Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-28T08:05:21.925Z Has data issue: false hasContentIssue false

Particle-Size Evidence for Source Areas of Charcoal Accumulation in Late Holocene Sediments of Eastern North American Lakes

Published online by Cambridge University Press:  20 January 2017

James S. Clark
Affiliation:
Department of Botany, Duke University, Durham, North Carolina 27706
P.Dan Royall
Affiliation:
Department of Geography, University of Tennessee, Knoxville, Tennessee 37996

Abstract

Two methods of analyzing charcoal in sediment reveal changes in charcoal accumulation across temperate eastern North America during the last several hundred years. In one method the analyst counts mostly small particles that reflect regional emissions; in the other, the analyst counts only larger particles derived mostly from such local sources as catchment fires. We used these methods to compare charcoal accumulation at 14 lakes from the prairie/forest border in Minnesota to eastern Maine. The two methods gave concordant accumulation rates for sediments of pre-1850 age at each of 4 lakes analyzed by both methods. This concordance is consistent with the interpretation that pre-1850 emissions were controlled by broad-scale factors, such as climatically controlled regional differences in fuels and moisture. Since 1900 large particles decreased greatly, and small particles decreased slightly, in Minnesota and Wisconsin. By contrast in the Northeast the large particle accumulation has remained at the low values measured in pre-1900 sediments at most sites, while small particles increased everywhere east of central New York and Pennsylvania. The observed patterns suggest that (1) large particles primarily reflect local fires that were common in the Midwest before fire suppression became effective, (2) large particles were rare in the Northeast, especially before extensive land clearance, (3) small particles reflect regional combustion that increased in the Northeast after extensive use of fire for land clearance and wood burning for industrial purposes of the 19th century, and (4) small particles remain abundant in the Midwest long after effective fire suppression, probably because these well-dispersed small particles have a large source area that extends beyond the local wildfires that account for large particles before European settlement.

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

Andreae, M. O. (1991). Biomass burning: Its history, use, and distribution and its impact on environmental quality and global climate. In“Global Biomass Burning: Atmospheric, Climatic, and Biospheric Implications” (Levin, J., Ed.), pp. 321. MIT Press, Cambridge.CrossRefGoogle Scholar
Bonny, A. P. (1976). Recruitment of pollen to the seston and sediment of some Lake District Lakes. Journal of Ecology 64, 859888.CrossRefGoogle Scholar
Bradbury, J. P. (1986). Effects of forest fire and other disturbances on wilderness lakes in northeastern Minnesota. II. Paleolimnology. Archiv fur Hydrobiologie 106, 203217.CrossRefGoogle Scholar
Chylek, P. Ramaswamy, V., and Srivastava, V. (1984). Graphitic carbon content of aerosols, clouds and snow, and its climatic implications. Science of the Total Environment 36, 117120.CrossRefGoogle Scholar
Clark, J. S. (1988a). Particle motion and the theory of charcoal analysis: source area, transport, deposition, and sampling. Quaternary Re-search 30, 8191.Google Scholar
Clark, J. S. (1988b). Charcoal-stratigraphic analysis on petrographic thin sections: Recent fire history in northwest Minnesota. Quaternary Research 30, 6780.CrossRefGoogle Scholar
Clark, J. S. (1990). Fire and climate change during the last 750 years in northwestern Minnesota. Ecological Monographs 60, 135159.CrossRefGoogle Scholar
Clark, J. S., and Robinson, J. (1993). Paleoecology of fire. In “Fire in the Environment: Its Ecological, Climatic, and Atmospheric Chem-ical Importance” (Crutzen, P. J. and Goldammer, J. G., Eds.), pp. 193214. Wiley, New York.Google Scholar
Clark, J. S., and Royall, P. D. (1994a). Pre-industrial particulate emissions and carbon sequestration from biomass burning in North America. Biogeochemistry 23, 117.Google Scholar
Clark, J. S., and Royall, P. D. (1994b). Transformation of a northern hardwood forest by aboriginal fire. The Holocene, in press.Google Scholar
Cope, M. J., and Chaloner, W. G. (1985). Wildfire: An interaction of biological and physical processes. In “Geological Factors and the Evolution of Plants” (Tiffncy, B. H., Ed.), pp. 257277. Yale Univ. Press, New Haven.Google Scholar
Crutzen, P. J., and Goldammer, J. G. (1993). “Fire in the Environment: Its Ecological, Climatic, and Atmospheric Chemical Importance.” Wiley, New York.Google Scholar
Gajewski, K. Winkler, M. G., and Swain, A. M. (1985a). Vegetation and fire history from three lakes with varved sediments in northwestern Wisconsin. Review of Palaeobotany andPalynology 44, 277292.CrossRefGoogle Scholar
Gajewski, K. Swain, A. M. Winkler, M. G. Peterson, G. M., and Steventon, R. (1985b). Late Holocene pollen data from lakes with varved sediments in northeastern and northcentral United States. IES Report 124, Center for Climatic Research, Madison, WI.Google Scholar
Goldberg, E. D. (1985). “Black Carbon in the Environment.” Wiley, New York.Google Scholar
Griffen, J. J., and Goldberg, E. D. (1975). The fluxes of elemental carbon in coastal marine sediments. Limnology and Oceanography 20, 456463.CrossRefGoogle Scholar
Grimm, E. C. (1983). Chronology and dynamics of vegetation change in the prairie-woodland region of southern Minnesota, USA. New Phytologist 9 3, 311350.CrossRefGoogle Scholar
Haines, D. A. Johnson, V. J., and Main, W. A. (1975). Wildfire atlas of the Northeastern and North Central states. United States Department of Agriculture Forest Service General Technical Report NC-16.Google Scholar
Harris, T, M. (1958). Forest fire in the Mesozoic. Journal of Ecology 46, 447453.CrossRefGoogle Scholar
Heinselman, M. L. (1973). Fire in the virgin forest of the Boundary Waters Canoe Area, Minnesota. Quaternary Research 3, 329382.CrossRefGoogle Scholar
Herring, J, R. (1985). Charcoal fluxes into sediments of the North Pacific ocean: The Cenozoic record of burning. In “The Carbon Cycle and Atmospheric C02: Natural Variations, Archaen to Present” (Sundquist, E. T. and Broecker, W. S., Eds.), pp. 419442. Geophysical Monograph 32, American Geophysical Union, Washington, DC.Google Scholar
MacDonald, G. M. Larsen, C. P. S., and Moser, K. A. (1991). The reconstruction of boreal forest fire history from lake sediments: A comparison of charcoal, pollen sedimentological, and geochemical indices. Quaternary Science Reviews 10, 5371.CrossRefGoogle Scholar
Malingreau, J.-P. Albini, F. A. Andreae, M. O. Brown, S. Levine, J. S. Lobert, J. Kulbusch, T. A. Radke, L. Setzer, A. Vitousek, P. M. Ward, D. E., and Warnatz, J. (1993). Quantification of fire characteristics from local to global scales. In “Fire in the Environment: Its Ecological, Climatic, and Atmospheric Chemical Importance” (Crutzen, P. J. and Goldammer, J. G., Eds.). Wiley, New York.Google Scholar
Marziani, G. P. L., and Iannone, A. (1986). A new method for cutting thin sections from prehistoric charcoal specimen. Review of Paiaeobotany and Palynology 48, 295301.CrossRefGoogle Scholar
McAndrews, J. H., and Boyko-Diakonow, M. (1989). Pollen analysis of varved sediment at Crawford Lake Ontario: Evidence of Indian and European farming. In “Quaternary Geology of Canada and Greenland” (Fulton, R. J., Ed.), pp. 528530. Geological Survey of Canada.Google Scholar
Patterson, W. A., and Backman, A. E. (1988). Fire and disease history of forest. In “Vegetation History” (Huntley, B. and Webb, T. III, Eds.), pp. 603632. Kluwer, Dordrecht,CrossRefGoogle Scholar
Patterson, W. A. Edwards, K. J., and Maguire, D. J. (1987). Microscopic charcoal as a fossil indicator of fire. Quaternary Science Reviews 6, 323.CrossRefGoogle Scholar
Patterson, W. A. III, and Sassaman, K. E. (1988). Indian fires in the prehistory of New England. In “Holocene Human Ecology in North-eastern North America.” (Nicholas, G. P., Ed.), pp. 107135. Plenum, New York.CrossRefGoogle Scholar
Peck, R. M. (1973). Pollen budget studies in a small Yorkshire catchment. In “Quaternary Plant Ecology” (Birks, H. J. B. and West, R. G., Eds.), pp. 4360. Blackwell, