Hostname: page-component-77c89778f8-n9wrp Total loading time: 0 Render date: 2024-07-22T11:34:17.182Z Has data issue: false hasContentIssue false

Holocene climate–fire–vegetation interactions at a subalpine watershed in southeastern British Columbia, Canada

Published online by Cambridge University Press:  20 January 2017

Colin J. Courtney Mustaphi*
Affiliation:
Ottawa-Carleton Geoscience Centre, Department of Earth Sciences, Carleton University, Ottawa, Ontario K1S 5B6, Canada
Michael F.J. Pisaric
Affiliation:
Ottawa-Carleton Geoscience Centre, Department of Earth Sciences, Carleton University, Ottawa, Ontario K1S 5B6, Canada Department of Geography and Environmental Studies, Carleton University, Ottawa, Ontario K1S 5B6, Canada
*
*Corresponding author at: York Institute for Tropical Ecosystems, Environment Department, University of York, Heslington, York YO10 5DD, United Kingdom. E-mail address:colin.courtney-mustaphi@york.ac.uk (C.J. Courtney Mustaphi).

Abstract

Vegetation assemblages and associated disturbance regimes are spatially heterogeneous in mountain ecosystems throughout the world due to the complex terrain and strong environmental gradients. Given this complexity, numerous sites describing postglacial vegetation and fire histories are needed to adequately understand forest development and ecosystem responses to varying climate and disturbance regimes. To gain insight into long-term historical climate–fire–vegetation interactions in southeastern British Columbia, Canada, sedimentological and paleoecological analyses were performed on a sediment core recovered from a small subalpine lake. The pollen assemblages, stomata, and macroremains indicate that from 9500 to 7500 cal yr BP, Pinus-dominated forests occurred within the catchment and Alnus was also present. Climate was an important control of fire and fire frequency was highest at this time, peaking at 8 fires 1000 yr− 1, yet charcoal accumulation rates were low, indicative of low terrestrial biomass abundance. From 7500 to 4600 cal yr BP, Pinus decreased as Picea, Abies and Larix increased and fire frequencies decreased to 3–6 fires 1000 yr− 1. Since 7500 cal yr BP the fire regime varied at a millennial scale, driven by forest biomass abundance and fuel accumulation changes. Local scale (bottom-up) controls of fire increased in relative importance since at least 6000 cal yr BP.

Type
Research Article
Copyright
University of Washington

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

Bacon, C.R. Eruptive history of Mount Mazama and Crater Lakes caldera, Cascade Range, USA. Journal of Volcanology and Geothermal Research 18, (1983). 57115.CrossRefGoogle Scholar
Bamber, R.N. Sodium hexametaphosphate as an aid in benthic sample sorting. Marine Environmental Research 7, (1982). 251255.CrossRefGoogle Scholar
Bassett, I.J., Crompton, C.W., and Parmelee, J.A. An atlas of airborne pollen grains and common fungus spores of Canada. Canada Department of Agriculture Research Branch monograph no. 18. (1978). Supply and Services Canada, Hull, Canada.Google Scholar
Berger, A., and Loutre, M.F. Insolation values for the climate of the last 10 million years. Quaternary Science Reviews 10, (1991). 297317.CrossRefGoogle Scholar
Blaauw, M., and Christen, J.A. Flexible paleoclimate age-depth models using an autoregressive gamma process. Bayesian Analysis 6, (2011). 457474.CrossRefGoogle Scholar
