratio of the oxygen-18 isotope to the more common oxygen-16 isotope in water, which is influenced by climate.[176]
  9. ^ The term Babad refers to Javanese and Balinese chronicles. These babads are not original works but recompilations of older works that were presumably written around the 14th century.[190]
  10. Sasak people.[196]

References

  1. ^ "Rinjani". Global Volcanism Program. Smithsonian Institution. Retrieved 22 January 2020.
  2. ^ Newhall, Self & Robock 2018, p. 572.
  3. ^ Newhall, Self & Robock 2018, p. 573.
  4. ^ a b c d "Rinjani Dari Evolusi Kaldera hingga Geopark". Geomagz (in Indonesian). 4 April 2016. Archived from the original on 22 February 2018. Retrieved 3 March 2018.
  5. ^ a b c Métrich et al. 2018, p. 2258.
  6. ^ a b c Rachmat et al. 2016, p. 109.
  7. ^ Fontijn et al. 2015, p. 2.
  8. ^ Mutaqin et al. 2019, pp. 338–339.
  9. ^ a b c d Rachmat et al. 2016, p. 107.
  10. ^ a b c d e Rachmat et al. 2016, p. 108.
  11. ^ a b Mutaqin et al. 2019, p. 339.
  12. ^ a b c d e f Lavigne et al. 2013, p. 16743.
  13. , retrieved 10 December 2021
  14. ^ a b c d e Vidal et al. 2015, p. 3.
  15. ^ a b c d Vidal et al. 2015, p. 2.
  16. . Retrieved 19 October 2018.
  17. ^ Métrich et al. 2018, p. 2260.
  18. ^ Métrich et al. 2018, p. 2264.
  19. ^ Métrich et al. 2018, p. 2263.
  20. ^ a b Rachmat et al. 2016, p. 110.
  21. ^ a b c Malawani et al. 2022, p. 6.
  22. ^ a b Stevenson et al. 2019, p. 1547.
  23. .
  24. ^ Büntgen et al. 2022, p. 532.
  25. ^ Vidal et al. 2015, pp. 21–22.
  26. ^ a b Vidal et al. 2015, p. 18.
  27. ^ Vidal et al. 2015, pp. 17–18.
  28. .
  29. ^ a b Vidal et al. 2015, p. 5.
  30. ^ Mutaqin & Lavigne 2019, p. 5.
  31. ^ a b c d Vidal et al. 2015, p. 7.
  32. ^ Malawani et al. 2023, p. 2102.
  33. ^ Mutaqin et al. 2019, p. 344.
  34. ^ Vidal et al. 2015, p. 17.
  35. ^ Lavigne et al. 2013, p. 16744.
  36. ^ Malawani et al. 2023, p. 2110.
  37. ^ Mutaqin et al. 2019, p. 348.
  38. ^ Alloway et al. 2017, p. 87.
  39. ^ Alloway et al. 2017, p. 90.
  40. ^ Vidal et al. 2015, p. 8.
  41. ^ Vidal et al. 2015, p. 12.
  42. ^ Vidal et al. 2015, p. 16.
  43. ^ a b Vidal et al. 2015, p. 19.
  44. ISSN 1871-1014
    .
  45. ^ Fontijn et al. 2015, p. 8.
  46. .
  47. .
  48. ^ a b Margalef et al. 2018, p. 5.
  49. ^ Vidal et al. 2015, p. 14.
  50. ^ Vidal et al. 2016, p. 2.
  51. ^ Métrich et al. 2018, p. 2278.
  52. ^
    PMID 26097277
    .
  53. ^ a b c d Lavigne et al. 2013, p. 16745.
  54. .
  55. ^ Lavigne et al. 2013, Table S1.
  56. ^ a b Alloway et al. 2017, p. 86.
  57. ^ .
  58. ^ Vidal et al. 2015, pp. 14–15.
  59. S2CID 226971090
    .
  60. .
  61. ^ a b Bufanio 2022, p. 19.
  62. ^ .
  63. ^ a b c d Lavigne et al. 2013, p. 16742.
  64. ^ a b c d e f Hamilton 2013, p. 39.
  65. ^ Oppenheimer 2003, p. 417.
  66. hdl:11681/5296. Archived from the original
    (PDF) on 18 November 2016. Retrieved 29 January 2019.
  67. ^ Oppenheimer 2003, p. 418.
  68. ^ a b Vidal et al. 2016, p. 1.
