To install click the Add extension button. That's it.

The source code for the WIKI 2 extension is being checked by specialists of the Mozilla Foundation, Google, and Apple. You could also do it yourself at any point in time.

4,5
Kelly Slayton
Congratulations on this excellent venture… what a great idea!
Alexander Grigorievskiy
I use WIKI 2 every day and almost forgot how the original Wikipedia looks like.
Live Statistics
English Articles
Improved in 24 Hours
Added in 24 Hours
What we do. Every page goes through several hundred of perfecting techniques; in live mode. Quite the same Wikipedia. Just better.
.
Leo
Newton
Brights
Milds

1257 Samalas eruption

From 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 volcano on the Indonesian island of Lombok. The event had a probable 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.

YouTube Encyclopedic

  • 1/5
    Views:
    9 043
    9 377 816
    92 060
    144 448
    81 168
  • 1257 SAMALAS eruption - WikiVidi Documentary
  • Volcano Eruptions Size Comparison (2021)
  • 1257: le monde bascula... | L'Histoire nous le dira # 208
  • The Mystery Volcanic Eruption in 536; The Year of Hell
  • The Colossal Mystery Volcanic Eruption in 1808; Found at Last?

Transcription

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 Lesser Sunda Islands[8] in the Sunda Arc[9] of Indonesia,[10] a subduction zone where the Australian plate subducts beneath 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 dacitic, with a SiO
2
content of 62–63 percent by weight.[10] Volcanic rocks in the Banda arc 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

A small cone rising above a greenish lake within a large crater on a mountain
The 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 (Plinian) and P4 (pyroclastic flows).[25] The duration of the P1 and P3 phases is not known individually, but the two phases combined (not including P2) lasted between 12 and 15 hours.[26] The eruption column reached a height of 39–40 kilometres (24–25 mi) during the first stage (P1),[27] and of 38–43 kilometres (24–27 mi) during the third stage (P3);[26] it was high enough that SO2 in it and its sulfur isotope ratio was influenced by photolysis at high altitudes.[28]

Event

The eruption began with a phreatic (steam explosion powered) stage that deposited 3 centimetres (1.2 in) of ash over 400 square kilometres (150 sq mi) of northwest Lombok. A magmatic stage followed, and lithic-rich pumice rained down, the fallout reaching a thickness of 8 centimetres (3.1 in) both upwind on East Lombok and on Bali.[14] This was followed by lapilli rock as well as ash fallout, and pyroclastic flows that were partially confined within the valleys on Samalas's western flank. Some ash deposits were eroded by the pyroclastic flows, which created furrow structures in the ash. Pyroclastic flows crossed 10 kilometres (6.2 mi) of the Bali Sea, reaching the Gili Islands to the northwest of Samalas[29] and Taliwang east of Lombok,[21] while pumice blocks presumably covered the Alas Strait between Lombok and Sumbawa.[30] The deposits show evidence of interaction of the lava with water, so this eruption phase was probably 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]

Pyroclastic flows descended the northern slopes of Samalas; on the southern slopes they split into two branches that proceeded to the Alas Strait to the east and the Bali Strait to the west.[32] Coral reefs were buried by the pyroclastic flows; some flows crossed the Alas Strait between Sumbawa and Lombok and formed deposits on Sumbawa.[33] These pyroclastic flows reached volumes of 29 cubic kilometres (7.0 cu mi) on Lombok,[34] and thicknesses of 35 metres (115 ft) as far as 25 kilometres (16 mi) from Samalas.[35] The pyroclastic flows altered the geography of Lombok; they and sediments eroded from the Samalas deposits extended the shorelines of the island[36] and buried 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.[11] Tephra in the form of layers of fine ash from the eruption fell as far away as Java, forming part of the Muntilan Tephra, which was found on the slopes of other volcanoes of Java, but could not be linked to eruptions in these volcanic systems. This tephra is now considered to be a product of the 1257 eruption and is thus also known as the Samalas Tephra.[31][38] It reaches 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[39] and 12–17 centimetres (4.7–6.7 in) on Bali's Agung volcano.[40] In Lake Logung 340 kilometres (210 mi) away from Samalas[31] on Java it is 3 centimetres (1.2 in) thick. Most of the tephra was deposited west-southwest of Samalas.[41] Considering the thickness of Samalas Tephra found at Mount Merapi, the total volume may have reached 32–39 cubic kilometres (7.7–9.4 cu mi).[42] 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 these were a Plinian eruption and an Ultraplinian eruption respectively.[43]

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).[49] The total dense rock equivalent volume of the whole eruption was at least 40 cubic kilometres (9.6 cu mi).[43] The magma erupted was trachydacitic and contained amphibole, apatite, clinopyroxene, iron sulfide, orthopyroxene, plagioclase, and titanomagnetite. It formed out of 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 Volcanic Explosivity Index of 7,[52] making it one of the largest eruptions of the current epoch, the Holocene.[53] Eruptions of comparable intensity include the Kurile lake eruption (in Kamchatka, Russia) in the 7th millennium BC, the Mount Mazama (United States, Oregon) eruption in the 6th millennium BC,[53] the Cerro Blanco (Argentina) eruption about 4,200 years ago,[54] the Minoan eruption (in Santorini, Greece)[53] between 1627 and 1600 BC,[55] the Tierra Blanca Joven eruption of Lake Ilopango (El Salvador) in the 6th century, and 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.[14] Early research considered that the caldera-forming eruption occurred between 1210 and 1300. In 2013, Lavigne suggested that the eruption occurred between May and October 1257, resulting in the 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;[61][62][63] specifically increased sulfate concentrations were found[64] in 1980 within the Crête ice core[65] (Greenland, drilled in 1974[66]) associated with a deposit of rhyolitic ash.[67] The 1257–1258 layer is the third largest sulfate signal at Crête;[68] at first a source in a volcano near Greenland had been considered[64] but Icelandic records made no mention of eruptions around 1250 and it was found in 1988 that ice cores in Antarctica—at Byrd Station and the South Pole—also contained sulfate signals.[69] Sulfate spikes were also found in ice cores from Ellesmere Island, Canada,[70] and the Samalas sulfate spikes were used as stratigraphic markers for ice cores even before the volcano that caused them was known.[71]

The ice cores indicated a large sulfate spike, accompanied by tephra deposition,[72] around 1257–1259,[73][72] the largest[d] in 7,000 years and twice the size of the spike due to the 1815 eruption of Tambora.[73] In 2003, a dense rock equivalent volume of 200–800 cubic kilometres (48–192 cu mi) was estimated for this eruption,[75] but it was also proposed that the eruption might have been somewhat smaller and richer in sulfur.[76][61] The volcano responsible was thought to be located in the Ring of Fire[77] 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.[78] A volcanic eruption in 1256 at Harrat al-Rahat near Medina was also too small to trigger these events.[79] Other proposals included several simultaneous eruptions.[80] The diameter of the caldera left by the eruption was estimated to be 10–30 kilometres (6.2–18.6 mi),[81] and the location was estimated to be close to the equator and probably north of it.[82]

While at first no clear-cut climate anomaly could be correlated to the 1257 sulfate layers,[83][84] in 2000[83] 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.[83] 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.[85] 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[86]