Blarquez, O., and Carcaillet, C. Fire, fuel composition and subalpine ecosystem resilience threshold in subalpine ecosystem. PLoS ONE 5, (2010). e12480 Google Scholar
Blott, S.J., and Pye, K. GRADISTAT: a grain size distribution and statistics package for the analysis of unconsolidated sediments. Earth Surface Processes and Landforms 26, (2001). 12371248.Google Scholar
Booth, R.K. Testate amoebae as proxies for mean annual water-table depth in Sphagnum-dominated peatlands of North America. Journal of Quaternary Science 23, (2008). 4357.Google Scholar
Carter, V.A., Brunelle, A., Minckley, T.A., Dennison, P.E., and Power, M.J. Regionalization of fire regimes in the Central Rocky Mountains, USA. Quaternary Research 80, (2013). 406416.Google Scholar
Chase, M., Bleskie, C., Walker, I.R., Gavin, D.G., and Hu, F.S. Midge-inferred Holocene summer temperatures in Southeastern British Columbia, Canada. Palaeogeography, Palaeoclimatology, Palaeoecology 257, (2008). 244259.Google Scholar
Clague, J.J., and Mathewes, R.W. Early Holocene thermal maximum in western North America: new evidence from Castle Peak, British Columbia. Geology 17, (1989). 277280.2.3.CO;2>CrossRefGoogle Scholar
Clark, J.S. Particle motion and the theory of charcoal analysis: source area, transport, deposition, and sampling. Quaternary Research 30, (1988). 6780.CrossRefGoogle Scholar
Conedera, M., Tinner, W., Neff, C., Meurer, M., Dickens, A.F., and Krebs, P. Reconstructing past fire regimes: methods, applications, and relevance to fire management and conservation. Quaternary Science Reviews 28, (2009). 555576.CrossRefGoogle Scholar
Coupé, R.A., Stewart, A.C., and Wikeem, B.M. Engelmann spruce–subalpine fir zone. Meidinger, D.V., and Pojar, J. Ecosystems of British Columbia. Special Report Series 6, (1991). BC Ministry of Forests, Victoria. 223236.Google Scholar
Courtney Mustaphi, C.J. A landscape-scale assessment of Holocene fire regime controls in south-eastern British Columbia, Canada. (PhD thesis) (2013). Carleton University, Ottawa, Ontario, Canada.Google Scholar
Courtney Mustaphi, C.J., and Pisaric, M.F.J. Synchronous forest fire regimes between watersheds with similar aspect during the late Holocene, southeastern British Columbia, Canada. Journal of Biogeography 40, (2013). 19831996. http://dx.doi.org/10.1111/jbi.12143Google Scholar
Da Silva, E. Wildfire history and its relationship with top-down and bottom-up controls in the Joseph and Gold Creek watersheds, Kootenay Mountains, British Columbia. (Unpublished MSc. Thesis) (2009). Department of Geography, University of Guelph, Guelph, Canada.Google Scholar
Daniau, A.-L., Tinner, W., Bartlein, P.J., Harrison, S.P., Prentice, I.C., Brewer, S., Friedlingstein, P., Harrison-Prentice, T.I., Inoue, J., Marlon, J.R. et al. Predictability of biomass burning in response to climate changes. Global Biogeochemical Cycles 26, (2012). GB4007 Google Scholar
Day, L.T., Oswald, W.W., Doughty, E.D., and Foster, D.R. Analysis of hemlock pollen size in Holocene lake sediments from New England. Quaternary Research 79, (2013). 362365.Google Scholar
Dean, W.E. Determination of carbonate and organic matter in calcareous sediments and sedimentary rocks by loss on ignition: comparison with other methods. Journal of Sedimentary Petrology 44, (1974). 242248.Google Scholar
Dearing, J.A. Magnetic susceptibility. Walden, J., Oldfield, F., and Smith, J. Environmental Magnetism: a Practical Guide. Technical Guide No. 6. (1999). Quaternary Research Association, London. 3562.Google Scholar
