Hostname: page-component-848d4c4894-nmvwc Total loading time: 0 Render date: 2024-06-28T15:50:30.516Z Has data issue: false hasContentIssue false

Tracking recorded fires using charcoal morphology from the sedimentary sequence of Prosser Lake, British Columbia (Canada)

Published online by Cambridge University Press:  20 January 2017

Mihaela D. Enache*
Affiliation:
Paleoecological Environmental Assessment and Research Laboratory (PEARL), Department of Biology, Queen's University, Kingston, ON, Canada K7L 3N6
Brian F. Cumming
Affiliation:
Paleoecological Environmental Assessment and Research Laboratory (PEARL), Department of Biology, Queen's University, Kingston, ON, Canada K7L 3N6
*
*Corresponding author. Fax: +1 613 533 6617.E-mail address:enachem@biology.queensu.ca(M.D. Enache).

Abstract

Quantitative analyses of variations in morphological features of charcoal were undertaken in a 210Pb-dated sediment core from Prosser Lake (British Columbia, Canada). Seven morphological types of charcoal were defined by particle shape, major–minor axis ratio, apparent porosity and progradation to unburned material. The distribution of morphotypes and total charcoal abundances were assessed as a proxy for fire events recorded between 1919 and 2000 and to subsequent mechanisms of transportation–sedimentation to lake sediments. Charcoal morphotypes showed distinct relationships to recorded area burned by fires. Fragile charcoal fragments with highly irregular porosity (termed Type M) displayed the strongest correlation to burned area (r2 = 0.51; P = 0.0001) and did not produce any false-positive signal for fires recorded within a radius of 20 km around the lake. We infer that high porosity and low density Type M fragments are aerially transported and directly deposited on the lake, and that the fragility of Type M charcoal prevents significant quantities from being secondarily transported and incorporated into the sedimentary record. We propose that charcoal morphology is an important but underutilized technique that can yield important insights into fire type, proximity and transportation–sedimentation processes.