The conclusive link between these events and an eruption of Samalas was made in 2013 on the basis of[62] radiocarbon dating of trees on Lombok[87] and the Babad Lombok, a series of writings in Old Javanese on palm leaves[62] that described a catastrophic volcanic event on Lombok which occurred before 1300.[12] These findings induced Franck Lavigne,[64] a geoscientist of the Pantheon-Sorbonne University[88] who had already suspected that a volcano on that island may be responsible, to conclude that the Samalas volcano was this volcano.[64] 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.[57] Later, geochemical similarities between tephra found in polar ice cores and eruption products of Samalas reinforced this localization.[89][90]

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;[91] one reconstruction even considers it the strongest of the last 2500 years.[92] It is about eight times stronger than that of Krakatau.[64] In the northern hemisphere it is only exceeded by the signal of the destructive 1783/1784 Laki eruption.[91] The ice core sulfate spikes have been used as a time marker in chronostratigraphic studies.[93] Ice cores from Illimani in Bolivia contain thallium[94] and sulfate spikes from the eruption.[95] For comparison, the 1991 eruption of Pinatubo ejected only about a tenth of the amount of sulfur erupted by Samalas.[96] Sulfate deposition from the Samalas eruption has been noted at Svalbard,[97] and the fallout of sulfuric acid from the volcano may have directly affected peatlands in northern Sweden.[98]

In addition, the sulfate aerosols may have extracted large amounts of the beryllium isotope 10
Be
from the stratosphere; such an extraction event and the subsequent deposition in ice cores may mimic changes in solar activity.[99] The amount of sulfur dioxide released by the eruption has been estimated to be 158 ± 12 million tonnes.[50] Whether the mass release was higher or lower than for Tambora is contentious; Tambora might have produced more sulfur[100] but Samalas may have been more effective at injecting tephra into the stratosphere.[101] After the eruption, it probably took weeks to months for the fallout to reach large distances from the volcano.[77] When large scale volcanic eruptions inject aerosols into the atmosphere, they can form stratospheric veils. These reduce the amount of light reaching the surface and cause lower temperatures, which can lead to poor crop yields.[102] 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.[103]

Other records of the eruption's impact include decreased tree growth in Mongolia between 1258 and 1262 based on tree ring data,[104] frost rings (tree rings damaged by frost during the growth season[105]), light tree rings in Canada and northwestern Siberia from 1258 and 1259 respectively,[106] thin tree rings in the Sierra Nevada, California, U.S.[107] cooling in sea surface temperature records off the Korean Peninsula[108] and in lake sediments of northeastern China,[109] a very wet monsoon in Vietnam,[87] droughts in many places in the Northern Hemisphere[110] as well as in southern Thailand cave records,[e][111] and a decade-long thinning of tree rings in Norway and Sweden.[112] Cooling may have lasted for 4–5 years based on simulations and tree ring data.[113]

Another effect of the eruption-induced climate change may have been a brief decrease in atmospheric carbon dioxide concentrations.[80] 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,[114] and increased productivity of the biosphere due to increased scattered sunlight and the fertilization of oceans by volcanic ash.[115]

The Samalas signal is only inconsistently reported from tree ring climate information,[116][117] and the temperature effects were likewise limited, probably because the large sulfate output altered the average size of particles and thus their radiative forcing.[118] 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.[119][120] Better modelling with a general circulation model that includes a detailed description of the aerosol indicated that the principal temperature anomaly occurred in 1258 and continued until 1261.[120] Climate models tend to overestimate the climate impact of a volcanic eruption;[121] one explanation is that climate models tend to assume that aerosol optical depth increases linearly with the quantity of erupted sulfur[122] when in reality self-limiting processes limit its growth.[123] The possible occurrence of an El Niño before the eruption may have further reduced the cooling.[124]

The Samalas eruption, together with 14th century cooling, is thought to have set off a growth of ice caps and sea ice,[125] and glaciers in the Alps, Bhutan Himalaya, the Pacific Northwest and the Patagonian Andes grew in size.[126][127] The advances of ice after the Samalas eruption may have strengthened and prolonged the climate effects.[98] 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.[128] The glacier advances triggered by the Samalas eruption are documented on Baffin Island, where the advancing ice killed and then incorporated vegetation, conserving it.[129] Likewise, a change in Arctic Canada from a warm climate phase to a colder one coincides with the Samalas eruption.[130]

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.[83] 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.[131] For comparison, the radiative forcing of Pinatubo's 1991 eruption was about a seventh of that of the Samalas eruption.[132] Sea surface temperatures too decreased by 0.3–2.2 °C (0.54–3.96 °F),[133] triggering changes in the ocean circulations. Ocean temperature and salinity changes may have lasted for a decade.[134] Precipitation and evaporation both decreased, evaporation reduced more than precipitation.[135]

Volcanic eruptions can also deliver bromine and chlorine into the stratosphere, where they contribute to the breakdown of ozone through their oxides chlorine monoxide and bromine monoxide. While most bromine and chlorine erupted would have been scavenged by the eruption column and thus would not have entered the stratosphere, the quantities that have been modelled for the Samalas halogen release (227 ± 18 million tonnes of chlorine and up to 1.3 ± 0.3 million tonnes of bromine) would have reduced stratospheric ozone[50] although only a small portion of the halogens would have reached the stratosphere.[136] One hypothesis is that the resulting increase in 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.[137]

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.[138] After an early warm winter 1257–1258[f][139] resulting in the early flowering of violets according to reports from the Kingdom of France,[140] European summers were colder after the eruption,[142] and winters were long and cold.[143]

The Samalas eruption came after the Medieval Climate Anomaly,[144] a period early in the last millennium with unusually warm temperatures,[145] 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.[146] The time period 1250–1300 was heavily disturbed by volcanic activity[128] from four eruptions in 1230, 1257, 1276 and 1286,[147] and is recorded by a moraine from a glacial advance on Disko Island,[148] although the moraine may indicate a pre-Samalas cold spell.[149] 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,[151][152] though this theory is not without disagreement.[153] The Samalas eruption in Europe is sometimes used as a chronological marker for the beginning of the Little Ice Age.[154]

Other inferred effects of the eruption are:

Other regions such as Alaska were mostly unaffected.[181] There is little evidence that tree growth was influenced by cold in what is now the Western United States,[182] where the eruption may have interrupted a prolonged drought period.[183] The climate effect in Alaska may have been moderated by the nearby ocean.[184] In 1259, Western Europe and the west coastal North America had mild weather[131] and there is no evidence for summer precipitation changes in Central Europe.[185] Tree rings do not show much evidence of precipitation changes.[186]

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.[102]

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.[187] 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.[188][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.[189]

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[190]

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,[191] which also mentions reconstruction and recovery efforts after the eruption,[192] and there is no clear-cut evidence that the kingdom itself was destroyed by the eruption, as the history there is poorly known in general.[187] Thousands of people died during the eruption[12] although it is possible that the population of Lombok fled before the eruption.[193] In Bali the number of inscriptions[j] dropped off after the eruption,[195] and Bali and Lombok may have been depopulated by it,[196] possibly for generations, allowing King Kertanegara of Singhasari on Java to conquer Bali in 1284 with little resistance.[140][195] It might have taken about a century for Lombok to recover from the eruption.[197] 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.[198]

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.[167]