Dorner, B. Natural landscape pattern and range of natural variability in the Arrow TSA. Arrow Innovative Forest Practices Agreement (IFPA). Final project report. (2001). (Internal document) Google Scholar
Enache, M.D., and Cumming, B.F. Tracking recorded fires using charcoal morphology from the sedimentary sequence of Prosser Lake, British Columbia (Canada). Quaternary Research 65, (2006). 282292.CrossRefGoogle Scholar
Enache, M.D., and Cumming, B.F. Charcoal morphotypes in lake sediments from British Columbia (Canada): an assessment of their utility for the reconstruction of past fire and precipitation. Journal of Paleolimnology 38, (2007). 347363.Google Scholar
Enache, M.D., and Cumming, B.F. Extreme fires under warmer and drier conditions inferred from sedimentary charcoal morphotypes from Opatcho Lake, central British Columbia, Canada. The Holocene 19, (2009). 835846.CrossRefGoogle Scholar
Ersek, V., Clark, P.U., Mix, A.C., Cheng, H., and Edwards, R.L. Holocene winter climate variability in mid-latitude western North America. Nature Communications 3, (2012). 1219 http://dx.doi.org/10.1038/ncomms2222Google Scholar
Fægri, K., and Iversen, J. Textbook of Pollen Analysis. (1989). John Wiley and Sons, New York.Google Scholar
Foit, F.F., Gavin, D.G., and Hu, F.S. The tephra stratigraphy of two lakes in south-central British Columbia, Canada and its implications for mid–late Holocene volcanic activity at Glacier Peak and Mount St. Helens, Washington, USA. Canadian Journal of Earth Sciences 41, (2004). 14011410.Google Scholar
Gavin, D.G., Brubaker, L.B., McLachlan, J.S., and Oswald, W.W. Correspondence of pollen assemblages with forest zones across steep environmental gradients, Olympic Peninsula, Washington, USA. The Holocene 15, (2005). 648662.Google Scholar
Gavin, D.G., Hu, F.S., Lertzmen, K., and Corbett, P. Weak climatic control of stand-scale fire history during the late Holocene. Ecology 87, (2006). 17221732.CrossRefGoogle ScholarPubMed
Gavin, D.G., Hallett, D.J., Hu, F.S., Lertzman, K.P., Prichard, S.J., Brown, K.J., Lynch, J.A., Bartlein, P.J., and Peterson, D.L. Forest fire and climate change in western North America: insights from sediment charcoal records. Frontiers in Ecology and the Environment 5, (2007). 499506.Google Scholar
Gavin, D.G., Henderson, A.C.G., Westover, K.S., Fritz, S.C., Walker, I.R., Leng, M.J., and Hu, F.S. Abrupt Holocene climate change and potential response to solar forcing in western Canada. Quaternary Science Reviews 30, (2011). 12431255.Google Scholar
Gedalof, Z. Climate and spatial patterns of wildfire. McKenzie, D., Miller, C., and Falk, D. The Landscape Ecology of Fire. (2011). Springer, 89116.Google Scholar
Gedalof, Z., Peterson, D.L., and Mantua, N.J. Columbia River flow and drought since 1750. Journal of the American Water Resources Association 40, (2004). 15791592.CrossRefGoogle Scholar
Gedalof, Z., Peterson, D.L., and Mantua, N.J. Atmospheric, climatic and ecological controls on extreme wildfire years in the northwestern United States. Ecological Applications 15, (2005). 154174.CrossRefGoogle Scholar
Gedye, S.J., Jones, R.T., Tinner, W., Ammann, B., and Oldfield, F. The use of mineral magnetism in the reconstruction of fire history: a case study from Lago di Origlio, Swiss Alps. Palaeogeography, Palaeoclimatology, Palaeoecology 164, (2000). 101110.CrossRefGoogle Scholar
Glew, J.R. A portable extruding device for close interval sectioning of unconsolidated core samples. Journal of Paleolimnology 1, (1988). 235239.CrossRefGoogle Scholar
