Hostname: page-component-848d4c4894-4hhp2 Total loading time: 0 Render date: 2024-04-30T12:47:58.759Z Has data issue: false hasContentIssue false

Using cryptotephras to extend regional tephrochronologies: An example from southeast Alaska and implications for hazard assessment

Published online by Cambridge University Press:  20 January 2017

Richard Payne*
Affiliation:
Department of Geography, Queen Mary, University of London, Mile End Road, London, E1 4NS, UK Department of Biology, University of Bergen, Allègaten 41, N-5007 Bergen, Norway Geography, School of Environment and Development, The University of Manchester, Oxford Road, Manchester, M13 9PL, UK
Jeffrey Blackford*
Affiliation:
Geography, School of Environment and Development, The University of Manchester, Oxford Road, Manchester, M13 9PL, UK
Johannes van der Plicht*
Affiliation:
Centre for Isotope Research, University of Groningen, P.O. Box 72, 9700 AB Groningen, The Netherlands Faculty of Archaeology, Leiden University, The Netherlands
*
*Corresponding author. Geography, School of Environment and Development, The University of Manchester, Oxford Road, Manchester, M13 9PL, UKE-mail addresses:Richard.Payne-2@manchester.ac.uk (R. Payne).
*Corresponding author. Geography, School of Environment and Development, The University of Manchester, Oxford Road, Manchester, M13 9PL, UKE-mail addresses:Richard.Payne-2@manchester.ac.uk (R. Payne).
*Corresponding author. Geography, School of Environment and Development, The University of Manchester, Oxford Road, Manchester, M13 9PL, UKE-mail addresses:Richard.Payne-2@manchester.ac.uk (R. Payne).

Abstract

Cryptotephrochronology, the use of hidden, diminutive volcanic ash layers to date sediments, has rarely been applied outside western Europe but has the potential to improve the tephrochronology of other regions of the world. Here we present the first comprehensive cryptotephra study in Alaska. Cores were extracted from five peatland sites, with cryptotephras located by ashing and microscopy and their glass geochemistry examined using electron probe microanalysis. Glass geochemical data from nine tephras were compared between sites and with data from previous Alaskan tephra studies. One tephra present in all the cores is believed to represent a previously unidentified eruption of Mt. Churchill and is named here as the ‘Lena tephra’. A mid-Holocene tephra in one site is very similar to Aniakchak tephra and most likely represents a previously unidentified Aniakchak eruption, ca. 5300–5030 cal yr BP. Other tephras are from the late Holocene White River eruption, a mid-Holocene Mt. Churchill eruption, and possibly eruptions of Redoubt and Augustine volcanoes. These results show the potential of cryptotephras to expand the geographic limits of tephrochronology and demonstrate that Mt. Churchill has been more active in the Holocene than previously appreciated. This finding may necessitate reassessment of volcanic hazards in the region.

Type
Research Article
Copyright
Elsevier Inc.