Climate change triggered by the Samalas eruption and the beginning of the Little Ice Age may have led to people in Polynesia migrating southwestward in the 13th century. The first settlement of New Zealand most likely occurred 1230–1280 AD and the arrival of people there and on other islands in the region may reflect such a climate-induced migration.[199]

Europe, Near East and Middle East

Contemporary chronicles in Europe mention unusual weather conditions in 1258.[200] Reports from 1258 in France and England indicate a dry fog, giving the impression of a persistent cloud cover to contemporary observers.[201] Medieval chronicles say that in 1258, the summer was cold and rainy, causing floods and bad harvests,[63] with cold from February to June.[202] Frost occurred in the summer 1259 according to Russian chronicles.[106] In Europe and the Middle East, changes in atmospheric colours, storms, cold, and severe weather were reported in 1258–1259,[203] with agricultural problems extending to North Africa.[204] In Europe, excess rain, cold and high cloudiness damaged crops and caused famines followed by epidemics,[205][206][87] although 1258–1259 did not lead to famines as bad as some other famines such as the Great Famine of 1315–17.[207]

The price for cereal increased in Britain,[203] France,[208] and Italy, augmented by price speculation.[209] Outbreaks of disease occurred during this time in the Middle East, England[208] and Italy, including typhus.[210] During and after the winter of 1258–59, exceptional weather was reported less commonly, but the winter of 1260–61 was very severe in Iceland, Italy, and elsewhere.[211] The disruption caused by the eruption may have influenced the onset of the Mudéjar revolt of 1264–1266 in Iberia.[212]

England and Italy

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[213]

A famine in London has been linked to this event;[52] this food crisis was not extraordinary[214] and there were issues with harvests already before the eruption.[215][216] The famine occurred at a time of political crisis between King Henry III of England and the English magnates.[217] 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.[87] Matthew Paris of St Albans described how until mid-August 1258, the weather alternated between cold and strong rain, causing high mortality.[213] The resulting famine was severe enough that grain was imported from Germany and Holland.[218]

In Italy, bad weather including intense rains in 1258 caused crop failures throughout the peninsula, as documented by numerous chronicles,[219] although impacts varied between regions.[210] Relative to most of Europe, the impact in Italy hit a year later.[220] In 1259, a cold wave led to high mortality throughout Italy.[221] The cities of Bologna and Siena in Italy attempted to manage the food crisis by buying and subsidizing grain, banning its export and limiting its price.[222] Siena also initiated diplomatic relations with Manfred, King of Sicily, ostensibly to help manage the food crisis,[223] while a political crisis set in in Bologna, which was also weakened geopolitically.[224] Parma ordered the sale of grain and tasked officials with monitoring markets, including closing them on Saturdays,[225] and banned food exports.[226] It is likely that the overthrow of the podestá (lord) of Parma Giberto da Gente [it] in 1259 was facilitated by the crisis, which induced his supporters to remain passive.[227] In Pavia, where a political crisis was already underway in 1257,[228] various economical and police measures were taken during the following two years to secure food supplies.[229] The city of Como in northern Italy repaired river banks that had been damaged by flooding,[230] and acquired grain for its consumption.[231] In Perugia, there were three years of food crisis between 1257 and 1260,[232] and the question of food supply played a major role in city politics and drove increased social control.[233] Perugia is also where the Flagellant movement arose;[234] 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.[235]

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[236] by restraining navigation and agriculture, perhaps allowing further climate shocks around 1425 to end the existence of the Norse settlement in Greenland.[237] 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.[238]

Four Corners region, North America

The 1257 Samalas eruption took place during the Pueblo III Period in southwestern North America, during which the Mesa Verde region on the San Juan River was the site of the so-called cliff dwellings. Several sites were abandoned after the eruption.[239] The eruption took place during a time of decreased precipitation and lower temperatures and when population was declining.[240] The Samalas eruption[241] was one among several eruptions during this period which may have triggered climate stresses[242] such as a colder climate,[239] which in turn caused strife within the society of the Ancestral Puebloans; possibly they left the northern Colorado Plateau as a consequence.[242]

Altiplano, South America

In the Altiplano of South America, a cold and dry interval between 1200 and 1450 has been associated with the Samalas eruption and the 1280 eruption of Quilotoa volcano in Ecuador. The use of rain-fed agriculture increased in the area between the Salar de Uyuni and the Salar de Coipasa despite the climatic change, implying that the local population effectively coped with the effects of the eruption.[243]

East Asia

Problems were also recorded in China, Japan, and Korea.[87] In Japan, the Azuma Kagami chronicle mentions that rice paddies and gardens were destroyed by the cold and wet weather,[244] and the so-called Shôga famine—which among other things stimulated the Japanese religious reformer Nichiren[245] may have been aggravated by bad weather in 1258 and 1259.[207] Along with the Mongol invasions of Korea, hardship caused by the Samalas eruption may have precipitated the downfall of the Goryeo military regime and of its last Choe dictator, Choe Ui.[246] 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.[247] Other effects of the eruption may have[248] included a total darkening of the Moon in May 1258 during a lunar eclipse,[249] 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.[250]

Mongol Empire

Increased precipitation triggered by the eruption may have facilitated the Mongol invasions of the Levant[251] but later the return of the pre-Samalas climate would have reduced the livestock capacity of the region, thus reducing their military effectiveness[252] and paving the way to their military defeat in the Battle of Ain Jalut.[253] The effects of the eruption, such as famines, droughts and epidemics[254] may also have hastened the decline of the Mongol Empire, although the volcanic event is unlikely to have been the sole cause.[167] It may have altered the outcome of the Toluid Civil War[254] and shifted its centre of power towards the Chinese part dominated by Kublai Khan which was more adapted to cold winter conditions.[255]

Central Asia and the Black Death

The eruption of Samalas and other volcanoes caused climate disturbances in Central Asia, including a cooling[256] 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,[257] which about 1268 began diversifying and eventually yielded the strain that caused the Black Death.[258] Human populations may have been weakened by volcanic cooling-induced food crises and political/military unrest, facilitating the establishment of the outbreak.[259]

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.[74]
  5. ^ Although the Thailand droughts appear to continue past the point where the effects of the Samalas aerosols should have ceased.[111]
  6. ^ Winter warming is frequently observed after tropical volcanic eruptions,[139] due to dynamic effects triggered by the sulfate aerosols.[140][141]
  7. ^ The Little Ice Age was a period of several centuries during the last millennium during which global temperatures were depressed;[145] the cooling was associated with volcanic eruptions.[150]
  8. ^ δ18O is the ratio of the oxygen-18 isotope to the more common oxygen-16 isotope in water, which is influenced by climate.[174]
  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.[188]
  10. ^ And on Lombok, the historical record of the Sasak people.[194]