Glew, J.R., Smol, J.P., and Last, W.M. Sediment core collection and extrusion. Last, W.M., and Smol, J.P. Tracking Environmental Change Using Lake Sediments. Basin Analysis, Coring, and Chronological Techniques 1, (2001). Kluwer Academic Publishers, Dordrecht. 73105.Google Scholar
Greene, G.A. Historical fire regime of the Darkwoods: quantifying the past to plan for the future. (MSc Thesis) (2011). Department of Geography, University of British Columbia, Vancouver.Google Scholar
Hallett, D.J., and Anderson, R.S. Paleofire reconstruction for high-elevation forests in the Sierra Nevada, California, with implications for wildfire synchrony and climate variability in the late Holocene. Quaternary Research 73, (2010). 180190.Google Scholar
Hallett, D.J., and Hills, L.V. Holocene vegetation dynamics, fire history, lake level and climate change in the Kootenay Valley, southeastern British Columbia, Canada. Journal of Paleolimnology 35, (2006). 351357.CrossRefGoogle Scholar
Hansen, B.C.S. Conifer stomata analysis as a paleoecological tool: an example from the Hudson Bay Lowlands. Canadian Journal of Botany 73, (1995). 244252.CrossRefGoogle Scholar
Hebda, R.J. British Columbia vegetation and climate history with focus on 6 ka BP. Géographie Physique et Quaternaire 49, (1995). 5579.Google Scholar
Heiri, O., Lotter, A.F., and Lemcke, G. Loss on ignition as a method for estimating organic and carbonate content in sediments: reproducibility and comparability of results. Journal of Paleolimnology 25, (2001). 101110.CrossRefGoogle Scholar
Hessl, A.E. Pathways for climate change effects on fire: models, data, and uncertainties. Progress in Physical Geography 35, (2011). 393407.Google Scholar
Heyerdahl, E.K., Brubaker, L.B., and Agee, J.K. Spatial controls of historical fire regimes: a multiscale example from the interior west, USA. Ecology 82, (2001). 660678.Google Scholar
Heyerdahl, E.K., Brubaker, L.B., and Agee, J.K. Annual and decadal climate forcing of historical fire regimes in the interior Pacific Northwest, USA. Holocene 12, (2002). 597604.Google Scholar
Heyerdahl, E.K., Lertzman, K., and Karpuk, S. Local-scale controls of a low-severity fire regime (1750–1950), southern British Columbia, Canada. Ecoscience 14, (2007). 4047.Google Scholar
Higuera, P.E. CharAnalysis 0.9: diagnostic and analytical tools for sediment-charcoal analysis [User's Guide]. Available from http://CharAnalysis.googlepages.com (2009). Google Scholar
Higuera, P.E., Peters, M.E., Brubaker, L.B., and Gavin, D.G. Understanding the origin and analysis of sediment-charcoal records with a simulation model. Quaternary Science Reviews 26, (2007). 17901809.Google Scholar
Higuera, P.E., Brubaker, L.B., Anderson, P.M., Hu, F.S., and Brown, T.A. Vegetation mediated the impacts of postglacial climate change on fire regimes in the south-central Brooks Range, Alaska. Ecological Monographs 79, (2009). 201219.Google Scholar
Higuera, P.E., Gavin, D.G., Bartlein, P.J., and Hallett, D.J. Peak detection in sediment-charcoal records: impacts of alternative data analysis methods on fire-history interpretations. International Journal of Wildland Fire 19, (2010). 9961014.Google Scholar
Higuera, P.E., Whitlock, C., and Gage, J.A. Linking tree-ring and sediment-charcoal records to reconstruct fire occurrence and area burned in subalpine forests of Yellowstone National Park, USA. Holocene 21, (2010). 327341.Google Scholar
