Mount Churchill

Summary

Mount Churchill is a dormant volcano in the Saint Elias Mountains and the Wrangell Volcanic Field (WVF) of eastern Alaska. Churchill and its neighbor Mount Bona are both ice-covered volcanoes with Churchill having a 2.7-by-4.2-kilometre-wide (1.7 mi × 2.6 mi) caldera just east of its summit. There are sparse outcrops of lava flows and tephra, mostly dacite.

Mount Churchill
U.S. Geological Survey climbing party ascending the Klutlan Glacier en route to Mount Churchill
Highest point
Elevation4,765 to 4,767 m (15,633 to 15,640 ft) Edit this on Wikidata
Prominence1,188 ft (362 m)
Coordinates61°25′9″N 141°42′55″W / 61.41917°N 141.71528°W / 61.41917; -141.71528
Geography
Parent rangeSaint Elias Mountains
Topo mapUSGS McCarthy B-2 Quadrangle
Geology
Mountain typeStratovolcano with caldera
Volcanic fieldWrangell Volcanic Field
Last eruption700 AD ± 200 years
Climbing
First ascentAugust 20, 1951 by R. Gates, J. Lindberg
Easiest routesnow climb

Subduction of the Pacific Plate beneath southeastern Alaska has largely ceased during the last one million years, causing a decline of the volcanic activity in the WVF. Churchill appears to be fed by melts derived from a stagnant slab in the mantle, left over by the previous subduction.

The volcano erupted several times during the Holocene. The most notable eruptions are the two White River Ash eruptions, deposited during two of the largest volcanic eruptions in North America during the past two millennia. The northern lobe was emplaced about 1,890 years ago, while the larger eastern lobe erupted in winter 852/853. The White River Ash covers vast expanses of Alaska and western Canada and has been found as far as Europe, and there is evidence that the Athabaskan people migrated out of the region and into the present-day United States as a consequence of the eruption.

Geography and geomorphology edit

The mountain is in the University Mountains sub-range[1] of the St. Elias Mountains of Alaska,[2] 64 kilometres (40 mi) east of McCarthy, Alaska,[3] and 25 kilometres (16 mi)[4] or 40 kilometres (25 mi) from the border with Canada.[5] The area is part of the Wrangell-St. Elias National Park and Preserve.[6] It is extremely remote[7] and there are no roads from which it is visible.[3] The mountain was first ascended in 20 August 1951 by R. Gates and J. Lindberg[8] and named in 1965 after the English politician Winston Churchill[9] and is also known as Klutlan Glacier, Churchill-Bona, or White River volcano.[10]

Various measurements have yielded summit heights[a] of 4,744 metres (15,564 ft),[9] 4,766 metres (15,636 ft),[2][12] 4,767 metres (15,640 ft)[6] or 4,768 metres (15,643 ft).[13] It is a mountain in a glaciated,[b] rugged mountain massif[15] that rises sharply above the surrounding land.[16] It is the tenth-highest peak in the United States.[17] The mountain is mostly covered by[18] ice hundreds of meters thick,[19] but lava flows with columnar jointing and tephra deposits form outcrops,[20] indicating that Mount Churchill may be a stratovolcano.[21] East of Mount Churchill,[15] 300 metres (980 ft) below the summit,[22] is a 2.7-by-4.2-kilometre-wide (1.7 mi × 2.6 mi) caldera,[11] which forms a flat amphitheater open to the northeast. Numerous outcrops of light-colored pumice with embedded lithics[c] occur around the amphitheater,[15] which is otherwise entirely ice-covered.[24] There are further outcrops of tephra in areas protected from erosion around the volcano; the largest such outcrop covers an area exceeding 3 square kilometres (1.2 sq mi).[25] Pumice forms terraces above the sides of the Klutlan Glacier, over a length of more than 20 kilometres (12 mi).[26] Their position above the present-day glacier surface may indicate that at the time of their deposition, the ice was thicker than present-day. Alternatively,[27] they could have deposited during floods over the ice, perhaps after an eruption or the breach of a pumice-dammed lake.[28] A 90-metre-high (300 ft) pumice mound on the other side of the glacier,[29] 16 kilometres (9.9 mi) from Mount Churchill,[30] was formed by tephra building up on a bedrock bench.[31] It was once considered the vent of the White River Ash.[30] In some places, 1-metre-thick (3.3 ft) volcanic ash covers the pumice.[27]

The older[11] Mount Bona is 3.2 kilometres (2 mi) southwest of Mount Churchill.[9] With a summit height of 5,005 metres (16,421 ft)[d] above sea level,[2] it is the highest mountain in the Wrangell Mountains[33][34] and the highest volcano in the Wrangell volcanic field[35] and the United States in general.[36] A snow-covered col at 4,400 metres (14,400 ft) elevation separates the two mountains.[32] Both mountains are covered with about 5 cubic kilometres (1.2 cu mi) of ice.[37] The Russell and Klutlan Glaciers run along the northern-western and eastern[38]-southern side of Mount Churchill, respectively.[38][25] The Klutlan Glacier is flanked by moraines and talus deposits.[39] Both glaciers eventually discharge into the White River.[6] Glaciers on the southern flank of Mount Bona discharge into the Chitina River.[37] Ice on Mount Churchill is up to 800, and possibly 1500, years old.[19]

Geology edit

The more than 500-kilometre-long (310 mi)[40] Wrangell volcanic field (WVF) has been active for the past 30 million years[41] in the Wrangell and St. Elias Mountains.[42] The Wrangell volcanic field features numerous large shield volcanoes,[43] which are among the largest arc volcanoes on Earth.[44] Mount Drum[45] and other volcanoes in the WVF during the middle Pleistocene had eruptions even larger than the White River Ash eruptions.[46] Mount Churchill[3] and Mount Wrangell are the only volcanoes in the WVF with Holocene eruptions.[47] With the exception of Mount Churchill, volcanism in the Wrangell volcanic field has migrated northwestwards[48] and declined as[49] plate configuration changed about 200,000 years ago[50] and subduction ceased.[49]

Mount Churchill and Mount Bona consist of andesitic lava flows. University Peak is a 8.4 million years old volcanic intrusion, now exposed through erosion.[51] The basement under Mount Bona is formed by a plateau consisting of Permian to Pennsylvanian-age rocks[52] and Tertiary granites; most of Mount Bona may be formed by these nonvolcanic rocks.[53]

Off the western coast of southeastern Alaska, the Pacific Plate used to subduct under the North American Plate, giving rise to the WVF.[42] Since the Jurassic,[54] seven separate terranes were transported to Alaska by the Pacific Plate and attached to the continent:[55] Windy terrane, the various Wrangellia terranes,[54] Chugach, Prince William and most recently the Yakutat Block, which is in the process of being accreted.[42] The collision with the Yakutat Block caused the cessation of subduction, with plate motion now occurring along strike-slip faults like the Denali and Totschunda Faults[42] while subduction continues farther west in the Aleutian megathrust.[56] The intersection of the Totschunda Fault with the Connector and Duke River fault(s) may be the point where magma ascends into Mount Churchill.[41]

Composition and origin of magmas edit

Churchill rocks are dacitic[e] and define a calc-alkaline[59] adakite suite.[60] There is a moderate quantity of phenocrysts, including biotite, hornblende, ilmenite, hypersthene, magnetite and plagioclase, with little apatite and orthopyroxene.[61][62][63] Several different rock chemistries contribute to each of the White River Ash lobes,[26] which are otherwise very similar to each other[64] and thus difficult to distinguish.[65] The particles in the eastern lobe are coarser than in the northern[66] and show evidence of two separate chemical trends;[60] the deposits on Mount Churchill match the composition of the eastern lobe.[64] Reconstructed magma temperatures are 950–990 °C (1,740–1,810 °F) for the northern lobe magma and 995–1,030 °C (1,823–1,886 °F) for the eastern lobe magma.[4]

The Wrangell slab[67] left over from the subduction may have stalled in the mantle, and was heated by asthenosphere flowing through a slab window until it melted and gave rise to the Mount Churchill magmas,[68][69] which thus have an adakitic composition typical for melts derived from subducted basalts at high temperatures.[70] During ascent, the magmas were further modified by interaction with the underlying basement[71] of the Alexander terrane.[72] Each of the White River Ash eruptions probably involved several different magma batches, rather than one layered magma chamber.[68][69]

Ice cores and climate edit

Several ice cores have been taken from the Bona-Churchill massif[73] and are an important source of information on the climate of the Pacific Northwest.[74] An ice core taken in 2002 from the col between Mount Churchill and Mount Bona[75] is the longest non-polar ice core as of 2006,[76] being 460.96 metres (1,512.3 ft) long.[77]

The ice cores record evidence of volcanic eruptions, including of Katmai, Krakatau, Laki and Tambora, and of climate variations like the Medieval Warm Period and the Little Ice Age. Other processes recorded in the Bona-Churchill ice cores are dust emissions in China,[73] wildfires in Alaska,[78] North Pacific sea surface temperatures, position of the Aleutian Low weather system[79] and Arctic sea ice cover.[80] Shallow ice and snow has been used to reconstruct dust composition at the St. Elias Mountains.[81]

The Chugach Mountains block the maritime airmasses, leading to a continental climate in the region.[16] Mean annual temperature on Mount Churchill is about −23 °C (−9 °F).[82] Annually, about 1 metre (3 ft 3 in) of snow water equivalent falls on Mount Churchill.[73]

Eruption history edit

The age of the Churchill-Bona massif is unknown[51] but Mount Churchill began erupting during the late Pleistocene. Potassium-argon dating has yielded an age of 119,000±17,000 years for a dacite lava close to the summit.[15] The 190,000 years old[83] Sheep Creek tephra sub-unit "F"[84] in Canada and Alaska[85] may have originated at Mount Churchill, but more likely at Mount Drum.[84] The appearance and height of Mount Churchill (and neighbouring Bona) imply that they were constructed in recent time.[86] The mountain may have looked very different before the White River Ash eruptions.[4]

There are six Holocene volcanic eruptions that may be attributed to Mount Churchill.[69] Ash emplaced around 647±55 CE may come either from Mount Churchill or Redoubt volcano,[87] and European tephras emplaced around 2,350 BCE and Greenland-Europe tephras from around 1100 CE resemble these of Mount Churchill.[88][89] Two tephra layers in southeastern Alaska, the 300 years old "Lena ash" and the 6,330 years old "MTR-146" ash, resemble the White River Ash[90] and may have been produced by eruptions of Mount Churchill;[91][92] tephra with similar composition to the "Lena tephra" has been found in Europe.[89] If the 1650 CE "Lena ash" comes from the volcano, it would be its youngest eruption.[93] Beyond these, volcanic activity was uncommon in the region.[94]

White River Ash eruptions edit

Mount Churchill is the source[f] of two of the largest volcanic eruptions of the past two millennia in North America.[2] The first eruption about 1,890 years ago emplaced the northern lobe of the White River Ash ("Northern White River Ash"),[2] the better known second eruption[98] in winter 852/853[g][99] emplaced the eastern lobe ("Eastern White River Ash").[2] Both were very violent[4] Plinian eruptions[62] with a volcanic explosivity index of 6.[101]

Deposits from the eruption were first discovered in 1883 along the upper Yukon River. After Mount Wrangell had been ruled out as its source in 1892, Mount Natazhat was proposed instead as the source vent and in 1965 Mount Bona. Only in 1984 and 1995 was Mount Churchill identified as the source.[7] The eruptions produced about 25–50 cubic kilometres (6.0–12.0 cu mi) of tephra[2] and covered an area exceeding 540,000 square kilometres (210,000 sq mi)[102] in Alaska,[h] Yukon Territory and Northwest Territories.[104][105] The present-day towns of Dawson City and Whitehorse, Canada, are within the 25-millimetre (1 in) thickness area of the northern and eastern lobe, respectively.[30] The ash is located at shallow depth in the ground,[106] unless carried deeper underground by soil processes.[i][108] It forms conspicuous layers along the Alaska Highway,[105] in riverbanks[109] of the Yukon, Tanana and their tributaries.[110] The ash layers affect the properties of the soil they are in;[111] they contribute to the formation of soils[112] and sometimes they are the detachment surface of landslides.[111] Closer to the US-Alaska border at the Klutlan Glacier it thickens to form dune-covered ash fields[113] and areas lacking vegetation, as the ash is an unsuitable ground for plant growth.[114] Stumps of trees killed by the fallout emerge from the ash layers close to Mount Churchill.[115] Ash is frequently reworked and redeposited,[66][116] and forms soils in the St. Elias Mountains.[117] Glaciers such as the Barnard and Klutlan Glaciers have captured and transported pumice and ash,[118][119] or eroded ash layers when they advanced;[120] some moraines at the foot of the St. Elias Mountains are formed mainly by White River Ash.[121] Ash is washed away by the Klutlan and White River, contributing in no small part (together with glacial flour) to its distinctive color that gives the White River its name.[122] The ash deposits have been used as a time marker in tephrochronology to obtain dates for natural events and archaeological sites[123] from Alaska and Yukon[124] as well as Greenland (correlation of ice cores)[125] and Ireland in Europe.[126]

Apart from direct physical effects, the Mount Churchill eruption likely had a strong psychological effect on the people in the affected area. The eruption column would have been visible for many hundreds of kilometers. Soon after it began, the sky would have turned dark for days and noise and lightning would have been heard and seen in Yukon.[127][128] About 500 people might have been living in the directly affected area.[129] It is probable that there were no direct casualties from the eruption; pyroclastic flows and other direct effects of the eruption were limited to the uninhabited surroundings of Mount Churchill, and the structures humans lived in at the time were unlikely to collapse under ash accumulation.[130][94] There is disagreement on whether oral tradition referring to the White River Ash eruptions can be identified among the Athapaskans.[131][132][14]

Northern White River Ash edit

The northern White River Ash extends along the Alaska–Canada border[133] and reaches a thickness of 5–10 centimetres (2.0–3.9 in) 380 kilometres (240 mi) west of the volcano, declining to 2.5 centimetres (0.98 in) 580 kilometres (360 mi) north of Mount Churchill.[134] The White River Ash is a formal stratigraphic unit in Alaska,[135] and particles from it have been detected as far as the northern Brooks Range in Alaska.[65] The widespread "PWS tephra" in Prince William Sound was emplaced between 2,039 and 1,520 years ago and resembles the northern White River Ash.[136] The eruption may have occurred during summer, when winds blow from the south,[137][128] and the eruption column might have been 30–35 kilometres (19–22 mi) high.[4] While not as well studied as the east lobe eruption,[138] its impact on human populations was relatively modest, with few signs of population or culture shifts.[139][140]

Eastern White River Ash edit

The eastern White River Ash is better studied[134] and was larger.[2] Its intensity was intermediate between the Mount Mazama eruption and the 1883 eruption of Krakatau.[141] It was more than twice the size of the 1912 Novarupta/Katmai eruption[142] and was ten times larger than the 1991 eruption of Mount Pinatubo.[143] A 40–45-kilometre-high (25–28 mi) eruption column rose over the volcano,[99] injecting ash into the stratosphere;[144] ash fell more than a thousand kilometers away[99] and sulfate and chloride precipitated in the Greenland Ice Sheet.[100] Strong westerly winds carried the ash cloud eastward,[145] where it may have mixed with snow as it fell out.[128] The eastern lobe of the White River Ash is 2.5 centimetres (0.98 in) thick at 600 kilometres (370 mi) distance from Mount Churchill,[98] extending to the Great Slave and Great Bear Lakes.[5] The eastern White River Ash has a color ranging from white to beige.[66]

Ash deposits from the eastern White River Ash have been detected across North America and into Europe, where it is identical to the "AD860B" ash found in Ireland,[99] Great Britain, Scandinavia, Germany, Poland[j][148] and Greenland. Other findings are in Nova Scotia,[149] south-central Alaska,[150] southeastern Alaska and the adjacent Pacific Ocean,[151] Newfoundland[152] and Maine.[145] These findings 7,000 kilometres (4,300 mi) from the volcano make the White River Ash one of the most extensive tephra deposits of the past 100,000 years,[153] and drew attention to the potential for intercontinental spread of volcanic ash[154] even by once-per-century eruptions.[155]

Territories impacted by the ashfall may have needed decades to recover,[156] with century-long changes in vegetation, aquatic and peat productivity[99] as forests opened up in some areas with ashfall.[157] In lakes, volcanic ash can either bury organisms,[133] or release nutrients such as phosphorus and thus increase productivity; both effects have been noted for the White River Ash.[158] Burial of food sources and ingestion of ash and fluoride would have impacted caribou, goats, moose and sheep populations,[159] forcing them to move away; genomic data indicate a large shift in caribou populations after the eastern White River Ash eruption,[160] although this theory is not uncontested.[161] Ash fall into rivers and the remobilization of ash fallen on land would have disrupted waterfowl, salmon runs and other fish populations,[162] although anadromous fish populations would have recovered within a short timeframe.[163]

Southern Yukon was depopulated by the eruption.[99] Local hunter-gatherer populations probably left the worst-hit areas and sought refuge in unaffected regions, returning only when conditions had improved[163] or not at all. Archaeological data indicate that some important trade routes were abandoned and new ones established after the eastern White River Ash eruption, implying that the displacement fostered a re-evaluation of economic activity and that displaced people had set up new trade networks.[164][165] The use of copper[k][167] and bows and arrows may have arrived in the Yukon territory that way,[168] and Dene people moved into coastal areas, sometimes coming into conflict with previously established people there and sometimes establishing new kin and commercial networks.[169] Other Dene people migrated south and east[l] after the eruption, driving the Athabaskan expansion and spreading the Na-Dene languages across the continent. By the arrival of the Europeans,[142][171][172][99] Athabaskans like the Apache[m] and Navajo[173][174] had spread between subarctic Canada and the Great Basin of the southwestern United States, bringing their languages with them.[175]

The eruption produced sulfate aerosols,[176] which can dim the Sun and cause a cooling of Earth's climate, creating a volcanic winter.[177] The sulfur yield, 2.5 teragrams, was relatively modest, one third of that from the 1991 eruption of Mount Pinatubo.[178] Climate models imply a maximum cooling of 0.3 °C (0.54 °F),[179] reaching 0.8 °C (1.4 °F) in some models,[180] with no clear changes in precipitation.[181] There are widespread reports of bad weather and resulting hardships such as famines during that decade in Europe,[182] and a clear link to the Mount Churchill eruption is not established;[183] at worst, it would have aggravated a pre-existent climate disturbance.[184] A link between the White River Ash and the mid-6th century cooling (Late Antique Little Ice Age) has been ruled out.[185]

Hazards edit

Mount Churchill is one of Canada's most dangerous volcanoes,[186] despite being outside of the country,[187] owing to the size of its eruptions. Renewed large-scale activity would be extremely hazardous for northwestern Canada and adjoining Alaska.[64] Smaller eruptions could threaten the White River valley and the Alaska Highway there[186] with ash fall and floods[37] caused by blockages in the White River.[188] Similar flood hazards exist in the Chitina and Copper River valleys south of Mount Churchill.[37] The United States Geological Service ranks Mount Churchill as a "high threat" volcano.[n][191]

Ashfall could damage machinery, forests and waterbodies, and cause breathing problems.[23] Even small eruptions of the high volcano could cause disturbances in air travel.[37] In addition, the intercontinental spread of ash would cause severe disruption, similar but on a larger scale to the 2010 eruptions of Eyjafjallajökull, with resultant consequences to transportation and the airline industry.[152] Aircraft routes between Asia, Europe and North America pass through the extent of the White River Ash plume.[192]

Notes edit

  1. ^ The value of 5,005 metres (16,421 ft) given by the Global Volcanism Program[11] actually refers to Mount Bona[2]
  2. ^ About 90% of the area is covered by snow and ice.[14]
  3. ^ Derived presumably from the magma conduit[23]
  4. ^ Sometimes its elevation is given as 5,029 metres (16,499 ft)[32]
  5. ^ The White River Ash has also been described as rhyodacitic[57] or rhyolitic.[58]
  6. ^ The absence of White River Ash in the Bona-Churchill ice core has been cited as an argument against it being the source,[95] but there is evidence of reworked White River Ash in the core.[96] Evidence is more definitive for the eastern lobe than the northern one.[97]
  7. ^ Between September 852[99] and January 853[100]
  8. ^ As far as the Brooks Range[103]
  9. ^ It can end up in early Holocene sediments, thus creating problems with tephrochronological correlations[107]
  10. ^ Where ash deposition is recorded through speleothems[146] in Kletno Bear Cave[147]
  11. ^ Erosion caused by the volcanic events may have led to the discovery of copper in the White River-Copper River area,[166] as stated by oral tradition[142]
  12. ^ Possibly also north[170]
  13. ^ Whose language probably began developing around the time of the eruption[173]
  14. ^ "High threat" is the second-highest in a five-class scale,[189] which considers both the threat posed by a volcano and the infrastructure/population/other human uses at risk.[190]

