Hostname: page-component-848d4c4894-xm8r8 Total loading time: 0 Render date: 2024-06-25T00:12:23.510Z Has data issue: false hasContentIssue false

Regional atmospheric circulation change in the North Pacific during the Holocene inferred from lacustrine carbonate oxygen isotopes, Yukon Territory, Canada

Published online by Cambridge University Press:  20 January 2017

Lesleigh Anderson*
Affiliation:
Department of Geosciences, University of Massachusetts Amherst, Amherst, MA 01003, USA
Mark B. Abbott*
Affiliation:
Department of Geology and Planetary Science, University of Pittsburgh, Pittsburgh, PA 15260-3332, USA
Bruce P. Finney*
Affiliation:
Institute of Marine Science, University of Alaska Fairbanks, Fairbanks, AK 99775, USA
Stephen J. Burns*
Affiliation:
Department of Geosciences, University of Massachusetts Amherst, Amherst, MA 01003, USA
*
*Corresponding author. Fax: +1 413 545 1200. E-mail addresses: land@geo.umass.edu (L. Anderson), mabbott1@pitt.edu (M.B. Abbott), finney@ims.uaf.edu (B.P. Finney), sburns@geo.umass.edu (S.J. Burns)
1Fax: +1 412 624 3914.
2Fax: +1 907 474 7204.
3Fax: +1 413 545 1200.
Rights & Permissions [Opens in a new window]

Abstract

Core share and HTML view are not available for this content. However, as you have access to this content, a full PDF is available via the ‘Save PDF’ action button.

Analyses of sediment cores from Jellybean Lake, a small, evaporation-insensitive groundwater-fed lake, provide a record of changes in North Pacific atmospheric circulation for the last ∽7500 yr at 5- to 30-yr resolution. Isotope hydrology data from the southern Yukon indicate that the oxygen isotope composition of water from Jellybean Lake reflects the composition of mean-annual precipitation, δ18Op. Recent changes in the δ18O of Jellybean sedimentary calcite (δ18Oca) correspond to changes in the North Pacific Index (NPI), a measure of the intensity and position of the Aleutian Low (AL) pressure system. This suggests that δ18Op variability was related to the degree of fractionation during moisture transport from the Gulf of Alaska across the St. Elias Mountains and that Holocene shifts were controlled by the intensity and position of the AL. Following this model, between ∽7500 and 4500 cal yr B.P., long-term trends suggest a predominantly weaker and/or westward AL. Between ∽4500 and 3000 cal yr B.P. the AL shifted eastward or intensified before shifting westward or weakening between ∽3000 and 2000 cal yr B.P. Rapid shifts eastward and/or intensification occurred ∽1200 and 300 cal yr B.P. Holocene changes in North Pacific atmospheric circulation inferred from Jellybean Lake oxygen isotopes correspond with late Holocene glacial advances in the St. Elias Mountains, changes in North Pacific salmon abundance, and shifts in atmospheric circulation over the Beaufort Sea.