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. ^ Corlett, Richard T. (27 June 2019), "Physical geography", The Ecology of Tropical East Asia, Oxford University Press, pp. 26–61, doi:10.1093/oso/9780198817017.003.0002, ISBN 978-0-19-881701-7, 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. ^ Pyle, David M. (2015). "Sizes of Volcanic Eruptions". The Encyclopedia of Volcanoes. pp. 257–264. doi:10.1016/B978-0-12-385938-9.00013-4. ISBN 9780123859389. 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. ^ 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.
  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. ^ 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. ^ 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. ^ Lowe, David J. (April 2011). "Tephrochronology and its application: A review". Quaternary Geochronology. 6 (2): 107. Bibcode:2011QuGeo...6..107L. doi:10.1016/j.quageo.2010.08.003. hdl:10289/4616. ISSN 1871-1014.
  45. ^ Fontijn et al. 2015, p. 8.
  46. ^ 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.
  47. ^ 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. Bibcode:2017JESS..126....7Y. doi:10.1007/s12040-016-0790-y. ISSN 0253-4126.
  48. ^ a b Margalef et al. 2018, p. 5.
  49. ^ Vidal et al. 2015, p. 14.
  50. ^ a b c 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.
  51. ^ Métrich et al. 2018, p. 2278.
  52. ^ a b 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. PMC 4470363. PMID 26097277.
  53. ^ a b c d Lavigne et al. 2013, p. 16745.
  54. ^ Fernandez-Turiel, J. L.; Perez–Torrado, F. J.; Rodriguez-Gonzalez, A.; Saavedra, J.; Carracedo, J. C.; Rejas, M.; Lobo, A.; Osterrieth, M.; Carrizo, J. I.; Esteban, G.; Gallardo, J.; Ratto, N. (8 May 2019). "La gran erupción de hace 4.2 ka cal en Cerro Blanco, Zona Volcánica Central, Andes: nuevos datos sobre los depósitos eruptivos holocenos en la Puna sur y regiones adyacentes". Estudios Geológicos. 75 (1): 26. doi:10.3989/egeol.43438.515.
  55. ^ Lavigne et al. 2013, Table S1.
  56. ^ a b Alloway et al. 2017, p. 86.
  57. ^ a b c d Reid, Anthony (2016). "Revisiting Southeast Asian History with Geology: Some Demographic Consequences of a Dangerous Environment". In Bankoff, Greg; Christensen, Joseph (eds.). Natural Hazards and Peoples in the Indian Ocean World. Palgrave Series in Indian Ocean World Studies. Palgrave Macmillan US. p. 33. doi:10.1057/978-1-349-94857-4_2. ISBN 978-1-349-94857-4.
  58. ^ Vidal et al. 2015, pp. 14–15.
  59. ^ Roverato, Matteo; Dufresne, Anja; Procter, Jonathan, eds. (2021). "Volcanic Debris Avalanches". Advances in Volcanology: 40. doi:10.1007/978-3-030-57411-6. ISBN 978-3-030-57410-9. ISSN 2364-3277. S2CID 226971090.
  60. ^ 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. Advances in Volcanology. p. 4. doi:10.1007/11157_2016_34. ISBN 978-3-319-44095-8.
  61. ^ a b Bufanio 2022, p. 19.
  62. ^ a b c d e f "Culprit Behind Medieval Eruption". Science. 342 (6154): 21.2–21. 3 October 2013. doi:10.1126/science.342.6154.21-b.
  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. ^ Langway, Chester C. (2008). "The history of early polar ice cores" (PDF). Cold Regions Science and Technology. 52 (2): 28. Bibcode:2008CRST...52..101L. doi:10.1016/j.coldregions.2008.01.001. hdl:11681/5296. Archived from the original (PDF) on 18 November 2016. Retrieved 29 January 2019.
  67. ^ Oppenheimer 2003, p. 418.
  68. ^ Hammer, Clausen & Langway 1988, p. 103.
  69. ^ Hammer, Clausen & Langway 1988, p. 104.
  70. ^ Hammer, Clausen & Langway 1988, p. 106.
  71. ^ Osipova, O. P.; Shibaev, Y. A.; Ekaykin, A. A.; Lipenkov, V. Y.; Onischuk, N. A.; Golobokova, L. P.; Khodzher, T. V.; Osipov, E. Y. (7 May 2014). "High-resolution 900 year volcanic and climatic record from the Vostok area, East Antarctica". The Cryosphere. 8 (3): 7. Bibcode:2014TCry....8..843O. doi:10.5194/tc-8-843-2014. ISSN 1994-0416. Archived from the original on 7 April 2019. Retrieved 7 April 2019.
  72. ^ a b Narcisi et al. 2019, p. 165.
  73. ^ a b 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. S2CID 118745561.
  74. ^ 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): 543–9. Bibcode:2015Natur.523..543S. doi:10.1038/nature14565. PMID 26153860. S2CID 4462058.
  75. ^ Oppenheimer 2003, p. 419.
  76. ^ Oppenheimer 2003, p. 420.
  77. ^ a b Campbell 2017, p. 113.
  78. ^ 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. S2CID 140540145.
  79. ^ Stothers 2000, p. 361.
  80. ^ a b Brovkin et al. 2010, p. 675.
  81. ^ Oppenheimer 2003, p. 424.
  82. ^ Hammer, Clausen & Langway 1988, p. 107.
  83. ^ a b c d Oppenheimer 2003, p. 422.
  84. ^ Zielinski, Gregory A. (1995). "Stratospheric loading and optical depth estimates of explosive volcanism over the last 2100 years derived from the Greenland Ice Sheet Project 2 ice core". Journal of Geophysical Research. 100 (D10): 20949. Bibcode:1995JGR...10020937Z. doi:10.1029/95JD01751.
  85. ^ 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.
  86. ^ Hamilton 2013, pp. 39–40.
  87. ^ a b c d e Hamilton 2013, p. 40.
  88. ^ "Centuries-old volcano mystery solved?". Science News. UPI. 18 June 2012. Archived from the original on 1 April 2019. Retrieved 11 March 2019.
  89. ^ Narcisi et al. 2019, p. 168.
  90. ^ Bufanio 2022, p. 20.
  91. ^ a b Kokfelt et al. 2016, p. 2.
  92. ^ Swingedouw et al. 2017, p. 28.
  93. ^ Boudon, Georges; Balcone-Boissard, Hélène; Solaro, Clara; Martel, Caroline (September 2017). "Revised chronostratigraphy of recurrent ignimbritic eruptions in Dominica (Lesser Antilles arc): Implications on the behavior of the magma plumbing system" (PDF). Journal of Volcanology and Geothermal Research. 343: 135. Bibcode:2017JVGR..343..135B. doi:10.1016/j.jvolgeores.2017.06.022. ISSN 0377-0273.
  94. ^ Kellerhals, Thomas; Tobler, Leonhard; Brütsch, Sabina; Sigl, Michael; Wacker, Lukas; Gäggeler, Heinz W.; Schwikowski, Margit (1 February 2010). "Thallium as a Tracer for Preindustrial Volcanic Eruptions in an Ice Core Record from Illimani, Bolivia". Environmental Science & Technology. 44 (3): 888–93. Bibcode:2010EnST...44..888K. doi:10.1021/es902492n. ISSN 0013-936X. PMID 20050662.
  95. ^ 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.
  96. ^ Fu et al. 2016, p. 2862.
  97. ^ 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.
  98. ^ a b Kokfelt et al. 2016, p. 6.
  99. ^ Baroni et al. 2019, p. 6.
  100. ^ Pouget, Manon; Moussallam, Yves; Rose-Koga, Estelle F.; Sigurdsson, Haraldur (25 October 2023). "A reassessment of the sulfur, chlorine and fluorine atmospheric loading during the 1815 Tambora eruption". Bulletin of Volcanology. 85 (11): 12. Bibcode:2023BVol...85...66P. doi:10.1007/s00445-023-01683-8. S2CID 264451181.
  101. ^ Vidal et al. 2015, p. 21.
  102. ^ a b Stothers 2000, p. 362.
  103. ^ Baroni et al. 2019, p. 21.
  104. ^ 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.
  105. ^ 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): 105. Bibcode:2015CliPa..11..105B. doi:10.5194/cp-11-105-2015. ISSN 1814-9324. Archived from the original on 20 October 2018. Retrieved 19 October 2018.
  106. ^ a b Hantemirov, Rashit M; Gorlanova, Ludmila A; Shiyatov, Stepan G (July 2004). "Extreme temperature events in summer in northwest Siberia since AD 742 inferred from tree rings". Palaeogeography, Palaeoclimatology, Palaeoecology. 209 (1–4): 161. Bibcode:2004PPP...209..155H. doi:10.1016/j.palaeo.2003.12.023. ISSN 0031-0182.
  107. ^ Scuderi, Louis A. (1990). "Tree-Ring Evidence for Climatically Effective Volcanic Eruptions". Quaternary Research. 34 (1): 73. Bibcode:1990QuRes..34...67S. doi:10.1016/0033-5894(90)90073-T. ISSN 1096-0287. S2CID 129758817.
  108. ^ Lee, Kyung Eun; Park, Wonsun; Yeh, Sang-Wook; Bae, Si Woong; Ko, Tae Wook; Lohmann, Gerrit; Nam, Seung-Il (1 September 2021). "Enhanced climate variability during the last millennium recorded in alkenone sea surface temperatures of the northwest Pacific margin". Global and Planetary Change. 204: 7. Bibcode:2021GPC...20403558L. doi:10.1016/j.gloplacha.2021.103558. ISSN 0921-8181.
  109. ^ Chu, Guoqiang; Sun, Qing; Wang, Xiaohua; Liu, Meimei; Lin, Yuan; Xie, Manman; Shang, Wenyu; Liu, Jiaqi (1 July 2012). "Seasonal temperature variability during the past 1600 years recorded in historical documents and varved lake sediment profiles from northeastern China". The Holocene. 22 (7): 787. Bibcode:2012Holoc..22..785C. doi:10.1177/0959683611430413. ISSN 0959-6836. S2CID 128544002.
  110. ^ Fei, Jie; Zhou, Jie (February 2016). "The drought and locust plague of 942–944 AD in the Yellow River Basin, China". Quaternary International. 394: 120. Bibcode:2016QuInt.394..115F. doi:10.1016/j.quaint.2014.11.053. ISSN 1040-6182.
  111. ^ a b Tan, Liangcheng; Shen, Chuan-Chou; Löwemark, Ludvig; Chawchai, Sakonvan; Edwards, R. Lawrence; Cai, Yanjun; Breitenbach, Sebastian F. M.; Cheng, Hai; Chou, Yu-Chen; Duerrast, Helmut; Partin, Judson W.; Cai, Wenju; Chabangborn, Akkaneewut; Gao, Yongli; Kwiecien, Ola; Wu, Chung-Che; Shi, Zhengguo; Hsu, Huang-Hsiung; Wohlfarth, Barbara (27 August 2019). "Rainfall variations in central Indo-Pacific over the past 2,700 y". Proceedings of the National Academy of Sciences. 116 (35): 17202, 17204. Bibcode:2019PNAS..11617201T. doi:10.1073/pnas.1903167116. ISSN 0027-8424. PMC 6717306. PMID 31405969.
  112. ^ Thun, Terje; Svarva, Helene (February 2018). "Tree-ring growth shows that the significant population decline in Norway began decades before the Black Death". Dendrochronologia. 47: 28. Bibcode:2018Dendr..47...23T. doi:10.1016/j.dendro.2017.12.002. ISSN 1125-7865.
  113. ^ Stoffel et al. 2015, p. 787.
  114. ^ Brovkin et al. 2010, p. 674.
  115. ^ Brovkin et al. 2010, pp. 674–675.
  116. ^ Guillet et al. 2017, p. 123.
  117. ^ 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.
  118. ^ Boucher, Olivier (2015). "Stratospheric Aerosols". Atmospheric Aerosols. Springer Netherlands. p. 279. doi:10.1007/978-94-017-9649-1_12. ISBN 978-94-017-9649-1.
  119. ^ Wade et al. 2020, p. 26651.
  120. ^ a b 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. 17: 1268. Bibcode:2015EGUGA..17.1268G.
  121. ^ Swingedouw et al. 2017, p. 30.
  122. ^ Stoffel et al. 2015, p. 785.
  123. ^ Wade et al. 2020, p. 26653.
  124. ^ Timmreck et al. 2009, p. 3.
  125. ^ Brewington, Seth D. (May 2016). "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.
  126. ^ Yang, Weilin; Li, Yingkui; Liu, Gengnian; Chu, Wenchao (21 September 2022). "Timing and climatic-driven mechanisms of glacier advances in Bhutanese Himalaya during the Little Ice Age". The Cryosphere. 16 (9): 3747. Bibcode:2022TCry...16.3739Y. doi:10.5194/tc-16-3739-2022. ISSN 1994-0416. S2CID 252451837.
  127. ^ Huston, Alan; Siler, Nicholas; Roe, Gerard H.; Pettit, Erin; Steiger, Nathan J. (1 April 2021). "Understanding drivers of glacier-length variability over the last millennium". The Cryosphere. 15 (3): 1647. Bibcode:2021TCry...15.1645H. doi:10.5194/tc-15-1645-2021. ISSN 1994-0416. S2CID 233737859.
  128. ^ a b 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. S2CID 54881452.
  129. ^ 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. S2CID 128567847.
  130. ^ 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.
  131. ^ a b Guillet et al. 2017, p. 126.
  132. ^ 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. S2CID 128149914.
  133. ^ 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" (PDF). Geophysical Research Letters. 43 (3): 1185. Bibcode:2016GeoRL..43.1184C. doi:10.1002/2015GL067359. Archived (PDF) from the original on 22 July 2018. Retrieved 16 December 2018.
  134. ^ 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.
  135. ^ Fu et al. 2016, p. 2859.
  136. ^ Wade et al. 2020, p. 26657.
  137. ^ Wade et al. 2020, p. 26656.
  138. ^ 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.
  139. ^ a b Newhall, Self & Robock 2018, p. 575.
  140. ^ a b c Lavigne et al. 2013, p. 16746.
  141. ^ a b Baldwin, Mark P.; Birner, Thomas; Brasseur, Guy; Burrows, John; Butchart, Neal; Garcia, Rolando; Geller, Marvin; Gray, Lesley; Hamilton, Kevin; Harnik, Nili; Hegglin, Michaela I.; Langematz, Ulrike; Robock, Alan; Sato, Kaoru; Scaife, Adam A. (1 January 2018). "100 Years of Progress in Understanding the Stratosphere and Mesosphere". Meteorological Monographs. 59: 27.36. Bibcode:2018MetMo..59...27B. doi:10.1175/AMSMONOGRAPHS-D-19-0003.1. ISSN 0065-9401.
  142. ^ 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): EPSC2016-4968. Bibcode:2016EGUGA..18.4968L. doi:10.1088/1748-9326/11/2/024001.
  143. ^ 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" (PDF). Quaternary Science Reviews. 122: 74–88. Bibcode:2015QSRv..122...74H. doi:10.1016/j.quascirev.2015.05.029.
  144. ^ Andres & Peltier 2016, p. 5783.
  145. ^ a b Andres & Peltier 2016, p. 5779.
  146. ^ Bradley, R. S.; Wanner, H.; Diaz, H. F. (22 January 2016). "The Medieval Quiet Period". The Holocene. 26 (6): 992. Bibcode:2016Holoc..26..990B. doi:10.1177/0959683615622552. S2CID 10041389.
  147. ^ Nicolussi, Kurt; Le Roy, Melaine; Schlüchter, Christian; Stoffel, Markus; Wacker, Lukas (July 2022). "The glacier advance at the onset of the Little Ice Age in the Alps: New evidence from Mont Miné and Morteratsch glaciers". The Holocene. 32 (7): 635. Bibcode:2022Holoc..32..624N. doi:10.1177/09596836221088247. hdl:20.500.11850/549477. ISSN 0959-6836. S2CID 248732759.
  148. ^ Jomelli et al. 2016, p. 3.
  149. ^ Jomelli et al. 2016, p. 5.
  150. ^ Wang, Zhiyuan; Wang, Jianglin; Zhang, Shijia (25 January 2019). "Variations of the global annual mean surface temperature during the past 2000 years: results from the CESM1". Theoretical and Applied Climatology. 137 (3–4): 8. Bibcode:2019ThApC.137.2877W. doi:10.1007/s00704-019-02775-2. S2CID 127578885.
  151. ^ a b Margalef et al. 2018, p. 4.
  152. ^ 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" (PDF). Geophysical Research Letters. 39 (2): L02708. Bibcode:2012GeoRL..39.2708M. doi:10.1029/2011GL050168. S2CID 15313398.
  153. ^ 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.
  154. ^ Jomelli, Vincent; Palacios, David; Hughes, Philip D.; Cartapanis, Olivier; Tanarro, Luis M. (2024). "The European glacial landscapes from the Late Holocene". European Glacial Landscapes. Elsevier. p. 569. doi:10.1016/b978-0-323-99712-6.00025-8. ISBN 978-0-323-99712-6.
  155. ^ a b Dätwyler et al. 2017, p. 2336.
  156. ^ Dätwyler et al. 2017, pp. 2321–2322.
  157. ^ Mark, Samuel Z.; Abbott, Mark B.; Rodbell, Donald T.; Moy, Christopher M. (1 September 2022). "XRF analysis of Laguna Pallcacocha sediments yields new insights into Holocene El Niño development". Earth and Planetary Science Letters. 593: 7. Bibcode:2022E&PSL.59317657M. doi:10.1016/j.epsl.2022.117657. ISSN 0012-821X. S2CID 249813841.
  158. ^ Emile-Geay et al. 2008, p. 3141.
  159. ^ Du, Xiaojing; Hendy, Ingrid; Hinnov, Linda; Brown, Erik; Schimmelmann, Arndt; Pak, Dorothy (2020). "Interannual Southern California Precipitation Variability During the Common Era and the ENSO Teleconnection". Geophysical Research Letters. 47 (1): 8. Bibcode:2020GeoRL..4785891D. doi:10.1029/2019GL085891. ISSN 1944-8007.
  160. ^ Emile-Geay et al. 2008, p. 3144.
  161. ^ Dee, Sylvia G.; Cobb, Kim M.; Emile-Geay, Julien; Ault, Toby R.; Edwards, R. Lawrence; Cheng, Hai; Charles, Christopher D. (27 March 2020). "No consistent ENSO response to volcanic forcing over the last millennium". Science. 367 (6485): 1477–1481. Bibcode:2020Sci...367.1477D. doi:10.1126/science.aax2000. ISSN 0036-8075. PMID 32217726. S2CID 214671146.
  162. ^ Yan, Qing; Korty, Robert; Zhang, Zhongshi (September 2015). "Tropical Cyclone Genesis Factors in a Simulation of the Last Two Millennia: Results from the Community Earth System Model". Journal of Climate. 28 (18): 7185. Bibcode:2015JCli...28.7182Y. doi:10.1175/jcli-d-15-0054.1. ISSN 0894-8755.
  163. ^ Wallace, E. J.; Donnelly, J. P.; Hengstum, P. J.; Wiman, C.; Sullivan, R. M.; Winkler, T. S.; d'Entremont, N. E.; Toomey, M.; Albury, N. (27 November 2019). "Intense Hurricane Activity Over the Past 1500 Years at South Andros Island, The Bahamas". Paleoceanography and Paleoclimatology. 34 (11): 15–16. Bibcode:2019PaPa...34.1761W. doi:10.1029/2019PA003665.
  164. ^ Hernández, Armand; Martin-Puertas, Celia; Moffa-Sánchez, Paola; Moreno-Chamarro, Eduardo; Ortega, Pablo; Blockley, Simon; Cobb, Kim M.; Comas-Bru, Laia; Giralt, Santiago; Goosse, Hugues; Luterbacher, Jürg; Martrat, Belen; Muscheler, Raimund; Parnell, Andrew; Pla-Rabes, Sergi; Sjolte, Jesper; Scaife, Adam A.; Swingedouw, Didier; Wise, Erika; Xu, Guobao (1 October 2020). "Modes of climate variability: Synthesis and review of proxy-based reconstructions through the Holocene". Earth-Science Reviews. 209: 20. Bibcode:2020ESRv..20903286H. doi:10.1016/j.earscirev.2020.103286. hdl:10261/221475. ISSN 0012-8252. S2CID 225632127.
  165. ^ Swingedouw et al. 2017, p. 41.
  166. ^ Toker, E.; Sivan, D.; Stern, E.; Shirman, B.; Tsimplis, M.; Spada, G. (January 2012). "Evidence for centennial scale sea level variability during the Medieval Climate Optimum (Crusader Period) in Israel, eastern Mediterranean". Earth and Planetary Science Letters. 315–316: 52. Bibcode:2012E&PSL.315...51T. doi:10.1016/j.epsl.2011.07.019.
  167. ^ a b c Newhall, Self & Robock 2018, p. 576.
  168. ^ Gangadharan, Nidheesh; Goosse, Hugues; Parkes, David; Goelzer, Heiko; Maussion, Fabien; Marzeion, Ben (17 October 2022). "Process-based estimate of global-mean sea-level changes in the Common Era". Earth System Dynamics. 13 (4): 1423. Bibcode:2022ESD....13.1417G. doi:10.5194/esd-13-1417-2022. ISSN 2190-4979. S2CID 249090169.
