Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-24T09:09:31.042Z Has data issue: false hasContentIssue false

Chronology and glass chemistry of tephra and cryptotephra horizons from lake sediments in northern Alaska, USA

Published online by Cambridge University Press:  31 July 2017

Alistair J. Monteath*
Affiliation:
Palaeoecology Laboratory (PLUS), School of Geography, University of Southampton, Southampton SO17 1BJ, United Kingdom
Maarten van Hardenbroek
Affiliation:
School of Geography, Politics and Sociology, University of Newcastle, Newcastle upon Tyne NE1 7RU, United Kingdom
Lauren J. Davies
Affiliation:
Department of Earth and Atmospheric Sciences, University of Alberta, Edmonton, Alberta T6G 2E3, Canada
Duane G. Froese
Affiliation:
Department of Earth and Atmospheric Sciences, University of Alberta, Edmonton, Alberta T6G 2E3, Canada
Peter G. Langdon
Affiliation:
Palaeoecology Laboratory (PLUS), School of Geography, University of Southampton, Southampton SO17 1BJ, United Kingdom
Xiaomei Xu
Affiliation:
Keck Carbon Cycle AMS Facility, University of California, Irvine, Irvine, California 92697-3100, USA
Mary E. Edwards
Affiliation:
Palaeoecology Laboratory (PLUS), School of Geography, University of Southampton, Southampton SO17 1BJ, United Kingdom Alaska Quaternary Center, College of Natural Sciences and Mathematics, University of Alaska Fairbanks, Fairbanks, Alaska 99775, USA
*
*Corresponding author at: Geography and Environment, Room 1067, Shackleton Building 44, University of Southampton, Southampton SO17 1BJ, United Kingdom. E-mail address: ali.monteath@soton.ac.uk (A.J. Monteath).

Abstract

Holocene tephrostratigraphy in Alaska provides independent chronology and stratigraphic correlation in a region where reworked old (Holocene) organic carbon can significantly distort radiocarbon chronologies. Here, we present new glass chemistry and chronology for Holocene tephras preserved in three Alaskan lakes: one in the eastern interior and two in the southern Brooks Range. Tephra beds in the eastern interior lake-sediment core are correlated with the White River Ash and the Hayes tephra set H (~4200–3700 cal yr BP), and an additional discrete tephra bed is likely from the Aleutian arc/Alaska Peninsula. Cryptotephras (nonvisible tephras) found in the Brooks Range include the informally named “Ruppert tephra” (~2700–2300 cal yr BP) and the Aniakchak caldera-forming event II (CFE II) tephra (~3600 cal yr BP). A third underlying Brooks Range cryptotephra is chemically indistinguishable from the Aniakchak CFE II tephra (4070–3760 cal yr BP) and is likely to be from an earlier eruption of the Aniakchak volcano.

Type
Research Article
Copyright
Copyright © University of Washington. Published by Cambridge University Press, 2017 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