References edit

  1. ^ Wood & Coombs 2001, p. 144.
  2. ^ a b c d e f g h i Richter et al. 1995, p. 741.
  3. ^ a b c Richter, Rosenkrans & Steigerwald 1995, p. 29.
  4. ^ a b c d e West & Donaldson 2001, p. 240.
  5. ^ a b Bunbury et al. 2008, p. 78.
  6. ^ a b c Winkler, p. 3.
  7. ^ a b Richter et al. 1995, p. 742.
  8. ^ Wood & Coombs 2001, p. 161.
  9. ^ a b c Map 1981.
  10. ^ GVP 2024, Synonyms & Subfeatures.
  11. ^ a b c GVP 2024, General Information.
  12. ^ Wood & Coombs 2001, p. 160.
  13. ^ Miller & Richter 1994, p. 772.
  14. ^ a b Sheets & Grayson 1979, p. 349.
  15. ^ a b c d Richter et al. 1995, p. 743.
  16. ^ a b Wiles et al. 2002, p. 896.
  17. ^ Mantell & Mantell 1976, p. 18.
  18. ^ Westgate et al. 2008, p. 201.
  19. ^ a b Fang et al. 2023, p. 4015.
  20. ^ Preece et al. 2014, p. 1039.
  21. ^ Preece & Hart 2004, p. 168.
  22. ^ McGimsey et al. 1992, p. 243.
  23. ^ a b Natural Resources Canada 2009.
  24. ^ McGimsey et al. 1992, p. 214.
  25. ^ a b Richter et al. 1995, p. 744.
  26. ^ a b Donaldson, Guerstein & Mueller 1996, p. 1234.
  27. ^ a b Donaldson, Guerstein & Mueller 1996, p. 1241.
  28. ^ Donaldson & Mueller 1994.
  29. ^ McGimsey et al. 1992, p. 213.
  30. ^ a b c McGimsey et al. 1992, p. 212.
  31. ^ Preece et al. 2014, p. 1025.
  32. ^ a b Holdsworth & Roy Krouse 2002, p. 32.
  33. ^ Miller & Richter 1994, p. 771.
  34. ^ Moffit 1938, p. 9.
  35. ^ Richter et al. 1990, p. 29.
  36. ^ Ringsmuth 2015, p. 12.
  37. ^ a b c d e McGimsey et al. 1995.
  38. ^ a b Lerbekmo 2008, p. 695.
  39. ^ Donaldson, Guerstein & Mueller 1996, p. 1235.
  40. ^ Trop 2012, p. 807.
  41. ^ a b Trop et al. 2022, p. 37.
  42. ^ a b c d Wilson et al. 2005, p. 3.
  43. ^ Preece & Hart 2004, p. 166.
  44. ^ Brueseke et al. 2019, p. 61.
  45. ^ Preece & Hart 2004, p. 167.
  46. ^ Jensen et al. 2006.
  47. ^ Winkler, p. 110.
  48. ^ Richter et al. 1990, p. 40.
  49. ^ a b Richter et al. 1990, p. 42.
  50. ^ Trop et al. 2022, p. 41.
  51. ^ a b Winkler, p. 113.
  52. ^ Miller & Richter 1994, p. 775.
  53. ^ Lerbekmo 2008, p. 694.
  54. ^ a b Wilson et al. 2005, p. 2.
  55. ^ Wilson et al. 2005, p. 1.
  56. ^ Trop 2012, p. 806.
  57. ^ Smith et al. 1999, p. 603.
  58. ^ Strickland et al. 2005, p. 638.
  59. ^ McGimsey et al. 1992, p. 215.
  60. ^ a b Preece et al. 2014, p. 1040.
  61. ^ Preece et al. 2014, p. 1027.
  62. ^ a b Robinson 2001, p. 158.
  63. ^ Richter et al. 1995, p. 745.
  64. ^ a b c Richter et al. 1995, p. 747.
  65. ^ a b Davies, Jensen & Kaufman 2022, p. 131.
  66. ^ a b c West & Donaldson 2001, p. 242.
  67. ^ Liang et al. 2024, p. 33.
  68. ^ a b Preece et al. 2014, p. 1038.
  69. ^ a b c Davies et al. 2016, p. 47.
  70. ^ Preece & Hart 2004, p. 184.
  71. ^ Preece & Hart 2004, p. 187.
  72. ^ Westgate et al. 2008, p. 202.
  73. ^ a b c Thompson et al. 2004.
  74. ^ Campbell et al. 2012, p. 100.
  75. ^ Lerbekmo 2008, pp. 693–694.
  76. ^ Bowen 2006, Alaska's Bona-Churchill.
  77. ^ Sierra-Hernández et al. 2022, p. 2.
  78. ^ Sierra-Hernández et al. 2022, p. 8.
  79. ^ Porter et al. 2017.
  80. ^ Porter, Mosley-Thompson & Thompson 2019, p. 10797.
  81. ^ Hinkley 1994, p. 3246.
  82. ^ Porter, Mosley-Thompson & Thompson 2019, p. 10786.
  83. ^ Westgate et al. 2008, p. 203.
  84. ^ a b Lowe 2011, p. 131.
  85. ^ Lowe 2011, p. 121.
  86. ^ Miller & Richter 1994, p. 773.
  87. ^ Zander et al. 2013, pp. 766, 768.
  88. ^ Plunkett et al. 2023, p. 8.
  89. ^ a b Plunkett & Pilcher 2018, p. 27.
  90. ^ Preece et al. 2014, p. 1020.
  91. ^ Payne, Blackford & van der Plicht 2008, p. 53.
  92. ^ Davies, Jensen & Kaufman 2022, p. 136.
  93. ^ Payne, Blackford & van der Plicht 2008, p. 51.
  94. ^ a b Sheets & Grayson 1979, p. 348.
  95. ^ Addison et al. 2010, p. 286.
  96. ^ Urmann 2009, Abstract.
  97. ^ Scott 2003, p. 366.
  98. ^ a b Reuther et al. 2020, p. 169.
  99. ^ a b c d e f g h Mackay et al. 2022, p. 1476.
  100. ^ a b Mackay et al. 2022, p. 1477.
  101. ^ GVP 2024, Eruptive History.
  102. ^ Robinson 2001, p. 159.
  103. ^ Monteath et al. 2017, p. 173.
  104. ^ Kristensen, Ives & Supernant 2021, p. 428.
  105. ^ a b Richter, Rosenkrans & Steigerwald 1995, p. 27.
  106. ^ Lerbekmo & Campbell 1969, p. 109.
  107. ^ Beierle 2002.
  108. ^ Bond, MacGregor & Lipovsky 2001, p. 25.
  109. ^ Capps 1916, p. 59.
  110. ^ Capps 1916, p. 60.
  111. ^ a b Lollino et al. 2015, p. 453.
  112. ^ Holloway 2020, p. 28.
  113. ^ Capps 1916, p. 61.
  114. ^ Rampton 1978, p. 133.
  115. ^ Rampton 1971, p. 976.
  116. ^ Lerbekmo 2008, p. 696.
  117. ^ Birks 1977, p. 2371.
  118. ^ Lerbekmo & Campbell 1969, p. 113.
  119. ^ Capps 1916, p. 64.
  120. ^ Rampton 1970, p. 1241.
  121. ^ Rampton 1970, p. 1249.
  122. ^ Bailey & Burn 2023, p. 728.
  123. ^ Robinson 2001, p. 160.
  124. ^ Scott 2003, p. 363.
  125. ^ Plunkett et al. 2023, p. 3.
  126. ^ McClung & Plunkett 2020, p. 152.
  127. ^ Kristensen, Beaudoin & Ives 2020, p. 21.
  128. ^ a b c Sheets & Grayson 1979, p. 350.
  129. ^ Ives 2019, p. 43.
  130. ^ Kristensen, Beaudoin & Ives 2020, p. 17.
  131. ^ Moodie, Catchpole & Abel 1992, p. 149.
  132. ^ Matson & Magne 2003, p. 556.
  133. ^ a b Bunbury & Gajewski 2013, p. 18.
  134. ^ a b Reuther et al. 2020, p. 171.
  135. ^ Pewe 1975, p. 19.
  136. ^ Wilbur et al. 1991.
  137. ^ Holloway 2020, p. 29.
  138. ^ Mullen 2012, p. 37.
  139. ^ Reuther et al. 2020, p. 185.
  140. ^ Harrod 2017, p. 149.
  141. ^ Kristensen, Beaudoin & Ives 2020, p. 2.
  142. ^ a b c Mullen 2012, p. 36.
  143. ^ Jensen et al. 2014, p. 875.
  144. ^ Pyne-O'Donnell et al. 2012, p. 9.
  145. ^ a b Patterson et al. 2017, p. 2.
  146. ^ Paine et al. 2021, p. 6.
  147. ^ Paine et al. 2021, p. 1.
  148. ^ Watson et al. 2017, p. 460.
  149. ^ Jensen et al. 2014, p. 876.
  150. ^ Zdanowicz et al. 2014, p. 48.
  151. ^ Addison et al. 2010, p. 278.
  152. ^ a b Jensen et al. 2014, p. 877.
  153. ^ Bourne et al. 2016, p. 5.
  154. ^ Plunkett & Pilcher 2018, p. 20.
  155. ^ Stevenson et al. 2015, p. 2075.
  156. ^ Kristensen et al. 2019, p. 13.
  157. ^ Kristensen, Ives & Supernant 2021, p. 441.
  158. ^ Bunbury & Gajewski 2013, p. 29.
  159. ^ Kristensen, Ives & Supernant 2021, p. 442.
  160. ^ Kristensen, Ives & Supernant 2021, p. 443.
  161. ^ Doering 2021, p. 3.
  162. ^ Kristensen, Beaudoin & Ives 2020, pp. 18–19.
  163. ^ a b Kristensen, Ives & Supernant 2021, p. 444.
  164. ^ Kristensen, Ives & Supernant 2021, p. 445.
  165. ^ Kristensen et al. 2019, p. 15.
  166. ^ Moodie, Catchpole & Abel 1992, p. 157.
  167. ^ Kristensen, Ives & Supernant 2021, p. 449.
  168. ^ Kristensen, Ives & Supernant 2021, p. 448.
  169. ^ Kristensen, Ives & Supernant 2021, p. 451.
  170. ^ Wilmeth 1979, p. 35.
  171. ^ Kristensen, Ives & Supernant 2021, p. 453.
  172. ^ Kristensen et al. 2019, p. 2.
  173. ^ a b Nash & Baxter 2023, p. 302.
  174. ^ Szykulski 2021, p. 283.
  175. ^ Coates 2020, p. 42.
  176. ^ Addison et al. 2010, p. 285.
  177. ^ Moodie, Catchpole & Abel 1992, p. 163.
  178. ^ Mackay et al. 2022, pp. 1477–1478.
  179. ^ Mackay et al. 2022, p. 1482.
  180. ^ Mackay et al. 2022, p. 1489.
  181. ^ Mackay et al. 2022, p. 1490.
  182. ^ Mackay et al. 2022, p. 1485.
  183. ^ Mackay et al. 2022, p. 1486.
  184. ^ Plunkett et al. 2023, p. 12.
  185. ^ Newfield 2018, p. 464.
  186. ^ a b Stasiuk, Hickson & Mulder 2003, p. 576.
  187. ^ Stasiuk, Hickson & Mulder 2003, p. 569.
  188. ^ Clague 1982, p. 35.
  189. ^ Ewert 2007, p. 122.
  190. ^ Ewert 2007, p. 112.
  191. ^ Ewert, Diefenbach & Ramsey 2018, p. 8.
  192. ^ Bourne et al. 2016, p. 2.