Hu, F.S., Brubaker, L.B., Gavin, D.G., Higuera, P.E., Lynch, J.A., Rupp, T.S., and Tinner, W. How climate and vegetation influence the fire regime of the Alaskan Boreal biome: The Holocene perspective. Mitigation and Adaptation Strategies for Global Change 11, (2006). 829846.Google Scholar
Johnson, E.A., and Wowchuk, D.R. Wildfires in the southern Canadian Rocky Mountains and their relationship to mid-tropospheric anomalies. Canadian Journal of Forest Research 23, (1993). 12131222.Google Scholar
Johnson, E.A., Fryer, G.I., and Heathcott, M.J. The influence of man and climate on frequency of fire in the interior wet belt forest, British Columbia. Journal of Ecology 78, (1990). 403412.Google Scholar
Journeay, J.M., Williams, S.P., Wheeler, J.O., (2000). Tectonic assemblage map, Kootenay Lake, British Columbia–Alberta–USA. Geological Survey of Canada, Open File 2948b. 1:1,000,000.Google Scholar
Jungen, J.R. Soil resources of the Nelson map area (82F). RAB Bulletin 20, Report No. 28 British Columbia Soil Survey. (1980). Province of British Columbia, Ministry of Environment, Resource Analysis Branch. Queen's Printer for BC, Victoria. 217 Google Scholar
Kelly, R.F., Higuera, P.E., Barrett, C.M., and Hu, F.S. A signal-to-noise-index to quantify the potential for peak detections in sediment-charcoal records. Quaternary Research 75, (2011). 1117.Google Scholar
Kendrew, W.G., and Kerr, D. The Climate of British Columbia and the Yukon Territory. (1955). Queen's Printer, Ottawa. 222 Google Scholar
Kitzberger, T., Brown, P.M., Heyerdahl, E.K., Swetnam, T.W., and Veblen, T.T. Contingent Pacific–Atlantic Ocean influence on multicentury wildfire synchrony over western North America. Proceedings of the National Academy of Sciences of the United States of America 104, (2007). 543548.Google Scholar
Long, C.J., Whitlock, C., Bartlein, P.J., and Millspaugh, S.H. A 9000-year fire history from the Oregon Coast Range, based on a high-resolution charcoal study. Canadian Journal of Forest Research 28, (1998). 774787.Google Scholar
Lowe, D.J. Tephrochronology and its application: a review. Quaternary Geochronology 6, (2008). 107153.CrossRefGoogle Scholar
Luckman, B.H., Holdsworth, G., and Osborn, G.D. Neoglacial glacier fluctuations in the Canadian Rockies. Quaternary Research 39, (1993). 144153.Google Scholar
MacDonald, G.M. Conifer stomata. Smol, J.P., Birks, H.J.B., Last, W.M., Bradley, R.S., and Alverson, K. Tracking environmental change using lake sediments. Terrestrial, Algal, and Siliceous Indicators 3, (2001). Kluwer Academic Publishers, Dordrecht. 3348.Google Scholar
Marlon, J., Bartlein, P.J., and Whitlock, C. Fire–fuel–climate linkages in the northwestern U.S. during the Holocene. Holocene 16, (2006). 10591071.Google Scholar
Marlon, J.R., Bartlein, P.J., Gavin, D.G., Long, C.J., Anderson, R.S., Brilese, C.E., Brown, K.J., Colombaroli, D., Hallett, D.J., Power, M.J., Scharf, E.A., and Walsh, M.K. Long-term perspective on wildfires in the western USA. Proceedings of the National Academy of Sciences of the United States of America 109, (2012). E535E543.Google ScholarPubMed
Marlon, J.R., Bartlein, P.J., Daniau, A.-L., Harrison, S.P., Maezumi, S.Y., Power, M.J., Tinner, W., and Vannière, B. Global biomass burning: a synthesis and review of Holocene paleofire records and their controls. Quaternary Science Reviews 65, (2013). 525.Google Scholar
McAndrews, J.H., Berti, A.A., and Norris, G. Key to the Quaternary pollen and spores of the Great Lakes region. (1973). University of Toronto Press, Toronto.Google Scholar