REFERENCES

Abbott, M.B., Stafford, T.W., 1996. Radiocarbon geochemistry of modern and ancient arctic lake systems, Baffin Island, Canada. Quaternary Research 45, 300311.Google Scholar
Anderson, R., Nuhfer, E., Dean, W.E., 1984. Sinking of volcanic ash in uncompacted sediment in Williams Lake, Washington. Science 225, 505508.CrossRefGoogle ScholarPubMed
Begét, J., Mason, O., Anderson, P., 1992. Age, extent and climatic significance of the c. 3400 BP Aniakchak tephra, western Alaska, USA. Holocene 2, 5156.CrossRefGoogle Scholar
Beierle, B., Bond, J., 2002. Density-induced settling of tephra through organic lake sediments. Journal of Paleolimnology 28, 433440.Google Scholar
Blaauw, M., Christen, J.A., 2011. Flexible paleoclimate age-depth models using an autoregressive gamma process. Bayesian Analysis 6, 457474.Google Scholar
Blockley, S.P.E., Pyne-O’Donnell, S.D.F., Lowe, J.J., Matthews, I.P., Stone, A., Pollard, A.M., Turney, C.S.M., Molyneux, E.G., 2005. A new and less destructive laboratory procedure for the physical separation of distal glass tephra shards from sediments. Quaternary Science Reviews 24, 19521960.Google Scholar
Brubaker, L.B., Anderson, P.A., Edwards, M.E., Lozhkin, A.V., 2005. Beringia as a glacial refugium for boreal trees and shrubs: new perspectives from mapped pollen data. Journal of Biogeography 32, 833848.Google Scholar
Brubaker, L.B., Garfinkel, H.L., Edwards, M.E., 1983. A late Wisconsin and Holocene vegetation history from the central Brooks Range: implications for Alaskan palaeoecology. Quaternary Research 20, 194214.Google Scholar
Carlson, L.J., Finney, B.P., 2004. A 13000-year history of vegetation and environmental change at Jan Lake, east-central Alaska. Holocene 14, 818827.Google Scholar
Cooper, A., Turney, C., Hughen, K.A., Brook, B.W., McDonald, H.G., Bradshaw, C.J.A., 2015. Abrupt warming events drove Late Pleistocene Holarctic megafaunal turnover. Science 349, 602606.Google Scholar
Coulter, S.E., Pilcher, J.R., Plunkett, G., Baillie, M., Hall, V.A., Steffensen, J.P., Vinther, B.M., Clausen, H.B., Johnsen, S.J., 2012. Holocene tephras highlight complexity of volcanic signals in Greenland ice cores. Journal of Geophysical Research: Atmospheres 117, D21303. http://dx.doi.org/10.1029/2012JD017698.CrossRefGoogle Scholar
Davies, L.J., Jensen, B.J.L., Froese, D.G., Wallace, K.L., 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, 2853.CrossRefGoogle Scholar
Davies, S.M., Elmquist, M., Bergman, J., Wohlfarth, B., Hammarlund, D., 2007. Cryptotephra sedimentation processes within two lacustrine sequences from west central Sweden. Holocene 17, 319330.Google Scholar
de Fontaine, C.S., Kaufman, D.S., Anderson, R.S., Werner, A., Waythomas, C.F., Brown, T.A., 2007. Late Quaternary distal tephra-fall deposits in lacustrine sediments, Kenai Peninsula, Alaska. Quaternary Research 68, 6478.Google Scholar
Denton, J.S., Pearce, N.J.G., 2008. Comment on “A synchronized dating of three Greenland ice cores through the Holocene” by B.M. Vinther et al.: no Minoan tephra in the 1642 B.C. layer of the GRIP ice core. Journal of Geophysical Research 113, D04303. http://dx.doi.org/10.1029/2007JD008970.Google Scholar
Edwards, M.E., Anderson, P.M., Garfinkel, H.L., Brubaker, L.B., 1985. Late Wisconsin and Holocene vegetation history of the upper Koyukuk region, central Brooks Range, Alaska. Canadian Journal of Botany 63, 616626.Google Scholar
Fierstein, J., Hildreth, W., 2008. Kaguyak dome field and its Holocene caldera, Alaska Peninsula. Journal of Volcanology and Geothermal Research 177, 340366.CrossRefGoogle Scholar