Sources edit

  • Addison, Jason A.; Beget, James E.; Ager, Thomas A.; Finney, Bruce P. (March 2010). "Marine tephrochronology of the Mt. Edgecumbe Volcanic Field, Southeast Alaska, USA". Quaternary Research. 73 (2): 277–292. Bibcode:2010QuRes..73..277A. doi:10.1016/j.yqres.2009.10.007. S2CID 59584705.
  • Bailey, Robert C.; Burn, Christopher R. (2023), "Yukon River Basin", Rivers of North America, Elsevier, pp. 714–745, doi:10.1016/b978-0-12-818847-7.00023-9, ISBN 978-0-12-818847-7, retrieved 2024-01-09
  • Beierle, Brandon D. (2002). Stratigraphic displacement of a tephra bed in organic lake sediments (Report). Geological Society of America Abstracts with Programs. Vol. 34.
  • Birks, H. J. B. (15 September 1977). "Modern pollen rain and vegetation of the St. Elias Mountains, Yukon Territory". Canadian Journal of Botany. 55 (18): 2367–2382. doi:10.1139/b77-270. ISSN 0008-4026.
  • Bond, Jeffrey David; MacGregor, Dylan B.; Lipovsky, Panya S. (2001). Quaternary geology and till geochemistry of the Anvil district (parts of 105K/2, 3, 5, 6 and 7), central Yukon Territory (PDF) (Report). Indian & Northern Affairs Canada, Yukon Region, Exploration & Geological Services Division. Retrieved 9 January 2024.
  • Bunbury, J.; Gajewski, K.; Weston, L. H.; Blackburn, L. R.; Lewis, L. L. (2008). "Variations in the depth and thickness of the White River Ash in lakes of the southwest Yukon" (PDF). Yukon Exploration and Geology: 77–84.
  • Bourne, A. J.; Abbott, P. M.; Albert, P. G.; Cook, E.; Pearce, N. J. G.; Ponomareva, V.; Svensson, A.; Davies, S. M. (21 July 2016). "Underestimated risks of recurrent long-range ash dispersal from northern Pacific Arc volcanoes". Scientific Reports. 6 (1): 29837. Bibcode:2016NatSR...629837B. doi:10.1038/srep29837. PMC 4956762. PMID 27445233.
  • Bowen, Mark (3 October 2006). Thin Ice: Unlocking the Secrets of Climate in the World's Highest Mountains. Henry Holt and Company. ISBN 978-1-4299-3270-7.
  • Brueseke, Matthew E.; Benowitz, Jeffrey A.; Trop, Jeffrey M.; Davis, Kailyn N.; Berkelhammer, Samuel E.; Layer, Paul W.; Morter, Bethany K. (February 2019). "The Alaska Wrangell Arc: ~30 Ma of subduction-related magmatism along a still active arc-transform junction". Terra Nova. 31 (1): 59–66. Bibcode:2019TeNov..31...59B. doi:10.1111/ter.12369. S2CID 133881900.
  • Bunbury, Joan; Gajewski, Konrad (2013). "Effects of the White River Ash Event on Aquatic Environments, Southwest Yukon, Canada". Arctic. 66 (1): 17–31. doi:10.14430/arctic4262. ISSN 0004-0843. JSTOR 23594603.
  • Campbell, Seth; Kreutz, Karl; Osterberg, Erich; Arcone, Steven; Wake, Cameron; Introne, Douglas; Volkening, Kevin; Winski, Dominic (2012). "Melt regimes, stratigraphy, flow dynamics and glaciochemistry of three glaciers in the Alaska Range". Journal of Glaciology. 58 (207): 99–109. Bibcode:2012JGlac..58...99C. doi:10.3189/2012JoG10J238.
  • Capps, S.R. (1916). An ancient volcanic eruption in the upper Yukon Basin (Report). Professional Paper 95-D. Washington, D.C.: United States Government Printing Office. pp. 59–72. doi:10.3133/pp95D.
  • Clague, John J. (1982), "The Role of Geomorphology in the Identification and Evaluation of Natural Hazards", in Craig, Richard G.; Craft, Jesse L. (eds.), Applied Geomorphology, Routledge, doi:10.4324/9781003027461, ISBN 978-1-003-02746-1, S2CID 214279352
  • Coates, Karen (2020). "Walking into New Worlds". Archaeology. 73 (5): 38–43. ISSN 0003-8113. JSTOR 27056755.
  • Davies, Lauren J.; Jensen, Britta J.L.; Froese, Duane G.; Wallace, Kristi L. (August 2016). "Late Pleistocene and Holocene tephrostratigraphy of interior Alaska and Yukon: Key beds and chronologies over the past 30,000 years". Quaternary Science Reviews. 146: 28–53. Bibcode:2016QSRv..146...28D. doi:10.1016/j.quascirev.2016.05.026.
  • Davies, Lauren J.; Jensen, Britta J. L.; Kaufman, Darrell S. (11 March 2022). "Late Holocene cryptotephra and a provisional 15 000-year Bayesian age model for Cascade Lake, Alaska". Geochronology. 4 (1): 121–141. Bibcode:2022GeChr...4..121D. doi:10.5194/gchron-4-121-2022. ISSN 2628-3697.
  • Doering, Briana N. (September 2021). "Subarctic landscape adaptations and paleodemography: A 14,000-year history of climate change and human settlement in central Alaska and Yukon". Quaternary Science Reviews. 268: 107139. Bibcode:2021QSRv..26807139D. doi:10.1016/j.quascirev.2021.107139.
  • Donaldson, J.A.; Mueller, W. (1994). Dispersal of pumice along Klutlan Glacier, Yukon Territory (Report). Geological Society of America, Abstracts with Programs. Vol. 26. p. 113.
  • Donaldson, J. Allan; Guerstein, Pablo G.; Mueller, Wulf (1 September 1996). "Facies analysis of a pumiceous terrace beside klutlan Glacier, Yukon Territory". Canadian Journal of Earth Sciences. 33 (9): 1233–1242. Bibcode:1996CaJES..33.1233D. doi:10.1139/e96-093. ISSN 0008-4077.
  • Ewert, John W. (November 2007). "System for Ranking Relative Threats of U.S. Volcanoes". Natural Hazards Review. 8 (4): 112–124. doi:10.1061/(ASCE)1527-6988(2007)8:4(112). ISSN 1527-6988.
  • Ewert, J.W.; Diefenbach, A.K.; Ramsey, D.W. (2018). "2018 update to the U.S. Geological Survey national volcanic threat assessment". U.S. Geological Survey Scientific Investigations Report 2018–5140 (Report). Scientific Investigations Report. p. 40. doi:10.3133/sir20185140. ISSN 2328-0328.
  • Fang, Ling; Jenk, Theo M.; Winski, Dominic; Kreutz, Karl; Brooks, Hanna L.; Erwin, Emma; Osterberg, Erich; Campbell, Seth; Wake, Cameron; Schwikowski, Margit (15 September 2023). "Early Holocene ice on the Begguya plateau (Mt. Hunter, Alaska) revealed by ice core 14C age constraints". The Cryosphere. 17 (9): 4007–4020. doi:10.5194/tc-17-4007-2023. ISSN 1994-0416.
  • "Churchill". Global Volcanism Program. Smithsonian Institution. Retrieved 7 January 2024.
  • Harrod, Ryan P. (2017), "The Decline of Social Control in the Pueblo World", The Bioarchaeology of Social Control, Cham: Springer International Publishing, pp. 145–161, doi:10.1007/978-3-319-59516-0_9, ISBN 978-3-319-59515-3, retrieved 2024-01-08
  • Hinkley, Todd K. (August 1994). "Composition and sources of atmospheric dusts in snow at 3200 meters in the St. Elias Range, southeastern Alaska, USA". Geochimica et Cosmochimica Acta. 58 (15): 3245–3254. Bibcode:1994GeCoA..58.3245H. doi:10.1016/0016-7037(94)90052-3.