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

Aaby, B., Digerfeldt, G., (1986). Sampling techniques for lakes and bogs. In: Berglund, B., (Ed.), Handbook of Holocene palaeoecology and palaeohydrology. Wiley, , Chichester.Google Scholar
Begét, J., Kienle, J., (1992). Cyclic formation of debris avalanches at Mount St Augustine volcano. Nature 356, 701704.Google Scholar
Begét, J., Motyka, R., (1998). New dates of late Pleistocene dacitic tephra from the Mount Edgecumbe Volcanic field, Southeastern Alaska. Quaternary Research 49, 123125.Google Scholar
Begét, J., Reger, R., Pinney, D., Gillispie, T., Campbell, K., (1991). Correlation of the Holocene Jarvis Creek, Tangle Lakes, Cantwell and Hayes tephras in south-central and central Alaska. Quaternary Research 35, 174189.Google Scholar
Begét, J., Mason, O., Anderson, P., (1992). Age, extent and climatic significance of the c. 3400 BP Aniakchak tephra, Western Alaska, USA. The Holocene 2, 5156.CrossRefGoogle Scholar
Begét, J.E., Stihler, S.D., Stone, D.B., (1994). A 500-year-long record of tephra falls from Redoubt volcano and other volcanoes in upper Cook Inlet, Alaska. Journal of Volcanology and Geothermal Research 62, 5567.Google Scholar
Bergman, J., Wastegård, S., Hammarlund, D., Wohlfarth, B., Roberts, S.J., (2004). Holocene tephra horizons at Klocka Bog, west-central Sweden: Aspects of reproducibility in subarctic peat deposits. Journal of Quaternary Science 19, 241249.CrossRefGoogle Scholar
Birks, H.H., Gulliksen, S., Haflidason, H., Mangerud, J., Possnert, G., (1996). New radiocarbon dates for the Vedde ash and the Saksunavtn ash from western Norway. Quaternary Research 45, 119127.Google Scholar
Blackford, J., Edwards, K., Dugmore, A., Cook, G., Buckland, P., (1992). Icelandic volcanic ash and mid-Holocene Scots pine (Pinus sylvestris) pollen decline in northern Scotland. The Holocene 2, 260265.Google Scholar
Blaauw, M., van Geel, B., Mauquoy, D., van der Plicht, J., (2004). Carbon-14 wiggle-match dating of peat deposits: Advantages and limitations. Journal of Quaternary Science 19, 177181.Google Scholar
Blockley, S.P.E., Pyne-O'Donell, S.D.F., Lowe, J.J., Pollard, A.M., Matthews, I.P., Molyneux, E.G., Turney, C.S.M., (2005). A new & less destructive laboratory procedure for the physical separation of distal glass tephra shards from sediments. Quaternary Science Reviews 24, 19521960.Google Scholar
Borchardt, G.A., Aruscavage, P.J., Miller, H.T., (1972). Correlation of the Bishop ash, a Pleistocene marker bed, using instrumental neutron activation analysis. Journal of Sedimentary Petrology 42, 301306.Google Scholar
Bronk Ramsey, C., (2005). OxCal program v.3.10. Oxford University .Google Scholar
Buckley, S., Walker, M.J.C., (2002). A mid-Flandrian tephra horizon, Cambrian Mountains, west Wales. Quaternary Newsletter 96, 511.Google Scholar
Caseldine, C.J., Hatton, J.M., Huber, U.M., Chiverrell, R., Woolley, N., (1998). Assessing the impact of volcanic activity on mid-Holocene climate in Ireland: The need for replicate data. The Holocene 8, 105111.Google Scholar
Chambers, F.M., Daniell, J.R.G., Hunt, J.B., Molloy, K., O 'Connell, M., (2004). Tephrostratigraphy of An Loch Mor, Inis Oirr, western Ireland: Implications for Holocene tephrochronology in the northeastern Atlantic region. The Holocene 14, 703720.Google Scholar
Child, J., Begét, J., Werner, A., (1998). Three Holocene tephras identified in Lacustrine sediment cores from the Wonder Lake area, Denali National Park and Preserve, Alaska USA. Arctic and Alpine Research 30, 8995.Google Scholar
Clague, J.J., Evans, S.G., Rampton, V.N., Woodsworth, G.J., (1995). Improved age estimates for the White River and Bridge River tephras, western Canada. Canadian Journal of Earth Sciences 32, 11721179.Google Scholar
Dachnowski-Stokes, A., (1941). Peat resources in Alaska. US Department of Agriculture Technical Bulletin 769, 184.Google Scholar
Davies, S.M., Hoek, W.Z., Bohncke, S.J.P., Lowe, J.J., Pyne O'Donnell, S., Turney, C.S.M., (2005). Detection of late-glacial distal tephra layers in the Netherlands. Boreas 34, 123135.Google Scholar
Downes, H., (1985). Evidence for magma heterogeneity in the White River Ash (Yukon Territory). Canadian Journal of Earth Sciences 22, 929934.Google Scholar
Dugmore, A., (1989). Icelandic volcanic ash in Scotland. Scottish Geographical Magazine 105, 168172.Google Scholar
Dugmore, A., Newton, A., (1992). Thin tephra layers in peat revealed by X-radiography. Journal of Archaeological Science 19, 163170.Google Scholar
Dugmore, A., Larsen, G., Newton, A., Sugden, D., (1992). Geochemical stability of fine-grained silicic Holocene tephra in Iceland and Scotland. Journal of Quaternary Science 7, 173183.Google Scholar
Dugmore, A., Larsen, G., Newton, A., (1995). Seven tephra isochrones in Scotland. The Holocene 5, 257266.Google Scholar