Guthrie, R.D., 2006. New carbon dates link climatic change with human colonization and Pleistocene extinctions. Nature 441, 207209.Google Scholar
Higuera, P.E., Brubaker, L.B., Anderson, P.M., Hu, F.S., Brown, T.A., 2009. Vegetation mediated the impacts of postglacial climate change on fire regimes in the south-central Brooks Range, Alaska. Ecological Monographs 79, 201219.Google Scholar
Jennings, A., Thordarson, T., Zalzal, K., Stoner, J., Hayward, C., Geirsdóttir, Á., Miller, G., 2014. Holocene tephra from Iceland and Alaska in SE Greenland shelf sediments. Geological Society, London, Special Publications 398, 157193.Google Scholar
Jensen, B.J.L., Froese, D.G., Preece, S.J., Westgate, J.A., Stachel, T., 2008. An extensive middle to late Pleistocene tephrochronologic record from east-central Alaska. Quaternary Science Reviews 27, 411427.CrossRefGoogle Scholar
Jensen, B.J.L., Pyne-O’Donnell, S., Plunkett, G., Froese, D.G., Hughes, P.D.M., Sigl, M., McConnell, J.R., et al. 2014. Transatlantic distribution of the Alaskan White River Ash. Geology 42, 875878.CrossRefGoogle Scholar
Kaufman, D.S., Jensen, B.J., Reyes, A.V., Schiff, C.J., Froese, D.G., Pearce, N.J., 2012. Late Quaternary tephrostratigraphy, Ahklun Mountains, SW Alaska. Journal of Quaternary Science 27, 344359.CrossRefGoogle Scholar
Kuehn, S.C., Froese, D.G., Shane, P.A.R., INTAV Intercomparison Participants. 2011. The INTAV intercomparison of electron-beam microanalysis of glass by tephrochronology laboratories: results and recommendations. Quaternary International 246, 1947.Google Scholar
Lerbekmo, J.F., 2008. The White River Ash: largest Holocene Plinian tephra. Canadian Journal of Earth Sciences 45, 693700.Google Scholar
Lowe, D.J., 2011. Tephrochronology and its application: a review. Quaternary Geochronology 6, 107153.Google Scholar
Mackay, H., Hughes, P.D.M., Jensen, B.J.L., Langdon, P.G., Pyne-O’Donnell, S.D.F., Plunkett, G., Froese, D.G., Coulter, S., 2016. A mid to late Holocene cryptotephra framework from eastern North America. Quaternary Science Reviews 132, 101113.Google Scholar
Mangerud, J., Lie, S.E., Furnes, H., Kristiansen, I.L., Lømo, L., 1984. A Younger Dryas ash bed in western Norway, and its possible correlations with tephra in cores from the Norwegian Sea and the North Atlantic. Quaternary Research 21, 85104.CrossRefGoogle Scholar
Miller, T.P., Smith, R.L., 1987. Late Quaternary caldera-forming eruptions in the eastern Aleutian arc, Alaska. Geology 15, 434438.Google Scholar
Neal, C.A., McGimsey, R.G., Miller, T.P., Riehle, J.R., Waythomas, C.F., 2001). Preliminary Volcano-Hazard Assessment for Aniakchak Volcano, Alaska. Open-File Report 00-519. U.S. Geological Survey, Anchorage, AK.Google Scholar
Oswald, W.W., Gavin, D.G., Anderson, P.M., Brubaker, L.B., Hu, F.-S., 2012. A 14,500-year record of landscape change from Okpilak Lake, northeastern Brooks Range, northern Alaska. Journal of Paleolimnology 48, 101113.CrossRefGoogle Scholar
Payne, R., Blackford, J., van der Plicht, J., 2008. Using cryptotephras to extend regional tephrochronologies: an example from southeast Alaska and implications for hazard assessment. Quaternary Research 69, 4255.Google Scholar
Payne, R.J., Blackford, J.J., 2004. Distal micro-tephra deposits in southeast Alaskan peatlands. In: Emond, D.S., Lewis, L.L. (Eds.), Yukon Exploration and Geology 2003. Yukon Geological Survey, Whitehorse, YT, Canada, pp. 191197.Google Scholar
Payne, R.J., Blackford, J.J., 2008. Extending the late Holocene tephrochronology of the Kenai Peninsula, Alaska. Arctic 61, 243254.Google Scholar