  • Holdsworth, Gerald; Roy Krouse, H. (2002). "Altitudinal variation of the stable isotopes of snow in regions of high relief". Journal of Glaciology. 48 (160): 31–41. Bibcode:2002JGlac..48...31H. doi:10.3189/172756502781831638.
  • Holloway, Caitlin R. (2020). "MIDDLE-LATE HOLOCENE ARCHAEOLOGY OF THE UPPER DIAMOND FORK VALLEY, YUKON-CHARLEY RIVERS NATIONAL PRESERVE" (PDF). Alaska Journal of Anthropology. 18 (2).
  • Ives, John W. (20 May 2019). A Theory of Northern Athapaskan Prehistory (1 ed.). Routledge. doi:10.4324/9780429044212. ISBN 978-0-429-04421-2. S2CID 197836650 – via ResearchGate.
  • Jensen, B. J.; Froese, D. G.; Preece, S. J.; Westgate, J. A. (December 2006). An Extensive Middle to Late Pleistocene Distal Tephra Record From East Central Alaska. American Geophysical Union, Fall Meeting 2006. Bibcode:2006AGUFM.V33B0662J. V33B-0662.
  • Jensen, Britta J.L.; Pyne-O’Donnell, Sean; Plunkett, Gill; Froese, Duane G.; Hughes, Paul D.M.; Sigl, Michael; McConnell, Joseph R.; Amesbury, Matthew J.; Blackwell, Paul G.; van den Bogaard, Christel; Buck, Caitlin E.; Charman, Dan J.; Clague, John J.; Hall, Valerie A.; Koch, Johannes; Mackay, Helen; Mallon, Gunnar; McColl, Lynsey; Pilcher, Jonathan R. (October 2014). "Transatlantic distribution of the Alaskan White River Ash". Geology. 42 (10): 875–878. Bibcode:2014Geo....42..875J. doi:10.1130/G35945.1.
  • Kristensen, Todd J.; Gregory Hare, P.; Gotthardt, Ruth M.; Easton, Norman A.; Ives, John W.; Speakman, Robert J.; Rasic, Jeffrey T. (December 2019). "The movement of obsidian in Subarctic Canada: Holocene social relationships and human responses to a large-scale volcanic eruption". Journal of Anthropological Archaeology. 56: 101114. doi:10.1016/j.jaa.2019.101114.
  • Kristensen, Todd J.; Beaudoin, Alwynne B.; Ives, John W. (2020). "Environmental and Hunter-Gatherer Responses to the White River Ash East Volcanic Eruption in the Late Holocene Canadian Subarctic". Arctic. 73 (2): 153–186. ISSN 0004-0843. JSTOR 26974890.
  • Kristensen, Todd J; Ives, John W; Supernant, Kisha (October 2021). "Power, security, and exchange: Impacts of a Late Holocene volcanic eruption in Subarctic North America". North American Archaeologist. 42 (4): 425–472. doi:10.1177/0197693120986822. S2CID 234270362.
  • Lerbekmo, J. F.; Campbell, F. A. (1 February 1969). "Distribution, composition, and source of the White River Ash, Yukon Territory". Canadian Journal of Earth Sciences. 6 (1): 109–116. Bibcode:1969CaJES...6..109L. doi:10.1139/e69-011. ISSN 0008-4077.
  • Lerbekmo, J. F. (June 2008). "The White River Ash: largest Holocene Plinian tephra". Canadian Journal of Earth Sciences. 45 (6): 693–700. Bibcode:2008CaJES..45..693L. doi:10.1139/E08-023. ISSN 0008-4077.
  • Liang, Xuran; Zhao, Dapeng; Hua, Yuanyuan; Xu, Yi-Gang (January 2024). "Big Mantle Wedge and Intraplate Volcanism in Alaska: Insight From Anisotropic Tomography". Journal of Geophysical Research: Solid Earth. 129 (1). Bibcode:2024JGRB..12927617L. doi:10.1029/2023JB027617.
  • Lollino, Giorgio; Manconi, Andrea; Clague, John; Shan, Wei; Chiarle, Marta, eds. (2015). Engineering Geology for Society and Territory - Volume 1: Climate Change and Engineering Geology. Cham: Springer International Publishing. doi:10.1007/978-3-319-09300-0. ISBN 978-3-319-09299-7. S2CID 53608652.
  • Lowe, David J. (April 2011). "Tephrochronology and its application: A review". Quaternary Geochronology. 6 (2): 107–153. Bibcode:2011QuGeo...6..107L. doi:10.1016/j.quageo.2010.08.003. hdl:10289/4616.
  • Mackay, Helen; Plunkett, Gill; Jensen, Britta J. L.; Aubry, Thomas J.; Corona, Christophe; Kim, Woon Mi; Toohey, Matthew; Sigl, Michael; Stoffel, Markus; Anchukaitis, Kevin J.; Raible, Christoph; Bolton, Matthew S. M.; Manning, Joseph G.; Newfield, Timothy P.; Di Cosmo, Nicola; Ludlow, Francis; Kostick, Conor; Yang, Zhen; Coyle McClung, Lisa; Amesbury, Matthew; Monteath, Alistair; Hughes, Paul D. M.; Langdon, Pete G.; Charman, Dan; Booth, Robert; Davies, Kimberley L.; Blundell, Antony; Swindles, Graeme T. (29 June 2022). "The 852/3 CE Mount Churchill eruption: examining the potential climatic and societal impacts and the timing of the Medieval Climate Anomaly in the North Atlantic region". Climate of the Past. 18 (6): 1475–1508. Bibcode:2022CliPa..18.1475M. doi:10.5194/cp-18-1475-2022. ISSN 1814-9324.
  • "Churchill". USGS National Map. 31 December 1981. Retrieved 7 January 2024.
  • Mantell, C. L.; Mantell, A. M. (1976). Our Fragile Water Planet. Boston, MA: Springer US. doi:10.1007/978-1-4684-0754-9. ISBN 978-1-4684-0756-3.
  • Matson, R. G.; Magne, Martin Paul Robert (31 May 2003). Athapaskans and Migrations: The Archaeology of Eagle Lake, British Columbia. doi:10.14288/1.0363461.
  • McClung, Lisa Coyle; Plunkett, Gill (2020). "Cultural change and the climate record in final prehistoric and early medieval Ireland". Proceedings of the Royal Irish Academy: Archaeology, Culture, History, Literature. 120C (1): 129–158. doi:10.1353/ria.2020.0014.
  • McGimsey, Robert G.; Richter, Donald H.; DuBois, Gregory D.; Miller, T. P. (1992). A postulated new source for the White River Ash, Alaska: A section in Geologic studies in Alaska by the US. Geological Survey, 1990 (Report). Bulletin 1999. Denver, CO: U.S. Geological Survey. pp. 212–218. doi:10.3133/70180193.
  • McGimsey, R.G.; Richter, D.H.; Waythomas, C.F.; Donaldson, J.A. (1995). Potential hazards from future eruptions of Mt. Churchill, Alaska (Report). Geological Society of America, Abstracts with Programs, Cordilleran Section. Vol. 27. p. 63.
  • Miller, Thomas P.; Richter, Donald H. (1994), Plafker, George; Berg, Henry C. (eds.), "Quaternary volcanism in the Alaska Peninsula and Wrangell Mountains, Alaska", The Geology of Alaska, Boulder, Colorado: Geological Society of America, pp. 759–779, doi:10.1130/dnag-gna-g1.759, ISBN 978-0-8137-5219-8, retrieved 2024-01-08
  • Moffit, F.H. (1938). Geology of the Chitina Valley and adjacent area, Alaska (PDF) (Report). U.S. Geological Survey Bulletin 894. p. 137.
  • Monteath, Alistair J.; van Hardenbroek, Maarten; Davies, Lauren J.; Froese, Duane G.; Langdon, Peter G.; Xu, Xiaomei; Edwards, Mary E. (September 2017). "Chronology and glass chemistry of tephra and cryptotephra horizons from lake sediments in northern Alaska, USA" (PDF). Quaternary Research. 88 (2): 169–178. Bibcode:2017QuRes..88..169M. doi:10.1017/qua.2017.38. S2CID 55217343.