Menounos, B., Osborn, G.D., Clague, J.J., and Luckman, B.H. Latest Pleistocene and Holocene glacier fluctuations in western Canada. Quaternary Science Reviews 28, (2009). 20492074.Google Scholar
Moore, P.D., and Webb, J.A. An Illustrated Guide to Pollen Analysis. (1978). Hodder and Stroughton, London.Google Scholar
Moos, M.T., and Cumming, B.F. Climate–fire interactions during the Holocene: a test of the utility of charcoal morphotypes in a sediment core from the boreal region of north-western Ontario (Canada). International Journal of Wildland Fire 21, (2012). 640652.CrossRefGoogle Scholar
Morris, J.L., Brunelle, A., DeRose, R.J., Seppä, H., Power, M.J., Carter, V., and Bares, R. Using fire regimes to delineate zones in a high-resolution lake sediment record from the western United States. Quaternary Research 79, (2013). 2436.Google Scholar
Mullineaux, D.R. Pre-1980 tephra-fall deposits erupted from Mount St. Helens, Washington, U.S. Geological Survey Professional Paper 1563, (1996). 199.Google Scholar
Nelson, D.B., Abbott, M.B., Steinman, B., Polissar, P.J., Stansell, N.D., Ortiz, J.D., Rosenmeier, M.F., Finney, B.P., and Riedel, J. Drought variability in the Pacific Northwest from a 6,000-yr lake sediment record. Proceedings of the National Academy of Sciences (2011). http://dx.doi.org/10.1073/pnas.1009194108Google Scholar
Nesbitt, J.H. Quantifying forest fire variability using tree rings Nelson, British Columbia 1700–Present. (MSc. Thesis) (2010). University of British Columbia, Vancouver.Google Scholar
Oke, T.R. Boundary Layer Climates. (1987). Methuen, London. 435 Google Scholar
Pollack, J., Quesnel, H., Hauk, C., and MacLean, H. A quantitative evaluation of natural age class distributions and stand replacement intervals in the Nelson Forest Region. Forest Sciences, Nelson Forest Region, B.C. Ministry of Forests, Technical Report TR-015. (1997). Google Scholar
Reimer, P.J., Baillie, M.G.L., Bard, E., Bayliss, A., Beck, J.W., Blackwell, P.G., Bronk Ramsey, C., Buck, C.E., Burr, G.S., Edwards, R.L. et al. IntCal09 and Marine09 radiocarbon age calibration curves, 0–50,000 years cal BP. Radiocarbon 51, (2009). 11111150.Google Scholar
Reyes, A.V., and Clague, J.J. Stratigraphic evidence for multiple Holocene advances of Lillooet Glacier, southern Coast Mountains, British Columbia. Canadian Journal of Earth Sciences 41, (2004). 903918.Google Scholar
Rodionov, S.N. A sequential algorithm for testing climate regime shifts. Geophysical Research Letters 31, (2004). L09204 Google Scholar
Rosenberg, S.M., Walker, I.R., and Mathewes, R.W. Postglacial spread of hemlock (Tsuga) and vegetation history in Mount Revelstoke National Park, British Columbia, Canada. Canadian Journal of Botany 81, (2003). 139151.Google Scholar
Rosenberg, S.M., Walker, I.R., Mathewes, R.W., and Hallett, D.J. Midge-inferred Holocene climate history of two subalpine lakes in southern British Columbia, Canada. Holocene 14, (2004). 258271.CrossRefGoogle Scholar
Scott, A.C., Cripps, J.A., Collinson, M.E., and Nichols, G.J. The taphonomy of charcoal following a recent heathland fire and some implications for the interpretation of fossil charcoal deposits. Palaeogeography, Palaeoclimatology, Palaeoecology 164, (2000). 131.Google Scholar
Steinman, B.A., Abbott, M.B., Mann, M.E., Stansell, N.D., and Finney, B.P. 1500-year quantitative reconstruction of precipitation in the Pacific Northwest. Proceedings of the National Academy of Sciences 109, (2012). 1161911623.Google Scholar
Steventon, J.D. Historic disturbance rates for interior biogeoclimatic subzones of the Prince Rupert Forest Region. Extension Note No. 26. (1997). Forest Sciences, Prince Rupert Forest Region, BC Forest Service, Smithers, BC, Google Scholar
Stockmarr, R.J. Tablets with spores used in absolute pollen analysis. Pollen et Spores 13, (1971). 615621.Google Scholar
Sweeney, C.A. A key for the identification of stomata of the native conifers of Scandinavia. Review of Palaeobotany and Palynology 128, (2004). 281290.CrossRefGoogle Scholar
Syvitski, J.P.M. Principles, methods, and Application of Particle Size Analysis. (1991). Press Syndicate of the Cambridge University Press, New York. 368 Google Scholar
Thevenon, F., and Anselmetti, F.S. Charcoal and fly-ash particles from Lake Lucerne sediments (Central Switzerland) characterized by image analysis: anthropologic, stratigraphic and environmental implications. Quaternary Science Reviews 26, (2007). 26312643.Google Scholar
Trouet, V., Taylor, A.H., Wahl, E.R., Skinner, C.N., and Stephens, S.L. Fire–climate interactions in the American West since 1400 CE. Geophysical Research Letters 37, (2010). L04702 Google Scholar
Wang, T., Hamann, A., Spittlehouse, D., and Murdock, T.N. Climate WNA—high-resolution spatial climate data for Western North America. Journal of Applied Meteorology and Climatology 61, (2012). 1629.Google Scholar
Westerling, A.L., Hidalgo, H.G., Cayan, D.R., and Swetnam, T.W. Warming and earlier spring increase western US forest wildfire activity. Science 313, (2006). 940943.Google Scholar
Whitlock, C., and Larsen, C.P.S. Charcoal as a fire proxy. Smol, J.P., Birks, H.J.B., Last, W.M., Bradley, R.S., and Alverson, K. Tracking environmental change using lake sediments. Terrestrial, Algal, and Siliceous Indicators 3, (2001). Kluwer Academic Publishers, Dordrecht. 7597.Google Scholar
Whitlock, C., Higuera, P.E., McWethy, D.B., and Briles, C.E. Paleoecological perspectives on fire ecology: revisiting the fire-regime concept. The Open Ecology Journal 3, (2010). 623.Google Scholar
Wilson, S.E., Cumming, B.F., and Smol, J.P. Diatom–salinity relationships in 111 lakes from the Interior Plateau of British Columbia, Canada: the development of diatom-based models for paleosalinity and paleoclimatic reconstructions. Journal of Paleolimnology 12, (1994). 197221.Google Scholar
Wong, C.M., Sandmann, H., and Dorner, B. Historical Variability of Natural Disturbances in British Columbia: a Literature Review. (2004). FORREX-Forest Research Extension Partnership, Kamloops, British Columbia, Canada.Google Scholar
Wright, H.E.J., Mann, D.H., and Glaser, P.H. Piston corers for peat and lake sediments. Ecology 65, (1984). 657659.Google Scholar
Yamaguchi, D.K. New tree ring dates for recent eruptions of Mount St. Helens. Quaternary Research 20, (1983). 246250.Google Scholar
Yamaguchi, D.K. Tree-ring evidence for a two-year interval between recent prehistoric explosive eruptions of Mount St. Helens. Geology 13, (1985). 554557.Google Scholar
Zdanowicz, C.M., Zielinski, G.A., and Germani, M.S. Mount Mazama eruption: calendrical age verified and atmospheric impact assessed. Geology 27, (1999). 621624.Google Scholar
Supplementary material: File

Courtney Mustaphi and Pisaric supplementary material

Supplementary Material

Download Courtney Mustaphi and Pisaric supplementary material(File)
File 1.5 KB