  69. ^ Hammer, Clausen & Langway 1988, p. 103.
  70. ^ Hammer, Clausen & Langway 1988, p. 104.
  71. ^ Hammer, Clausen & Langway 1988, p. 106.
  72. from the original on 7 April 2019. Retrieved 7 April 2019.
  73. ^ a b Narcisi et al. 2019, p. 165.
  74. ^
    S2CID 118745561
    .
  75. .
  76. ^ Oppenheimer 2003, p. 419.
  77. ^ Oppenheimer 2003, p. 420.
  78. ^ a b Campbell 2017, p. 113.
  79. S2CID 140540145
    .
  80. ^ Stothers 2000, p. 361.
  81. ^ a b Brovkin et al. 2010, p. 675.
  82. ^ Oppenheimer 2003, p. 424.
  83. ^ Hammer, Clausen & Langway 1988, p. 107.
  84. ^ a b c d Oppenheimer 2003, p. 422.
  85. .
  86. .
  87. ^ Hamilton 2013, pp. 39–40.
  88. ^ a b c d e Hamilton 2013, p. 40.
  89. ^ "Centuries-old volcano mystery solved?". Science News. UPI. 18 June 2012. Archived from the original on 1 April 2019. Retrieved 11 March 2019.
  90. ^ Narcisi et al. 2019, p. 168.
  91. ^ Bufanio 2022, p. 20.
  92. ^ a b Kokfelt et al. 2016, p. 2.
  93. ^ Swingedouw et al. 2017, p. 28.
  94. ISSN 0377-0273
    .
  95. .
  96. .
  97. ^ Fu et al. 2016, p. 2862.
  98. .
  99. ^ a b Kokfelt et al. 2016, p. 6.
  100. ^ Baroni et al. 2019, p. 6.
  101. ^ Vidal et al. 2016, p. 7.
  102. S2CID 264451181
    .
  103. ^ Vidal et al. 2015, p. 21.
  104. ^ a b Stothers 2000, p. 362.
  105. ^ Baroni et al. 2019, p. 21.
  106. .
  107. from the original on 20 October 2018. Retrieved 19 October 2018.
  108. ^ .
  109. .
  110. .
  111. .
  112. .
  113. ^ .
  114. .
  115. ^ Stoffel et al. 2015, p. 787.
  116. ^ Brovkin et al. 2010, p. 674.
  117. ^ Brovkin et al. 2010, pp. 674–675.
  118. ^ Guillet et al. 2017, p. 123.
  119. .
  120. .
  121. ^ Wade et al. 2020, p. 26651.
  122. ^ .
  123. ^ Swingedouw et al. 2017, p. 30.
  124. ^ Stoffel et al. 2015, p. 785.
  125. ^ Wade et al. 2020, p. 26653.
  126. ^ Timmreck et al. 2009, p. 3.
  127. .
  128. .
  129. .
  130. ^ .
  131. .
  132. .
  133. ^ a b Guillet et al. 2017, p. 126.
  134. S2CID 128149914
    .
  135. (PDF) from the original on 22 July 2018. Retrieved 16 December 2018.
  136. .
  137. ^ Fu et al. 2016, p. 2859.
  138. ^ Wade et al. 2020, p. 26657.
  139. ^ Wade et al. 2020, p. 26656.
  140. .
  141. ^ a b Newhall, Self & Robock 2018, p. 575.
  142. ^ a b c Lavigne et al. 2013, p. 16746.
  143. ^
    ISSN 0065-9401
    .
  144. .
  145. .
  146. ^ Andres & Peltier 2016, p. 5783.
  147. ^ a b Andres & Peltier 2016, p. 5779.
  148. S2CID 10041389
    .
  149. .
  150. ^ Jomelli et al. 2016, p. 3.
  151. ^ Jomelli et al. 2016, p. 5.
  152. S2CID 127578885
    .
  153. ^ a b Margalef et al. 2018, p. 4.
  154. S2CID 15313398
    .
  155. .
  156. .
  157. ^ a b Dätwyler et al. 2017, p. 2336.
  158. ^ Dätwyler et al. 2017, pp. 2321–2322.
  159. S2CID 249813841
    .
  160. ^ Emile-Geay et al. 2008, p. 3141.
  161. ISSN 1944-8007
    .
  162. ^ Emile-Geay et al. 2008, p. 3144.
  163. S2CID 214671146
    .