  • Moodie, D. Wayne; Catchpole, A. J. W.; Abel, Kerry (1992). "Northern Athapaskan Oral Traditions and the White River Volcano". Ethnohistory. 39 (2): 148–171. doi:10.2307/482391. ISSN 0014-1801. JSTOR 482391.
  • Mullen, Patrick O. (6 February 2012). "An Archaeological Test of the Effects of the White River Ash Eruptions". Arctic Anthropology. 49 (1): 35–44. doi:10.1353/arc.2012.0013. ISSN 0066-6939. S2CID 129873168 – via Project MUSE.
  • Nash, Stephen E.; Baxter, Erin L., eds. (2023). Pushing Boundaries in Southwestern Archaeology: Chronometry, Collections, and Contexts. University Press of Colorado. ISBN 978-1-64642-362-0.
  • "Mount Churchill". Natural Resources Canada. Archived from the original on 8 June 2009.
  • Newfield, Timothy P. (2018), White, Sam; Pfister, Christian; Mauelshagen, Franz (eds.), "The Climate Downturn of 536–50", The Palgrave Handbook of Climate History, London: Palgrave Macmillan UK, pp. 447–493, doi:10.1057/978-1-137-43020-5_32, ISBN 978-1-137-43019-9, retrieved 2024-01-08
  • Paine, Alice R.; Baldini, James U.L.; Wadsworth, Fabian B.; Lechleitner, Franziska A.; Jamieson, Robert A.; Baldini, Lisa M.; Brown, Richard J.; Müller, Wolfgang; Hercman, Helena; Gąsiorowski, Michał; Stefaniak, Krzysztof; Socha, Paweł; Sobczyk, Artur; Kasprzak, Marek (June 2021). "The trace-element composition of a Polish stalagmite: Implications for the use of speleothems as a record of explosive volcanism". Chemical Geology. 570: 120157. Bibcode:2021ChGeo.57020157P. doi:10.1016/j.chemgeo.2021.120157. S2CID 233542503.
  • Patterson, R. Timothy; Crann, Carley A.; Cutts, Jamie A.; Courtney Mustaphi, Colin J.; Nasser, Nawaf A.; Macumber, Andrew L.; Galloway, Jennifer M.; Swindles, Graeme T.; Falck, Hendrik (July 2017). "New occurrences of the White River Ash (east lobe) in Subarctic Canada and utility for estimating freshwater reservoir effect in lake sediment archives" (PDF). Palaeogeography, Palaeoclimatology, Palaeoecology. 477: 1–9. Bibcode:2017PPP...477....1P. doi:10.1016/j.palaeo.2017.03.031.
  • Payne, Richard; Blackford, Jeffrey; van der Plicht, Johannes (January 2008). "Using cryptotephras to extend regional tephrochronologies: An example from southeast Alaska and implications for hazard assessment" (PDF). Quaternary Research. 69 (1): 42–55. Bibcode:2008QuRes..69...42P. doi:10.1016/j.yqres.2007.10.007. S2CID 73704290.
  • Pewe, Troy Lewis (1975). Quaternary stratigraphic nomenclature in unglaciated central Alaska (Report). Professional Paper 862. U.S. Govt. Print. Off. doi:10.3133/pp862.
  • Plunkett, Gill; Pilcher, Jonathan R. (April 2018). "Defining the potential source region of volcanic ash in northwest Europe during the Mid- to Late Holocene" (PDF). Earth-Science Reviews. 179: 20–37. Bibcode:2018ESRv..179...20P. doi:10.1016/j.earscirev.2018.02.006. S2CID 85561033.
  • Porter, S. E.; Mosley-Thompson, E.; Thompson, L. G.; Bolzan, J. F. (December 2017). A Paleo Perspective on Arctic and Mid-latitude Linkages from a Southeast Alaska Ice Core. American Geophysical Union, Fall Meeting 2017. Bibcode:2017AGUFM.A42B..02P. #A42B-02.
  • Plunkett, Gill; Sigl, Michael; McConnell, Joseph R.; Pilcher, Jonathan R.; Chellman, Nathan J. (February 2023). "The significance of volcanic ash in Greenland ice cores during the Common Era". Quaternary Science Reviews. 301: 107936. Bibcode:2023QSRv..30107936P. doi:10.1016/j.quascirev.2022.107936.
  • Porter, Stacy E.; Mosley-Thompson, Ellen; Thompson, Lonnie G. (27 October 2019). "Ice Core δ 18 O Record Linked to Western Arctic Sea Ice Variability". Journal of Geophysical Research: Atmospheres. 124 (20): 10784–10801. doi:10.1029/2019JD031023.
  • Preece, Shari J.; Hart, William K. (November 2004). "Geochemical variations in the". Tectonophysics. 392 (1–4): 165–191. doi:10.1016/j.tecto.2004.04.011.
  • Preece, S.J.; McGimsey, R.G.; Westgate, J.A.; Pearce, N.J.G.; Hart, W.K.; Perkins, W.T. (October 2014). "Chemical complexity and source of the White River Ash, Alaska and Yukon". Geosphere. 10 (5): 1020–1042. Bibcode:2014Geosp..10.1020P. doi:10.1130/GES00953.1.
  • Pyne-O'Donnell, Sean D.F.; Hughes, Paul D.M.; Froese, Duane G.; Jensen, Britta J.L.; Kuehn, Stephen C.; Mallon, Gunnar; Amesbury, Matthew J.; Charman, Dan J.; Daley, Tim J.; Loader, Neil J.; Mauquoy, Dmitri; Street-Perrott, F. Alayne; Woodman-Ralph, Jonathan (October 2012). "High-precision ultra-distal Holocene tephrochronology in North America". Quaternary Science Reviews. 52: 6–11. Bibcode:2012QSRv...52....6P. doi:10.1016/j.quascirev.2012.07.024.
  • Rampton, Vern (1 October 1970). "Neoglacial fluctuations of the Natazhat and Klutlan Glaciers, Yukon Territory, Canada". Canadian Journal of Earth Sciences. 7 (5): 1236–1263. Bibcode:1970CaJES...7.1236R. doi:10.1139/e70-118. ISSN 0008-4077.
  • Rampton, V.N. (July 1978). "Holocene Glacial and Tree-Line Variations in the White River Valley and Skolai Pass, Alaska and Yukon Territory: A discussion". Quaternary Research. 10 (1): 130–134. Bibcode:1978QuRes..10..130R. doi:10.1016/0033-5894(78)90017-0. S2CID 129381047.
  • Rampton, Vern (1971). "Late Quaternary Vegetational and Climatic History of the Snag-Klutlan Area, Southwestern Yukon Territory, Canada". Geological Society of America Bulletin. 82 (4): 959. doi:10.1130/0016-7606(1971)82[959:LQVACH]2.0.CO;2.
  • Reuther, Joshua; Potter, Ben; Coffman, Sam; Smith, Holly; Bigelow, Nancy (February 2020). "Revisiting the Timing of the Northern Lobe of the White River Ash Volcanic Event in Eastern Alaska and Western Yukon". Radiocarbon. 62 (1): 169–188. Bibcode:2020Radcb..62..169R. doi:10.1017/RDC.2019.110. S2CID 203074680.
  • Richter, D H; Smith, J G; Lanphere, M A; Dalrymple, G B; Reed, B L; Shew, Nora (December 1990). "Age and progression of volcanism, Wrangell volcanic field, Alaska". Bulletin of Volcanology. 53 (1): 29–44. Bibcode:1990BVol...53...29R. doi:10.1007/BF00680318. S2CID 140169594.
  • Richter, D. H.; Preece, S. J.; Mcgimsey, R. G.; Westgate, J. A. (1 June 1995). "Mount Churchill, Alaska: source of the late Holocene White River Ash". Canadian Journal of Earth Sciences. 32 (6): 741–748. Bibcode:1995CaJES..32..741R. doi:10.1139/e95-063. ISSN 0008-4077.
  • Richter, D. H.; Rosenkrans, D. S.; Steigerwald, M. J. (1995). Guide to the Volcanoes of the Western Wrangell Mountains, Alaska: Wrangell-St. Elias National Park and Preserve (No. 2072) (PDF) (Report). US Government Printing Office.
  • Ringsmuth, Katherine Johnson (2015). Alaska's Skyboys: Cowboy Pilots and the Myth of the Last Frontier. University of Washington Press. ISBN 9780295995083.