  169. ^ Michel, Simon; Swingedouw, Didier; Chavent, Marie; Ortega, Pablo; Mignot, Juliette; Khodri, Myriam (3 March 2020). "Reconstructing climatic modes of variability from proxy records using ClimIndRec version 1.0". Geoscientific Model Development. 13 (2): 852. Bibcode:2020GMD....13..841M. doi:10.5194/gmd-13-841-2020. ISSN 1991-959X.
  170. ^ 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.
  171. ^ Knudsen, Karen Luise; Sha, Longbin; Zhao, Meixun; Seidenkrantz, Marit-Solveig; Björck, Svante; Jiang, Hui; Li, Tiegang; Li, Dongling (1 January 2018). "East Asian Winter Monsoon Variations and Their Links to Arctic Sea Ice During the Last Millennium, Inferred From Sea Surface Temperatures in the Okinawa Trough". Paleoceanography and Paleoclimatology. 33 (1): 68. Bibcode:2018PaPa...33...61L. doi:10.1002/2016PA003082. ISSN 2572-4525. S2CID 210097561.
  172. ^ Sanchez, Sara C.; Amaya, Dillon J.; Miller, Arthur J.; Xie, Shang-Ping; Charles, Christopher D. (10 April 2019). "The Pacific Meridional Mode over the last millennium". Climate Dynamics. 53 (5–6): 4. Bibcode:2019ClDy...53.3547S. doi:10.1007/s00382-019-04740-1. ISSN 1432-0894. S2CID 146254012.
  173. ^ Sousa, Pedro M.; Ramos, Alexandre M.; Raible, Christoph C.; Messmer, M.; Tomé, Ricardo; Pinto, Joaquim G.; Trigo, Ricardo M. (1 January 2020). "North Atlantic Integrated Water Vapor Transport—From 850 to 2100 CE: Impacts on Western European Rainfall". Journal of Climate. 33 (1): 267. Bibcode:2020JCli...33..263S. doi:10.1175/JCLI-D-19-0348.1. ISSN 0894-8755.
  174. ^ Stevenson et al. 2019, p. 1535.
  175. ^ Stevenson et al. 2019, p. 1548.
  176. ^ Zhang, Xuanze; Peng, Shushi; Ciais, Philippe; Wang, Ying-Ping; Silver, Jeremy D.; Piao, Shilong; Rayner, Peter J. (19 June 2019). "Greenhouse Gas Concentration and Volcanic Eruptions Controlled the Variability of Terrestrial Carbon Uptake Over the Last Millennium". Journal of Advances in Modeling Earth Systems. 11 (6): 1724. Bibcode:2019JAMES..11.1715Z. doi:10.1029/2018MS001566. PMC 6774283. PMID 31598188.
  177. ^ Banerji, Upasana S.; Padmalal, D. (1 January 2022). "12 – Bond events and monsoon variability during Holocene—Evidence from marine and continental archives". Holocene Climate Change and Environment. Elsevier: 322. doi:10.1016/B978-0-323-90085-0.00016-4. ISBN 9780323900850. S2CID 244441781.
  178. ^ a b Misios et al. 2022, p. 819.
  179. ^ Misios et al. 2022, p. 816.
  180. ^ Dai, Zhangqi; Wang, Bin; Zhu, Ling; Liu, Jian; Sun, Weiyi; Li, Longhui; Lü, Guonian; Ning, Liang; Yan, Mi; Chen, Kefan (9 September 2022). "Atlantic multidecadal variability response to external forcing during the past two millenniums". Journal of Climate. -1 (aop): 7. Bibcode:2022JCli...35.4503D. doi:10.1175/JCLI-D-21-0986.1. ISSN 0894-8755. S2CID 252249527.
  181. ^ 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. 18: EPSC2016–15250. Bibcode:2016EGUGA..1815250G.
  182. ^ 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.
  183. ^ Herweijer, Celine; Seager, Richard; Cook, Edward R.; Emile-Geay, Julien (April 2007). "North American Droughts of the Last Millennium from a Gridded Network of Tree-Ring Data". Journal of Climate. 20 (7): 1355. Bibcode:2007JCli...20.1353H. CiteSeerX 10.1.1.466.4049. doi:10.1175/jcli4042.1. ISSN 0894-8755. S2CID 129185669.
  184. ^ 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. S2CID 59361457.
  185. ^ Büntgen, Ulf; Urban, Otmar; Krusic, Paul J.; Rybníček, Michal; Kolář, Tomáš; Kyncl, Tomáš; Ač, Alexander; Koňasová, Eva; Čáslavský, Josef; Esper, Jan; Wagner, Sebastian; Saurer, Matthias; Tegel, Willy; Dobrovolný, Petr; Cherubini, Paolo; Reinig, Frederick; Trnka, Miroslav (April 2021). "Recent European drought extremes beyond Common Era background variability". Nature Geoscience. 14 (4): 194. Bibcode:2021NatGe..14..190B. doi:10.1038/s41561-021-00698-0. ISSN 1752-0908. S2CID 232237182.
  186. ^ Büntgen et al. 2022, p. 543.
  187. ^ a b Alloway et al. 2017, p. 98.
  188. ^ a b Mutaqin & Lavigne 2019, p. 2.
  189. ^ Mutaqin & Lavigne 2019, p. 4.
  190. ^ Lavigne et al. 2013, Supporting Information.
  191. ^ Hamilton 2013, p. 41.
  192. ^ Malawani et al. 2022, p. 8.
  193. ^ Mutaqin & Lavigne 2019, p. 9.
  194. ^ Kholis, Muhammad Arsyad Nur; Kurnia, Wahyu (26 November 2021). "Suling Dewa Sebagai Identitas Simbolik Masyarakat Sasak Kuto-Kute di Karang Bajo Bayan Lombok Utara". Jurnal Kajian Seni (in Indonesian). 8 (1): 19. doi:10.22146/jksks.64498. ISSN 2356-3001. S2CID 247378729.
  195. ^ a b Reid, Anthony (16 January 2017). "Population history in a dangerous environment: How important may natural disasters have been?". Masyarakat Indonesia. 39 (2): 520. ISSN 2502-5694. Archived from the original on 19 October 2018. Retrieved 18 October 2018.
  196. ^ Reid, Anthony (2016). "Building Cities in a Subduction Zone: Some Indonesian Dangers". In Miller, Michelle Ann; Douglass, Mike (eds.). Disaster Governance in Urbanising Asia. Springer Singapore. p. 51. doi:10.1007/978-981-287-649-2_3. ISBN 978-981-287-649-2.
  197. ^ Malawani et al. 2022, p. 11.
  198. ^ Mutaqin & Lavigne 2019, p. 7–8.
  199. ^ Anderson, Atholl (2016). The First Migration: Māori Origins 3000 BC – AD 1450. Bridget Williams Books. p. 18. ISBN 9780947492809.
  200. ^ Ludlow, Francis (2017). "Volcanology: Chronicling a medieval eruption". Nature Geoscience. 10 (2): 78–79. Bibcode:2017NatGe..10...78L. doi:10.1038/ngeo2881. ISSN 1752-0908.
  201. ^ Stothers 2000, p. 363.
  202. ^ 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. Vol. 139. Washington DC American Geophysical Union Geophysical Monograph Series. p. 259. Bibcode:2003GMS...139..255D. doi:10.1029/139GM16. ISBN 978-0-87590-998-1.
  203. ^ a b Dodds & Liddy 2011, p. 54.