Type
Original Articles
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

Agee, J., (1993). Fire Ecology of Pacific Northwest Forests. Island Press, Washington, D.C..Google Scholar
Bailey, J.G., Tate, A., Diesel, C.F.K., Wall, T.F., (1990). A char morphology system with applications to coal combustion. Fuel 69, 225233.Google Scholar
Valentine, K.W.G., Sprout, P.N., Baker, T.E., Lavkulich, L.M., (1994). The Soil Landscapes of British Columbia.Victoria B.C. Ministry of EnvironmentGoogle Scholar
Bégin, Y., Marguerie, D., (2002). Characterization of tree macroremains production in a recently burned conifer forest in northern Quebec, Canada. Plant Ecology 159, 143152.Google Scholar
Bengtsson, M., (1987). Combustion behaviour for a coal containing a high proportion of pseudovitrinite. Fuel Processing Technology 15, 201212.Google Scholar
Binford, M.W., (1990). Calculation and uncertainty analysis of 210Pb dates for PIRLA project lake sediment cores. Journal of Paleolimnology 3, 253267.CrossRefGoogle Scholar
Bryden, M., Hagge, M.J., (2003). Modelling the combined impact of moisture and char shrinkage on the pyrolysis of a biomass particle. Fuel 82, 16331644.Google Scholar
Clark, J.S., (1988). Stratigraphic charcoal analysis on petrographic thin sections: application to fire history in northwestern Minnesota. Quaternary Research 30, 8191.CrossRefGoogle Scholar
Clark, J.S., (1989). Ecological disturbance as a renewal process: theory and application to fire history. Oikos 56, 1730.CrossRefGoogle Scholar
Clark, J.S., Royall, P.D., (1995). Particle size evidence for source areas of charcoal accumulation rates in Late Holocene Sediments of Eastern North American Lakes. Quaternary Research 43, 8089.Google Scholar
Clark, J.S., Lynch, J., Stocks, B.J., Goldammer, J.G., (1998). Relationships between charcoal particles in air and sediments in West-Central Siberia. The Holocene 8, 1930.Google Scholar
Cloke, M., Lester, E., (1994). Characterization of coals for combustion using petrographic analysis: a review. Fuel 73, 315320.Google Scholar
Doubleday, N.C., Smol, J.P., (2005). Atlas and classification scheme of arctic combustion particles suitable for paleoenvironmental work. Journal of Paleolimnology 33, 393431.Google Scholar
Eakins, J.D., Morrison, R.T., (1978). A new procedure for the determination of lead-210 in lake and marine sediments. International Journal of Applied Radiation and Isotopes 29, 336531.Google Scholar
Farley, A.L., (1979). Atlas of British Columbia. The University of British Columbia Press, .Google Scholar
Fearnside, P.M., Graça, P.M.L.A., Filho, N.H., Rodrigues, F.J.A., Robinson, J.M., (1999). Tropical forest burning in Brazilian Amazonia: measurement of biomass loading, burning efficiency and charcoal formation at Altamira, Pará. Forest Ecology and Management 123, 6579.Google Scholar
Fearnside, P.M., Graça, P.M.L.A., Rodrigues, F.J.A., (2001). Burning of Amazonian rainforests: burning efficiency and charcoal formation in forest leared for cattle pasture near Manaus, Brazil. Forest Ecology and Management 146, 115128.Google Scholar
Gardner, J., Whitlock, C., (2001). Charcoal accumulation following a recent fire in the Cascade Range, northwestern USA, and its relevance for fire-history studies. The Holocene 11, 541549.CrossRefGoogle Scholar
Garstang, M., Tysson, P.D., (1996). Atmospheric circulation, vertical structure and transport. van Wilgen, B.W., Andreae, M.O., Goldammer, J.G., Lindesay, J.A., Fire in Southern African Savannas: Ecological and Atmospheric Perspectives Witwatersrand University Press, Johannesburg, South Africa.Google Scholar
Garstang, M., Tyson, P.D., Cachier, H., Radke, L., (1997). Atmospheric transport of particulate and gaseous products by fires. Clark, J.S., Cachier, H., Goldammer, J.G., Stocks, B., Sediment Records of Biomass Burning and Global Change 207250.CrossRefGoogle Scholar
Glew, J., (1991). Miniature gravity corer for recovering short sediment cores. Journal of Paleolimnology 5, 285287.CrossRefGoogle Scholar
Graça, P.M.L.A., Fearnside, P.M., Cerri, C.C., (1999). Burning of Amazonian forest in Ariquemes, Rondônia, Brazil: biomass, charcoal formation, and burning efficiency. Forest Ecology Management 120, 179191.Google Scholar
Grotkjær, T., Dam-Johansen, K., Jensen, A.D., Glarborg, P., (2003). An experimental study of biomass ignition. Fuel 82, 825833.CrossRefGoogle Scholar