Type
Research Article
Copyright
University of Washington

References

Anderson, L. (2005). Holocene climate of the southwest Yukon Territory, Canada, inferred from lake-level and isotope analyses of small carbonate lakes. Ph.D. Dissertation,University of Massachusetts Amherst, Department of Geosciences, Amherst, MA, USA.Google Scholar
Anderson, L., Abbott, M.B., and Finney, B.P. (2001). Holocene climate inferred from oxygen isotope ratios in lake sediments, Central Brooks Range, Alaska. Quaternary Research 55, 313321.CrossRefGoogle Scholar
Appleby, P.G. (2001). Chronostratigraphic techniques in recent sediments.Last, W., Smol, J.P. Tracking Environmental Change Using Lake Sediments Basin Analysis, Coring, and Chronological Techniques vol. 1, Kluwer Academic Press, Dordrecht.171203.Google Scholar
Bengtsson, L., and Enell, M. (1986). Chemical analyses.Berglund, B.E. Handbook of Holocene Paleoecology and Paleohydrology J. Wiley, Chichester.423451.Google Scholar
Biondi, F., Gershunov, A., and Cayan, D.R. (2001). North Pacific decadal climate variability since 1661. Journal of Climate 14, 510.2.0.CO;2>CrossRefGoogle Scholar
Bitz, C.M., and Battisti, D.S. (1999). Interannual to decadal variability in climate and the glacier mass balance in Washington, Western Canada and Alaska. Journal of Climate 12, 31813196.Google Scholar
Bowen, G.J., and Wilkinson, B. (2002). Spatial distribution of δ18O in meteoric precipitation. Geology 30, 315318.2.0.CO;2>CrossRefGoogle Scholar
Brown, T.A., Nelson, D.E., Mathewes, R.W., Vogel, J.S., and Southon, J.R. (1989). Radiocarbon dating of pollen by accelerator mass spectrometry. Quaternary Research 32, 205212.Google Scholar
Calkin, P.E., Wiles, G.C., and Barclay, D.J. (2001). Holocene coastal glaciation of Alaska. Quaternary Science Reviews 20, 449461.CrossRefGoogle Scholar
Cayan, D.R., and Peterson, D.H. (1989). The influence of North Pacific atmospheric circulation on streamflow in the west. Geophysical Monograph 55, 375379.Google Scholar
Clague, J.J., Evans, S.G., Rampton, V.N., and Woodsworth, G.J. (1995). Improved age estimates for the White River and Bridge River tephras, Western Canada. Canadian Journal of Earth Sciences 32, 11721179.CrossRefGoogle Scholar
Cwynar, L.C. (1988). Late Quaternary vegetation history of Kettlehole Pond, southwestern Yukon. Canadian Journal of Forest Research 18, 12701279.Google Scholar
Cwynar, L.C., and Spear, R.W. (1995). Paleovegetation and paleoclimatic changes in the Yukon at 6 kA BP. Geographie Physique et Quaternaire 49, 2935.CrossRefGoogle Scholar
Dansgaard, W. (1964). Stable isotopes in precipitation. Tellus 16, 436468.Google Scholar
D'Arrigo, R., Wiles, G., Jacoby, G., and Villaba, R. (1999). North Pacific sea surface temperatures: past variations inferred from tree rings. Geophysical Research Letters 26, 27572760.CrossRefGoogle Scholar
Davi, N.K., Jacoby, G.C., and Wiles, G.C. (2003). Boreal temperature variability inferred from maximum latewood density and tree-ring width data, Wrangell Mountain region, Alaska. Quaternary Research 60, 252262.CrossRefGoogle Scholar
Dean, W.E., and Stuiver, M. (1993). Stable carbon and oxygen isotope studies of the sediments of Elk Lake, Minnesota.Bradbury, J.P., Dean, W.F. Elk Lake, Minnesota: Evidence for Rapid Climate Change in the North-Central United States United States Geological Survey Special Paper vol. 276, 163180.Boulder, ColoradoGoogle Scholar
Denton, G.H., and Karlen, W. (1977). Holocene glacial and tree-line variations in the White River Valley and Skolai Pass, Alaska and Yukon Territory. Quaternary Research 7, 63111.CrossRefGoogle Scholar
Dyke, A.S., and Savelle, J.M. (2000). Holocene driftwood incursion to Southwestern Victoria Island, Canadian Arctic Archipelago, and its significance to Paleoceanography and Archeology. Quaternary Research 54, 113120.CrossRefGoogle Scholar
Dyke, A.S., and Savelle, J.M. (2001). Holocene history of the Bering Sea Bowhead whale (Balaena mysticetus) in its Beaufort Sea summer grounds off southwestern Victoria Island, western Canadian Arctic. Quaternary Research 55, 371379.Google Scholar