Dugmore, A.J., Newton, A.J., Edwards, K.J., Larsen, G., Blackford, J.J., Cook, G.T., (1996). Long-distance marker horizons from small-scale eruptions: British tephra deposits from the AD 1510 eruption of Hekla, Iceland. Journal of Quaternary Science 11, 511516.Google Scholar
Dwyer, R.B., Mitchell, F.J.G., (1997). Investigation of the environmental impact of remote volcanic activity on north Mayo, Ireland, during the mid-Holocene. The Holocene 7, 113118.Google Scholar
Froggatt, P.C., (1983). Toward a comprehensive upper Quaternary tephra and ignimbrite stratigraphy in New Zealand using electron microprobe analysis of glass shards. Quaternary Research 19, 188200.Google Scholar
Gehrels, M.J., Lowe, D.J., Hazell, Z.J., Newnham, R.M., (2006). A continuous 5300-yr Holocene cryptotephrostratigraphic record from northern New Zealand and implications for tephrochronology and volcanic hazard assessment. The Holocene 16, 173187.Google Scholar
Gronvold, K., Oskarsson, N., Johnsen, S.J., Clausen, H.B., Hammer, C.U., Bond, G., Bard, E., (1995). Ash layers from Iceland in the Greenland GRIP ice core correlated with oceanic and land based sediments. Earth and Planetary Science Letters 54, 238246.Google Scholar
Hodder, A.P.W., De Lange, P.J., Lowe, D.J., (1991). Dissolution and depletion of ferromagnesian minerals from Holocene tephra layers in an acid bog, New Zeland, and implications for tephra correlation. Journal of Quaternary Science 6, 195208.Google Scholar
Hunt, J., Hill, P., (1993). Tephra geochemistry: a discussion of some persistent analytical problems. The Holocene 3, 271278.Google Scholar
King, R.H., Kingston, M.S., Barnett, R.L., (1982). A numerical approach to the classification of magnetites from tephra in southern Alberta. Canadian Journal of Earth Sciences 19, 20122019.Google Scholar
Lerbekmo, J.F., Campbell, F.A., (1969). Distribution, composition and source of the White River Ash, Yukon Territory. Canadian Journal of Earth Sciences 6, 109116.Google Scholar
Lerbekmo, J.F., Westgate, J.A., Smith, D.G.W., Denton, G.H., (1975). New data on the character and history of the White River volcanic eruption, Alaska. Quaternary studies: Selected papers from IX INQUA congress. Royal Society of New Zealand Bulletin. vol. 13, 203209.Google Scholar
Lowe, D.J., Hunt, J.B., (2001). A summary of terminology used in tephra-related studies. Juvigné, E., Raynal, J-P., Tephra: Chronology, Archaeology. Les Dossiers de l'Archéo-Logis vol. 1, 1722.Google Scholar
Mangerud, J., Lie, S.E., Furnes, H., Kristiansen, I.L., Lomo, 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.Google Scholar
McKenzie, G., (1970). Some properties and age of volcanic ash in Glacier Bay National Monument. Arctic 23, 4649.Google Scholar
Merkt, J., Muller, H., Knabe, W., Muller, P., Weiser, T., (1993). The early Holocene Saksunarvatn Tephra found in lake sediments in N.W. Germany. Boreas 22, 93100.Google Scholar
Miller, T., Smith, R., (1987). Late Quaternary caldera-forming eruptions in the eastern Aleutian volcanic arc, Alaska. Geology 15, 434438.Google Scholar
Moodie, D.W., Catchpole, A.J.W., Abel, K., (1992). Northern Athapaskan oral traditions and the White River volcano. Ethnohistory 39, 148171.Google Scholar
Nilsson, M., Klarqvist, M., Bohlin, E., Possnert, G., (2001). Variation in 14C age of macrofossils and different fractions of minute peat samples dated by AMS. The Holocene 11, 579586.Google Scholar
Payne, R., Blackford, J., (2004). Distal tephra deposits in southeast Alaskan peatlands. Emond, D., Lewis, L., Yukon Exploration and Geology 2003. Yukon Geological Survey, Whitehorse., 191197.Google Scholar
Payne, R., Blackford, J., (2005). Microwave digestion and the geochemical stability of tephra. Quaternary Newsletter 106, 2433.Google Scholar
Payne, R., Kilfeather, A., van der Meer, J., Blackford, J., (2005). Experiments on the taphonomy of tephra in peatlands. Suo 56, 147156.Google Scholar
Persson, C., (1971). Tephrochronological investigations of peat deposits in Scandinavia and on the Faroe Islands. Sveriges Geologiska Undersokning 65, 334.Google Scholar
Pilcher, J., Hall, V., (1992). Towards a tephrochronology for the Holocene of the north of Ireland. The Holocene 2, 255259.Google Scholar
Pilcher, J., Hall, V., McCormac, F., (1995). Dates of Holocene Icelandic volcanic eruptions from tephra layers in Irish peats. The Holocene 5, 103110.Google Scholar
Pilcher, J.R., Hall, V.A., (1996). A tephrochronology for the Holocene of the north of England. The Holocene 6, 100105.Google Scholar
Pollard, A.M., Blockley, S.P.E., Ward, K.R., (2003). Chemical alteration of tephra in the depositional environment: theoretical stability modeling. Journal of Quaternary Science 18, 385394.Google Scholar