  164. .
  165. .
  166. .
  167. ^ Swingedouw et al. 2017, p. 41.
  168. .
  169. ^ a b c Newhall, Self & Robock 2018, p. 576.
  170. S2CID 249090169
    .
  171. .
  172. .
  173. .
  174. .
  175. .
  176. ^ Stevenson et al. 2019, p. 1535.
  177. ^ Stevenson et al. 2019, p. 1548.
  178. PMID 31598188
    .
  179. .
  180. ^ a b Misios et al. 2022, p. 819.
  181. ^ Misios et al. 2022, p. 816.
  182. S2CID 252249527
    .
  183. .
  184. .
  185. .
  186. .
  187. .
  188. ^ Büntgen et al. 2022, p. 543.
  189. ^ a b Alloway et al. 2017, p. 98.
  190. ^ a b Mutaqin & Lavigne 2019, p. 2.
  191. ^ Mutaqin & Lavigne 2019, p. 4.
  192. ^ Lavigne et al. 2013, Supporting Information.
  193. ^ Hamilton 2013, p. 41.
  194. ^ Malawani et al. 2022, p. 8.
  195. ^ Mutaqin & Lavigne 2019, p. 9.
  196. S2CID 247378729
    .
  197. ^ from the original on 19 October 2018. Retrieved 18 October 2018.
  198. .
  199. ^ Malawani et al. 2022, p. 11.
  200. ^ Mutaqin & Lavigne 2019, p. 7–8.
  201. .
  202. .
  203. ^ Stothers 2000, p. 363.
  204. .
  205. ^ a b Dodds & Liddy 2011, p. 54.
  206. from the original on 20 October 2018. Retrieved 20 October 2018.
  207. ^ Grillo 2021, p. 150.
  208. ^ Guillet et al. 2017, p. 124.
  209. ^ a b Guillet et al. 2017, p. 127.
  210. ^ a b Stothers 2000, p. 366.
  211. ^ Bufanio 2022, p. 23.
  212. ^ a b Bufanio 2022, p. 25.
  213. ^ Stothers 2000, p. 364.
  214. ISSN 1988-4230
    .
  215. ^ .
  216. ^ Campbell 2017, p. 91.
  217. ^ Bufanio 2022, p. 27.
  218. ^ Campbell 2017, p. 108.
  219. ^ Campbell 2017, p. 119.
  220. .
  221. ^ Bufanio 2022, pp. 23, 25.
  222. ^ Bufanio 2022, p. 26.
  223. ^ Moglia 2022, p. 53.
  224. S2CID 232354348
    .
  225. .
  226. .
  227. ^ Moglia 2022, p. 52.
  228. ^ Moglia 2022, p. 55.
  229. ^ Moglia 2022, p. 58.
  230. ^ Bertoni 2022, p. 37.
  231. ^ Bertoni 2022, p. 39.
  232. ^ Grillo 2021, p. 153.
  233. ^ Grillo 2021, p. 154.
  234. ^ Luongo 2022, p. 76.
  235. ^ Luongo 2022, p. 77.
  236. ^ Luongo 2022, p. 63.
  237. ^ Stothers 2000, pp. 367–368.
  238. ^ Harrison & Maher 2014, pp. 156–157.
  239. ^ Harrison & Maher 2014, p. 180.
  240. .
  241. ^ .
  242. , retrieved 10 December 2021
  243. ^ Salzer 2000, p. 308.
  244. ^ a b Salzer 2000, pp. 312–314.
  245. PMID 29279865
    .
  246. ^ Guillet et al. 2017, p. 125.
  247. ^ Jenkins 2021, p. 63.
  248. S2CID 259928765
    .
  249. ^ Jenkins 2021, p. 82.
  250. ^ Bufanio 2022, p. 22.
  251. ^ Timmreck et al. 2009, p. 1.
  252. ^ Alloway et al. 2017, p. 96.
  253. ^ Di Cosmo, Wagner & Büntgen 2021, p. 92.
  254. ^ Di Cosmo, Wagner & Büntgen 2021, p. 97.
  255. ^ Di Cosmo, Wagner & Büntgen 2021, p. 100.
  256. ^ .
  257. .
  258. ^ Fell et al. 2020, p. 41.
  259. ^ Fell et al. 2020, p. 42.
  260. ^ Fell et al. 2020, p. 40.
  261. ^ Fell et al. 2020, p. 43.

Sources

External links