Dyke, A.S., Andrews, J.T., Clark, P.U., England, J.H., Miller, G.H., Shaw, J., and Veillette, J.J. (2002). The Laurentide and Innuitian ice sheets during the last glacial maximum. Quaternary Science Reviews 21, 931.Google Scholar
Edwards, T.W.D., Wolfe, B.B., and MacDonald, G.M. (1996). Influence of changing atmospheric circulation on precipitation δ18O-temperature relations in Canada during the Holocene. Quaternary Research 46, 211218.Google Scholar
Edwards, M.E., Mock, C.J., Finney, B.P., Barber, V.A., and Bartlein, P.J. (2001). Potential analogues for paleoclimatic variations in eastern interior Alaska during the past 14,000 yr: atmospheric-circulation controls of regional temperature and moisture responses. Quaternary Science Reviews 20, 189202.CrossRefGoogle Scholar
Environment Canada, (2003). Environment Canada, 2003. Canadian Climate Normals Web site.http://www.msc-smc.ec.gc.ca/climate/climate_normals/index_e.cfm.Google Scholar
Epstein, S., Buchsbaum, R., Lowenstam, H.A., and Urey, H.C. (1953). Revised carbonate-water temperature scale. Geological Society of America Bulletin 62, 417426.Google Scholar
Finney, B.P., Gregory-Eaves, I., Sweetman, J., Douglas, M.S.V., and Smol, J.P. (2000). Impacts of climatic change and fishing on Pacific salmon abundance over the past 300 years. Science 290, 795799.Google Scholar
Finney, B.P., Gregory-Eaves, I., Douglas, M.S.V., and Smol, J.P. (2002). Fisheries productivity in the northeastern Pacific Ocean over the past 2,200 years. Nature 416, 729733.Google Scholar
Friedman, I., and O'Niel, J.R. (1977). Compilation of stable isotope fractionation factors of geochemical interest.Fleischer, M. Data of Geochemistry, Chapter KK. United States Geological Survey Professional Paper 440-KK, Boulder sixth ed.12 Google Scholar
Gonfiantini, R., Frohlich, K., Araguas-Araguas, L., and Rozanski, K. (1998). Isotopes in groundwater hydrology.Kendall, C., McDonnell, J.J. Isotope Tracers in Catchment Hydrology Elsevier, Amsterdam.203246.Google Scholar
Hammerlund, D., Barnekow, L., Birks, H.J.B., Buchardt, B., and Edwards, T.W.D. (2002). Holocene changes in atmospheric circulation recorded in the oxygen-isotope stratigraphy of lacustrine carbonates from northern Sweden. The Holocene 12, 339351.Google Scholar
Hammerlund, D., Björck, S., Buchardt, B., Israelson, C., and Thomsen, C.T. (2003). Rapid hydrological changes during the Holocene revealed by stable isotope records of lacustrine carbonates from Lake Igelsjön, Southern Sweden. Quaternary Science Reviews 22, 353370.Google Scholar
Hare, S.R., and Francis, R.C. (1994). Climate change and salmon production in the Northeast Pacific Ocean.Beamish, R.J. Climate Change and Northern Fish Populations Canadian Special Publication in Fish and Aquatic Science vol. 121, 133.Google Scholar
Heiri, O., Lotter, A.F., and Lemcke, G. (2001). Loss on ignition as a method for estimating organic and carbonate content in sediments: reproducibility and comparability of results. Journal of Paleolimnology 25, 101110.CrossRefGoogle Scholar
Herczeg, A.L., and Fairbanks, R.G. (1987). Anomalous carbon isotope fractionation between atmospheric CO2 and dissolved inorganic carbon induced by intense photosynthesis. Geochimica et Cosmochimica Acta 51, 895899.Google Scholar
Heusser, C.J., Heusser, L.E., and Peteet, D.M. (1985). Late-quaternary climatic change on the American North Pacific Coast. Nature 316, 485487.Google Scholar
Holdsworth, G., Krouse, H.R., and Nosal, M. (1992). Ice core climate signals from Mount Logan, Yukon A.D. 1700–1897.Bradley, R.S., Jones, P.D. Climate Since A.D. 500 Routelidge, London.483516.Google Scholar
Hu, F.S., and Shemesh, A. (2003). A biogenic-silica δ18O record of climatic change during the last glacial-interglacial transition in southwestern Alaska. Quaternary Research 59, 379385.Google Scholar
Hu, F.S., Ito, E., Brown, T.A., Curry, B.B., and Engstron, D.R. (2001). Pronounced climatic variations in Alaska during the last two millennia. Proceedings of the National Academy of Sciences 98, 1055210556.Google Scholar
Hu, F.S., Kaufman, D., Yoneji, S., Nelson, D., Shemesh, A., Huang, Y., Tian, J., Bond, G., Clegg, B., and Brown, T. (2003). Cyclic variation and solar forcing of Holocene climate in the Alaskan subarctic. Science 301, 18901893.CrossRefGoogle ScholarPubMed