  204. ^ Frey Sánchez, Antonio Vicente (2017). "¿Qué puede aportar el clima a la historia? El ejemplo del periodo cálido medieval en el Magreb almorávide y almohade". El Futuro del Pasado: Revista Electrónica de Historia (in Spanish). 6 (8): 221–266. doi:10.14516/fdp.2017.008.001.008. ISSN 1989-9289. Archived from the original on 20 October 2018. Retrieved 20 October 2018.
  205. ^ Grillo 2021, p. 150.
  206. ^ Guillet et al. 2017, p. 124.
  207. ^ a b Guillet et al. 2017, p. 127.
  208. ^ a b Stothers 2000, p. 366.
  209. ^ Bufanio 2022, p. 23.
  210. ^ a b Bufanio 2022, p. 25.
  211. ^ Stothers 2000, p. 364.
  212. ^ Frey Sánchez, Antonio Vicente (31 December 2014). "Ciudades y poder político en al-Andalus. Una hipótesis sobre el origen de las revueltas urbanas en Murcia en el siglo XIII". Anuario de Estudios Medievales (in Spanish). 44 (2): 854. doi:10.3989/aem.2014.44.2.06. ISSN 1988-4230.
  213. ^ a b John Gillingham (2014). Conquests, Catastrophe and Recovery: Britain and Ireland 1066–1485. Random House. p. 26. ISBN 978-1-4735-2233-6.
  214. ^ Campbell 2017, p. 91.
  215. ^ Bufanio 2022, p. 27.
  216. ^ Campbell 2017, p. 108.
  217. ^ Campbell 2017, p. 119.
  218. ^ Speed, Robert; Tickner, David; Lei, Gang; Sayers, Paul; Wei, Yu; Li, Yuanyuan; Moncrieff, Catherine; Pegram, Guy (2016). Drought risk management: a strategic approach. UNESCO Publishing. p. 44. ISBN 978-92-3-100094-2.
  219. ^ Bufanio 2022, pp. 23, 25.
  220. ^ Bufanio 2022, p. 26.
  221. ^ Moglia 2022, p. 53.
  222. ^ Degroot, Dagomar; Anchukaitis, Kevin; Bauch, Martin; Burnham, Jakob; Carnegy, Fred; Cui, Jianxin; de Luna, Kathryn; Guzowski, Piotr; Hambrecht, George; Huhtamaa, Heli; Izdebski, Adam; Kleemann, Katrin; Moesswilde, Emma; Neupane, Naresh; Newfield, Timothy; Pei, Qing; Xoplaki, Elena; Zappia, Natale (March 2021). "Towards a rigorous understanding of societal responses to climate change". Nature. 591 (7851): 545–546. Bibcode:2021Natur.591..539D. doi:10.1038/s41586-021-03190-2. ISSN 1476-4687. PMID 33762769. S2CID 232354348.
  223. ^ Domingues, Lidia L. Zanetti (30 October 2022). "Carestia, maltempo e alleanze politiche: Siena e Manfredi di Sicilia fra 1257 e 1260". Studi di storia medioevale e di diplomatica. Nuova Serie (in Italian): 104. doi:10.54103/2611-318X/18283. ISSN 2611-318X.
  224. ^ Bortoluzzi, Daniele (30 October 2022). "Bologna e gli Ordinamenta Bladi". Studi di storia medioevale e di diplomatica. Nuova Serie (in Italian): 89. doi:10.54103/2611-318X/18282. ISSN 2611-318X.
  225. ^ Moglia 2022, p. 52.
  226. ^ Moglia 2022, p. 55.
  227. ^ Moglia 2022, p. 58.
  228. ^ Bertoni 2022, p. 37.
  229. ^ Bertoni 2022, p. 39.
  230. ^ Grillo 2021, p. 153.
  231. ^ Grillo 2021, p. 154.
  232. ^ Luongo 2022, p. 76.
  233. ^ Luongo 2022, p. 77.
  234. ^ Luongo 2022, p. 63.
  235. ^ Stothers 2000, pp. 367–368.
  236. ^ Harrison & Maher 2014, pp. 156–157.
  237. ^ Harrison & Maher 2014, p. 180.
  238. ^ 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" (PDF). Quaternary Science Reviews. 136: 229–252. Bibcode:2016QSRv..136..229X. doi:10.1016/j.quascirev.2015.10.004.
  239. ^ a b Matson, R.G. (February 2016). "The nutritional context of the Pueblo III depopulation of the northern San Juan: Too much maize?". Journal of Archaeological Science: Reports. 5: 622–624. Bibcode:2016JArSR...5..622M. doi:10.1016/j.jasrep.2015.08.032. ISSN 2352-409X.
  240. ^ Windes, Thomas C.; Van West, Carla R. (2021), Van Dyke, Ruth M.; Heitman, Carrie C. (eds.), "Landscapes, Horticulture, and the Early Chacoan Bonito Phase", The Greater Chaco Landscape, Ancestors, Scholarship, and Advocacy, University Press of Colorado, p. 83, ISBN 978-1-64642-169-5, JSTOR j.ctv1m46ffr.6, retrieved 10 December 2021
  241. ^ Salzer 2000, p. 308.
  242. ^ a b Salzer 2000, pp. 312–314.
  243. ^ Cruz, Pablo; Winkel, Thierry; Ledru, Marie-Pierre; Bernard, Cyril; Egan, Nancy; Swingedouw, Didier; Joffre, Richard (1 December 2017). "Rain-fed agriculture thrived despite climate degradation in the pre-Hispanic arid Andes". Science Advances. 3 (12): 5. Bibcode:2017SciA....3E1740C. doi:10.1126/sciadv.1701740. ISSN 2375-2548. PMC 5738230. PMID 29279865.
  244. ^ Guillet et al. 2017, p. 125.
  245. ^ Jenkins 2021, p. 63.
  246. ^ Molnar, Aaron (June 2023). "Felled Forests and Fallowed Fields: Establishing a Narrative of Ecological and Climate Change in Mongol-Era Goryeo". Seoul Journal of Korean Studies. 36 (1): 225–226. doi:10.1353/seo.2023.a902140. S2CID 259928765.
  247. ^ Jenkins 2021, p. 82.
  248. ^ Bufanio 2022, p. 22.
  249. ^ Timmreck et al. 2009, p. 1.
  250. ^ Alloway et al. 2017, p. 96.
  251. ^ Di Cosmo, Wagner & Büntgen 2021, p. 92.
  252. ^ Di Cosmo, Wagner & Büntgen 2021, p. 97.
  253. ^ Di Cosmo, Wagner & Büntgen 2021, p. 100.
  254. ^ a b Kern, Zoltán; Pow, Stephen; Pinke, Zsolt; Ferenczi, László (1 April 2021). Samalas and the Fall of the Mongol Empire: A volcanic eruption's influence on the dissolution of history's largest contiguous empire. 23rd EGU General Assembly. EGU General Assembly Conference Abstracts. pp. EGU21–3460. Bibcode:2021EGUGA..23.3460K.
  255. ^ Hao, Zhixin; Zheng, Jingyun; Yu, Yingzhuo; Xiong, Danyang; Liu, Yang; Ge, Quansheng (1 October 2020). "Climatic changes during the past two millennia along the Ancient Silk Road". Progress in Physical Geography: Earth and Environment. 44 (5): 619–620. Bibcode:2020PrPG...44..605H. doi:10.1177/0309133319893919. ISSN 0309-1333. S2CID 213726073.
  256. ^ Fell et al. 2020, p. 41.
  257. ^ Fell et al. 2020, p. 42.
  258. ^ Fell et al. 2020, p. 40.
  259. ^ Fell et al. 2020, p. 43.

Sources

External links

This page was last edited on 4 March 2024, at 04:17
Basis of this page is in Wikipedia. Text is available under the CC BY-SA 3.0 Unported License. Non-text media are available under their specified licenses. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc. WIKI 2 is an independent company and has no affiliation with Wikimedia Foundation.