  • Robinson, Stephen D. (2001). "Extending the Late Holocene White River Ash Distribution, Northwestern Canada". Arctic. 54 (2): 157–161. doi:10.14430/arctic775. ISSN 0004-0843. JSTOR 40512371.
  • Scott, William E. (2003). "Quaternary volcanism in the United States". Developments in Quaternary Sciences. 1: 351–380. doi:10.1016/S1571-0866(03)01016-9. ISBN 978-0-444-51470-7.
  • Sheets, Payson D.; Grayson, Donald K. (1979). Volcanic activity and human ecology. New York: Academic Press. ISBN 978-0-12-639120-6.
  • Sierra-Hernández, M. Roxana; Beaudon, Emilie; Porter, Stacy E.; Mosley-Thompson, Ellen; Thompson, Lonnie G. (27 January 2022). "Increased Fire Activity in Alaska Since the 1980s: Evidence From an Ice Core-Derived Black Carbon Record". Journal of Geophysical Research: Atmospheres. 127 (2). Bibcode:2022JGRD..12735668S. doi:10.1029/2021JD035668. S2CID 245845893.
  • Smith, C. A. S.; Ping, C. L.; Fox, C. A.; Kodama, H. (1 November 1999). "Weathering characteristics of some soils formed in White River Tephra, Yukon Territory, Canada". Canadian Journal of Soil Science. 79 (4): 603–613. doi:10.4141/S98-066.
  • Stasiuk, Mark V.; Hickson, Catherine J.; Mulder, Taimi (2003). "The Vulnerability of Canada to Volcanic Hazards". Natural Hazards. 28 (2/3): 563–589. doi:10.1023/A:1022954829974. S2CID 129461798.
  • 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): 2069–2091. Bibcode:2015AMT.....8.2069S. doi:10.5194/amt-8-2069-2015. ISSN 1867-1381.
  • Strickland, A. J.; Arocena, J. M.; Sanborn, P.; Smith, C. A. S. (1 November 2005). "Secondary mineral formation in the White River tephra in grassland and forest soils in central Yukon Territory". Canadian Journal of Soil Science. 85 (5): 637–648. doi:10.4141/S05-001. ISSN 0008-4271.
  • Szykulski, Józef (3 September 2021). "Radosław Palonka, Sztuka i archeologia kultur indiańskich prekolumbijskiego Południowego Zachodu USA. Wydawnictwo Uniwersytetu Jagiellońskiego, Kraków 2019, ss. 530, ryc. 307, tabl. 15". Przegląd Archeologiczny (in Polish). 69: 279–283. doi:10.23858/PA69.2021.2402. ISSN 2657-4004.
  • Thompson, L. G.; Mosley-Thompson, E. S.; Zagorodnov, V.; Davis, M. E.; Mashiotta, T. A.; Lin, P. (December 2004). 1500 Years of Annual Climate and Environmental Variability as Recorded in Bona-Churchill (Alaska) Ice Cores. American Geophysical Union, Fall Meeting 2004. Bibcode:2004AGUFMPP23C..05T. PP23C-05.
  • Trop, Jeffrey M.; Benowitz, Jeff A.; Kirby, Carl S.; Brueseke, Matthew E. (1 February 2022). "Geochronology of the Wrangell Arc: Spatial-temporal evolution of slab-edge magmatism along a flat-slab, subduction-transform transition, Alaska-Yukon". Geosphere. 18 (1): 19–48. Bibcode:2022Geosp..18...19T. doi:10.1130/GES02417.1.
  • Trop, Jeffrey (2012). "Miocene basin development and volcanism along a strike-slip to flat-slab subduction transition: Stratigraphy, geochemistry, and geochronology of the central Wrangell volcanic belt, Yakutat–North America collision zone". Geosphere. 8 (4): 805. doi:10.1130/GES00762.1.
  • Urmann, D. (2009). Decadal Scale Climate Variability During The Last Millennium As Recorded By The Bona Churchill And Quelccaya Ice Cores. OhioLINK Electronic Theses and Dissertations Center (Doctoral dissertation). Ohio State University.
  • Watson, E. J.; Kołaczek, P.; Słowiński, M.; Swindles, G. T.; Marcisz, K.; Gałka, M.; Lamentowicz, M. (May 2017). "First discovery of Holocene Alaskan and Icelandic tephra in Polish peatlands: Holocene Alaskan and Icelandic tephra in Polish peatlands". Journal of Quaternary Science. 32 (4): 457–462. doi:10.1002/jqs.2945.
  • West, K. D.; Donaldson, J. A. (2001). "Resedimentation of the late Holocene White River tephra, Yukon Territory and Alaska" (PDF). Yukon Exploration and Geology: 239–247.
  • Westgate, J.A.; Preece, S.J.; Froese, D.G.; Pearce, N.J.G.; Roberts, R.G.; Demuro, M.; Hart, W.K.; Perkins, W. (February 2008). "Changing ideas on the identity and stratigraphic significance of the Sheep Creek tephra beds in Alaska and the Yukon Territory, northwestern North America". Quaternary International. 178 (1): 183–209. Bibcode:2008QuInt.178..183W. doi:10.1016/j.quaint.2007.03.009.
  • Wilbur, S.C.; Molinari, M.P.; Beget, J.E.; Hengesh, J.V. (1991). Four Holocene tephra from the Prince William Sound area, Alaska (Report). Geological Society of America, Abstracts with Programs. Vol. 23. p. 398.
  • Wiles, Gregory C.; Jacoby, Gordon C.; Davi, Nicole K.; McAllister, Ryan P. (July 2002). "Late Holocene glacier fluctuations in the Wrangell Mountains, Alaska". Geological Society of America Bulletin. 114 (7): 896–908. Bibcode:2002GSAB..114..896W. doi:10.1130/0016-7606(2002)114<0896:LHGFIT>2.0.CO;2.
  • Wilmeth, Roscoe (1979). "An Athabaskan Hypothesis". Canadian Journal of Archaeology / Journal Canadien d'Archéologie (3): 33–40. ISSN 0705-2006. JSTOR 41102195.
  • Wilson, Frederic H.; Labay, Keith A.; Shew, Nora B.; Preller, Cindi C.; Mohadjer, Solmaz; Richter, Donald H. (2005). Preliminary integrated geologic map databases for the United States: digital data for the geology of Wrangell-Saint Elias National Park and Preserve, Alaska (Report). U.S. Geological Survey Open-File Report 2005-1342.
  • Winkler, Gary R. A Geologic Guide to Wrangell—Saint Elias National Park and Preserve, Alaska: A Tectonic Collage of Northbound Terranes (PDF). USGS (Report). Retrieved 7 January 2024.
  • Wood, Michael; Coombs, Colby (January 2001). Alaska: A Climbing Guide. The Mountaineers Books. ISBN 978-0-89886-724-4.
  • Zander, Paul D.; Kaufman, Darrell S.; Kuehn, Stephen C.; Wallace, Kristi L.; Anderson, R. Scott (November 2013). "Early and late Holocene glacial fluctuations and tephrostratigraphy, Cabin Lake, Alaska". Journal of Quaternary Science. 28 (8): 761–771. Bibcode:2013JQS....28..761Z. doi:10.1002/jqs.2671. S2CID 131689652.
  • Zdanowicz, Christian; Fisher, David; Bourgeois, Jocelyne; Demuth, Mike; Zheng, James; Mayewski, Paul; Kreutz, Karl; Osterberg, Erich; Yalcin, Kaplan; Wake, Cameron; Steig, Eric J.; Froese, Duane; Goto-Azuma, Kumiko (2014). "Ice Cores from the St. Elias Mountains, Yukon, Canada: Their Significance for Climate, Atmospheric Composition and Volcanism in the North Pacific Region". Arctic. 67 (5): 35–57. doi:10.14430/arctic4352. ISSN 0004-0843. JSTOR 24363799.

External links edit

  • "Mount Churchill". Bivouac.com. Retrieved 2014-01-01.