IAEA/WMO,(2001). Global Network of Isotopes in Precipitation (GNIP) and Isotope Hydrology Information System (ISOHIS).International Atomic Energy Agency and World Meteorological Organization.http://www.isohis.iaea.org/.Google Scholar
Ingraham, N.L. (1998). Isotopic variations in precipitation.Kendall, C., McDonnell, J.J. Isotope Tracers in Catchment Hydrology Elsevier, Amsterdam.87118.Google Scholar
Jones, V.J., Leng, M.J., Solovieva, N., Sloane, H.J., and Tarasov, P. (2004). Holocene climate of the Kola Peninsula; evidence from the oxygen isotope record of diatom silica. Quaternary Science Reviews 23, 833839.CrossRefGoogle Scholar
Keenan, T.J., and Cwynar, L.C. (1992). Late Quaternary history of black spruce and grasslands in southwest Yukon Territory. Canadian Journal of Botany 70, 13361345.CrossRefGoogle Scholar
Lacourse, T., and Gajewski, K. (2000). Late Quaternary vegetation history of Sulfur Lake, southwest Yukon Territory, Canada. Arctic 53, 2735.Google Scholar
Latif, M., and Barnett, T.P. (1994). Causes of decadal climate variability over the North Pacific and North America. Science 266, 634637.Google Scholar
Leng, M.J., and Anderson, N.J. (2003). Isotopic variation in modern lake waters from western Greenland. The Holocene 13, 605611.Google Scholar
Mann, D.H., Crowell, A.L., Hamilton, T.D., and Finney, B.P. (1998). Holocene geologic and climatic history around the Gulf of Alaska. Arctic Anthropology 35, 112131.Google Scholar
Mantua, N.J., Hare, S.R., Zhang, Y., Wallace, J.M., and Francis, R.C. (1997). A Pacific interdecadal climate oscillation with impacts on salmon production. Bulletin of the American Meteorological Society 78, 10691079.Google Scholar
McKenzie, J.A. (1985). Carbon isotopes and productivity in the lacustrine and marine environment.Stumm, W. Chemical Processes in Lakes J. Wiley, Toronto.99118.Google Scholar
Mock, C.J., Bartlein, P.J., and Anderson, P.M. (1998). Atmospheric circulation patterns and spatial climatic variations in Beringia. International Journal of Climatology 18, 10851104.3.0.CO;2-K>CrossRefGoogle Scholar
Moore, G.W.K., Alverson, K., and Holdsworth, G. (2002a). )Variability in the climate of the Pacific Ocean and North America as expressed in the Mount Logan ice core. Annals of Glaciology 35, 423429.Google Scholar
Moore, G.W.K., Holdsworth, G., and Alverson, K. (2002b). )Climate change in the North Pacific region over the past three centuries. Nature 420, 401403.Google Scholar
Oana, S., and Deevey, E.S. (1960). Carbon-13 in lake waters, and its possible bearing on paleolimnology. American Journal of Science 258-A, 253272.Google Scholar
Pienitz, R., Smol, J.P., Last, W.M., Leavitt, P.R., and Cumming, B.F. (2000). Multi-proxy Holocene paleoclimatic record from a saline lake in the Canadian subarctic. The Holocene 10, 673686.CrossRefGoogle Scholar
Plummer, L.N., and Busenberg, E. (1982). The solubility of calcite, aragonite, and vaterite in CO2–H2O solutions between 0 and 90°C, and an evaluation of the aqueous model for CaCO3–CO2–H2O equilibria. Geochimica et Cosmochimica Acta 44, 10111040.Google Scholar
Rosqvist, G., Jonsson, C., Yam, R., Karlén, W., and Shemesh, A. (2004). Diatom oxygen isotopes in pro-glacial lake sediments from northern Sweden: a 5000 year record of atmospheric circulation. Quaternary Science Reviews 23, 851859.CrossRefGoogle Scholar
Rozanski, K., Araguas-Araguas, L., and Gonfiantini, R. (1992). Relation between long-term trends of oxygen-18 isotope composition of precipitation and climate. Science 258, 981985.CrossRefGoogle ScholarPubMed
Rozanski, K., Araguás-Araguás, L., and Gonfiantini, R. (1993). Isotopic patterns in modern global precipitation.Swart, P.K., Lohman, K.C., McKenzie, J., Savin, S. Climate Change in Continental Isotopic Records American Geophysical Union, Geophysical Monograph vol. 78, 136.Washington, DCGoogle Scholar
Rupper, S., Steig, E.J., and Roe, G. (2004). The relationship between snow accumulation at Mt. Logan, Yukon, Canada, and climate variability in the North Pacific. Journal of Climate 17, 47244739.Google Scholar