Richter, D.H., Preece, S.J., McGimsey, R.G., Westgate, J.A., (1995). Mount Churchill, Alaska: Source of the late Holocene White River Ash. Canadian Journal of Earth Sciences 32, 741748.Google Scholar
Riehle, J.R., (1985). A reconnaissance of the major Holocene tephra deposits in the upper Cook Inlet, Alaska. Journal of Volcanology and Geothermal Research 26, 3754.Google Scholar
Riehle, J.R., (1996). Mount Edgecumbe volcanic field: A geologic history. U.S. Forest Service, Juneau, Alaska.Google Scholar
Riehle, J.R., Brew, D.A., (1984). Explosive latest Pleistocene(?) and Holocene activity of the Mount Edgecumbe Volcanic Field, Alaska. U.S. Geological Survey Circular 939, 111114.Google Scholar
Riehle, J., Bowers, P., (1990). The Hayes tephra deposits, an upper Holocene marker horizon in south-central Alaska. Quaternary Research 33, 276290.Google Scholar
Riehle, J., Meyer, C., Ager, T., Kaufmann, D., Ackerman, R., (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., Mann, D.H., Peteet, D.M., Engstrom, D.R., Brew, D.A., Meyer, C.E., (1992). The Mount Edgecumbe tephra deposits, a marker horizon in southeastern Alaska near the Pleistocene–Holocene boundary. Quaternary Research 37, 183202.Google Scholar
Riehle, J.R., Meyer, C.E., Miyaoka, R., (1999). Data on Holocene tephra (volcanic ash) deposits in the Alaska peninsula and lower Cook Inlet region of the Aleutian volcanic arc, Alaska. U.S. Geological Survey Open-File 99135.Google Scholar
Riehle, J.R., Ager, T.A., Reger, R.D., Pinney, D.S., Kaufman, D.S., in press. Stratigraphic and compositional complexities of the late Quaternary Lethe tephra in south-central Alaska. Quaternary International. doi:10.1016/j.quaint.2006.09.006.Google Scholar
Rigg, G., (1914). Notes on the flora of some Alaskan Sphagnum bogs. The Plant World 17, 167182.Google Scholar
Rigg, G., (1937). Some raised bogs of Southeastern Alaska with notes on flat bogs and muskegs. American Journal of Botany 24, 194198.Google Scholar
Robinson, S.D., (2001). Extending the late Holocene White River Ash distribution, northwestern Canada. Arctic 54, 157161.Google Scholar
Robinson, S.D., Moore, T.R., (1999). Carbon and peat accumulation over the past 1200 years in a landscape with discontinuous permafrost, northwestern Canada. Global Biogeochemical cycles 13, 591601.Google Scholar
Rose, N., Golding, P., Battarbee, R., (1996). Selective concentration and enumeration of tephra shards from lake sediment cores. The Holocene 6, 243246.Google Scholar
Shane, P.A.R., Froggatt, P.C., (1994). Discriminant function analysis of New Zealand and North American tephra deposits. Quaternary Research 41, 7081.Google Scholar
Stockmarr, J., (1971). Tablets with Spores used in Absolute Pollen Analysis. Pollen et Spores 13, 615621.Google Scholar
Stokes, S., Lowe, D.J., (1988). Discriminant function analysis of late Quaternary tephras from five volcanoes in New Zealand using glass shard major element chemistry. Quaternary Research 30, 270283.Google Scholar
Telford, R.J., Heegaard, E., Birks, H.J.B., (2004). All age–depth models are wrong: But how badly?. Quaternary Science Reviews 23, 15.Google Scholar
Ter Braak, C., Šmilauer, P., CANOCO for Windows. Version 4.53, Biometris-Plant Research, Wageningen, , The Netherlands.Google Scholar
Van den Bogaard, C., Schmincke, H.-U., (2002). Linking the North Atlantic to central Europe: A high-resolution tephrochronological record from northern Germany. Journal of Quaternary Science 17, 320.Google Scholar
VanderHoek, R., Myron, R., (2004). Cultural Remains from a Catastrophic Landscape: An Archeological Overview and Assessment of Aniakchak National Monument and Preserve. United States Department of the Interior, National Park Service, Aniakchak National Monument and Preserve.Google Scholar
van den Bogaard, C., Schmincke, H.-U., (2002). Linking the North Atlantic to central Europe: a high resolution Holocene tephrochronological record from northern Germany. Journal of Quaternary Science 17, 320.Google Scholar
Waitt, R., Begét, J., in press. Geologic mapping at Augustine Volcano, Alaska. USGS Professional Paper.Google Scholar
Wastegård, S., Björck, S., Grauert, M., Hannon, G.E., (2001). The Mjáuvøtn tephra and other Holocene tephra horizons from the Faroe Islands: a link between the Icelandic source region, the Nordic Seas, and the European continent. The Holocene 11, 101109.Google Scholar
Workman, W.B., (1979). The significance of volcanism in the prehistory of subarctic northwest North America. In: Sheets, P., Grayson, D.K., (Eds.) Volcanic Activity and Human Ecology. Academic Press, , New York., pp. 339-371.,.CrossRefGoogle Scholar
Zoltai, S., (1988). Late Quaternary volcanic ash in the peatlands of central Alberta. Canadian Journal of Earth Science 26, 207214.CrossRefGoogle Scholar