Jones, T.P., Chaloner, W.G., Kuhlbusch, T.A.J., (1997). Proposed bio-geological and chemical based terminology in fire-altered plant matter. Clark, J.S., Cachier, H., Goldammer, J.G., Stocks, B., Sediment Records of Biomass Burning and Global Change 922.CrossRefGoogle Scholar
Kurosaki, F., Ishimaru, K., Hata, T., Bronsveld, P., Kobayashi, E., Imamura, Y., (2003). Microstructure of wood charcoal by flash heating. Carbon 41, 30573062.Google Scholar
Lester, E., Cloke, M., Allen, M., (1996). Char characterization using image analysis techniques. Energy and Fuels 10, 696703.Google Scholar
Long, C.J., Whitlock, C., (2002). Fire and vegetation history from the Coastal Rain Forest of the Western Oregon Coast Range. Quaternary Research 58, 215225.Google Scholar
Lynch, J.A., Clark, J.S., Stocks, B.J., (2004). Charcoal production, dispersal, and deposition from the Fort Providence experimental fire: interpreting fire regimes from charcoal records in boreal forests. Canadian Journal of Forest Research 34, 16421656.CrossRefGoogle Scholar
Millspaugh, S.H., Whitlock, C., (1995). A 750-year fire history based on lake sediment records in central Yellowstone National Park. The Holocene 5, 283292.Google Scholar
Mlaouhi, A., Khouaja, A., Saoudi, H., Depeyre, D., (1999). Trials of wood carbonization of some forest and fruit-bearing species. Renewable Energy 16, 11181121.Google Scholar
Nichols, G.J., Cripps, J.A., Collinson, M.E., Scott, A.C., (2000). Experiments in waterlogging and sedimentology of charcoal: results and implications. Palaeogeography, Palaeoclimatology, Palaeoecology 164, 4356.Google Scholar
Oka, N., Murayama, T., Yamada, S., Yamada, T., Shinozaki, S., Shibaoka, M., Thomas, C.G., (1987). The influence of rank and maceral composition and char burnout of pulverized coal. Fuel Processing Technology 15, 213224.CrossRefGoogle Scholar
Órfão, J.J.M., Antunes, F.J.A., Figueiredo, J.L., (1999). Pyrolysis kinetics of lignocellulosic materials—Three independent reactions model. Fuel 78, 349358.CrossRefGoogle Scholar
Patterson, W.A., Edwards, K.J., Maguire, D.J., (1987). Microscopic charcoal as an indicator of fire. Quaternary Science Reviews 6, 323.CrossRefGoogle Scholar
Pisaric, M.F.J., (2002). Long distance transport of terrestrial plant material by convection resulting from forest fires. Journal of Paleolimnology 28, 349354.Google Scholar
Scott, A.C., Cripps, J.A., Collinson, M.E., Nichols, G.J., (2000). The taphonomy of charcoal following a recent heatland fire and some implications for the interpretation of fossil charcoal deposits. Palaeogeography, Palaeoclimatology, Palaeoecology 164, 131.Google Scholar
Séro-Guillaume, O., Margerit, J., (2002). Modelling forest fires. Part I: a complete set of equations derived by extended irreversible thermodynamics. International Journal of Heat Mass Transfer 45, 17051722.Google Scholar
Umbanhowar, C.E. Jr. McGrath, M.J., (1998). Experimental production and analysis of microscopic charcoal from wood, leaves, and grasses. The Holocene 8, 341346.CrossRefGoogle Scholar
Vaughan, A., Nichols, G., (1995). Controls on the deposition of charcoal: implications for sedimentary accumulations of fusain. Journal of Sedimentary Research 65 02, 129135.Google Scholar
Vernet, J.-L., Ogereau, P., Figueiral, I., Machado Yanes, C., Uzquiano, P., (2001). Guide d'identification des charbons de bois préhistoriques et récents.. Sud-Ouest de l'Europe: France, Péninsule Ibérique et Îles Canaries.Google Scholar
Ward, D., (2001). Combustion chemistry and smoke. Johnson, E.A., Miyanishi, K., Forest Fires: Behavior and Ecological Effects Academic Press, San Diego.5577.CrossRefGoogle Scholar
Whitlock, C., Larsen, C., (2001). Charcoal as a fire proxy. Smol, John P., Birks, H. John B., Last, William M., Tracking Environmental Change Using Lake Sediments Kluwer Academic Publishers, 7598.Google Scholar
Whitlock, C., Millspaugh, S.H., (1996). Testing the assumptions of fire-history studies: an examination of modern charcoal accumulation in Yellowstone National Park, USA. The Holocene 6, 715.Google Scholar
Whitlock, C., Shafer, S.L., Marlon, J., (2003). The role of climate and vegetation change in shaping past and future fire regimes in the northwestern US and the implications for ecosystem management. Forest Ecology and Management 178, 521.Google Scholar
Zunckel, M., Hong, Y., Brassel, K., O'Beirne, S., (1996). Characteristics of the nocturnal boundary layer: Okaukuejo, Namibia during SAFARI-92. J. Geophys. Res. 101, 2375723766.Google Scholar