Shemesh, A., Rosqvist, G., Rietti-Shati, M., Rubensdotter, L., Bigler, C., Yam, R., and Karlén, W. (2001). Holocene climatic change in Swedish Lapland inferred from an oxygen-isotope record of lacustrine biogenic silica. The Holocene 11, 447454.Google Scholar
Siegenthaler, U., and Eicher, U. (1986). Stable oxygen and carbon isotope analyses.Berglund, B.E. Handbook of Holocene Paleoecology and Paleohydrology J. Wiley, Chichester.407422.Google Scholar
Spear, R.W., and Cwynar, L.C. (1997). Late Quaternary vegetation history of White Pass, Northern British Columbia, Canada. Arctic and Alpine Research 29, 4552.Google Scholar
Spooner, I.S., Barnes, S., Baltzer, K.B., Raeside, R., Osborn, G.D., and Mazzucchi, D. (2003). The impact of air mass circulation dynamics on Late Holocene paleoclimate in northwestern North America. Quaternary International 108, 7783.Google Scholar
Streten, N.A. (1974). Some features of the summer climate of interior Alaska. Arctic 27, 272286.Google Scholar
Stuiver, M., Reimer, P.J., Bard, E., Beck, J.W., Burr, G.S., Hughen, K.A., Kromer, B., McCormac, F.G., v.d.Plicht, J., and Spurk, M. (1998). Radiocarbon age calibration 24,000-0 cal BP. Radiocarbon 40, 10411083.Google Scholar
Szeicz, J.M., and MacDonald, G.M. (1995). Dendroclimatic reconstruction of summer temperature in Northwestern Canada since A.D. 1638 based on age-dependent modeling. Quaternary Research 44, 257266.CrossRefGoogle Scholar
Szeicz, J.M., and MacDonald, G.M. (1996). A 930-year ring-width chronology from moisture sensitive white spruce (Picea glauca Moench) in Northwestern Canada. The Holocene 6, 345351.Google Scholar
Talbot, M.R. (1990). A review of the palaeohydrological interpretation of carbon and oxygen isotopic ratios in primary lacustrine carbonates. Chemical Geology, Isotope Geosciences Section 80, 261279.Google Scholar
Thompson, R. (1986). Paleomagnetic dating.Berglund, B.E. Handbook of holocene paleoecology and paleohydrology J. Wiley, Chichester.313328.Google Scholar
Trenbirth, K.E., and Hurrell, J.W. (1994). Decadal atmosphere-ocean variations in the Pacific. Climate Dynamics 9, 303319.Google Scholar
von Grafenstein, U., Erlenkeuser, H., Muller, J., Trimborn, P., and Alefs, J. (1996). A 200 lake sediments: an extension of the δ18Op-year mid-European air temperature record preserved in air temperature relation into the past. Geochimica et Cosmochimica Acta 60, 40254036.Google Scholar
Wahl, H.E., Fraser, D.B., Harvey, R.C., and Maxwell, J.B. (1987). Climate of the Yukon, Climatological Studies Number 40. Environment Canada, Ottawa.323 Google Scholar
Wake, C.P., Yalcin, K., and Gundestrup, N.S. (2002). The climate signal recorded in the oxygen-isotope, accumulation and major-ion time series from the Eclipse ice core, Yukon Territory, Canada. Annals of Glaciology 35, 416422.Google Scholar
Wang, X-C., and Geurts, M. (1991). Late Quaternary pollen records and vegetation history of the southwest Yukon Territory: a review. Geographié Physique et Quaternaire 45, 175193.Google Scholar
Wiles, G.C., D'Arrigo, R.D., and Jacoby, G.C. (1998). Gulf of Alaska atmosphere-ocean variability over recent centuries inferred from coastal tree-ring records. Climatic Change 38, 289306.CrossRefGoogle Scholar
Wiles, G.C., Barclay, D.J., and Calkin, P.E. (1999). Tree-ring-dated ‘Little Ice Age’ histories of maritime glaciers from western Prince William Sound, Alaska. The Holocene 9, 163173.Google Scholar
Wiles, G.C., D'Arrigo, R.D., Villalba, R., Calkin, P.E., and Barclay, D.J. (2004). Century-scale solar variability and Alaskan temperature change over the past millennium. Geophysical Research Letters 31, LI5203 10.1029/2004GL020050CrossRefGoogle Scholar
Wolfe, B.B., Edwards, T.W.D., Jiang, H., MacDonald, G.M., Gervais, B.R., and Snyder, J.A. (2003). Effect of varying oceanicity on early- to mid-Holocene palaeohydrology, Kola Peninsula, Russia: isotopic evidence from treeline lakes. The Holocene 13, 153160.Google Scholar
Yu, Z., McAndrews, J.H., and Eicher, U. (1997). Middle Holocene dry climate caused by change in atmospheric circulation patterns: evidence from lake levels and stable isotopes. Geology 25, 251254.2.3.CO;2>CrossRefGoogle Scholar