Pearce, C., Varhelyi, A., Wastegård, S., Muschitiello, F., Barrientos, N., O’Regan, M., Cronin, T., et al. 2016. The 3.6 ka Aniakchak tephra in the Arctic Ocean: a constraint on the Holocene radiocarbon reservoir age in the Chukchi Sea. Climate of the Past Discussions (in review). http://dx.doi.org/10.5194/cp-2016-112.Google Scholar
Preece, S.J., McGimsey, R.G., Westgate, J.A., Pearce, N.J.G., Hart, W.K., Perkins, W.T., 2014. Chemical complexity and source of the White River Ash, Alaska and Yukon. Geosphere 10, 10201042.Google Scholar
Pyne O’Donnell, S.D.F., 2011. The taphonomy of Last Glacial–Interglacial Transition (LGIT) distal volcanic ash in small Scottish lakes. Boreas 40, 131145.CrossRefGoogle Scholar
Pyne O’Donnell, S.D.F., Hughes, P.D.M., Froese, D.G., Jensen, B.J.L., Kuehn, S.C., Mallon, G., Amesbury, M.J., et al. 2012. High-precision ultra-distal Holocene tephrochronology in North America. Quaternary Science Reviews 52, 611.Google Scholar
Reimer, P., Bard, E., Bayliss, A., Beck, J., Blackwell, P., Bronk Ramsey, C., Buck, C., Cheng, H., Edwards, R., Friedrich, M., 2013. IntCal13 and Marine13 radiocarbon age calibration curves 0-50,000 years cal BP. Radiocarbon 55, 18691887.Google Scholar
Riehle, J.R., 1985. A reconnaissance of the major Holocene tephra deposits in the upper Cook Inlet region, Alaska. Journal of Volcanology and Geothermal Research 26, 3774.Google Scholar
Riehle, J.R., 1994. Heterogeneity, correlatives, and proposed stratigraphic nomenclature of Hayes tephra set H, Alaska. Quaternary Research 41, 285288.Google Scholar
Riehle, J.R., Meyer, C.E., Ager, T.A., Kaufman, D.S., Ackerman, R.E., 1987. The Aniakchak tephra deposit, a late Holocene marker horizon in western Alaska. U.S. Geological Survey Circular 998, 1922.Google Scholar
Riehle, J.R., Waitt, R.B., Meyer, C.E., Calk, L.C., 1998. Age of formation of Kaguyak caldera, eastern Aleutian arc, Alaska, estimated by tephrochronology. U.S. Geological Survey Professional Paper 1595, 161–168.Google Scholar
Serreze, M.C., Lynch, A.H., Clark, M.P., 2001. The Arctic frontal zone as seen in the NCEP-NCAR reanalysis. Journal of Climate 14, 15501567.Google Scholar
Turney, C.S.M., 1998. Extraction of rhyolitic component of Vedde microtephra from minerogenic lake sediments. Journal of Paleolimnology 19, 199206.Google Scholar
Wallace, K., Coombs, M.L., Hayden, L.A., Waythomas, C.F., 2014. Significance of a Near-Source Tephra-Stratigraphic Sequence to the Eruptive History of Hayes Volcano, South-Central Alaska. Scientific Investigations Report 2014-5133. U.S. Geological Survey, Reston, VA.Google Scholar
Wright, H.E. Jr., Mann, D.H., Glaser, P.H., 1984. Piston corers for peat and lake sediments. Ecology 65, 657659.Google Scholar
Zander, P.D., Kaufman, D.S., Kuehn, S.C., Wallace, K.L., Anderson, R.S., 2013. Early and late Holocene glacial fluctuations and tephrostratigraphy, Cabin Lake, Alaska. Journal of Quaternary Science 28, 761771.CrossRefGoogle Scholar
Zdanowicz, C., Fisher, D., Bourgeois, J., Demuth, M., Zheng, J., Mayewski, P.A., Kreutz, K., et al. 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, 3557.Google Scholar
Zdanowicz, C.M., Zielinski, G.A., Germani, M.S., 1999. Mount Mazama eruption: calendrical age verified and atmospheric impact assessed. Geology 27, 621624.Google Scholar
Supplementary material: Image

Monteath supplementary material

Monteath supplementary material 1

Download Monteath supplementary material(Image)
Image 40.7 MB
Supplementary material: Image

Monteath supplementary material

Monteath supplementary material 2

Download Monteath supplementary material(Image)
Image 31.9 MB
Supplementary material: PDF

Monteath supplementary material

Monteath supplementary material 3

Download Monteath supplementary material(PDF)
PDF 2.2 MB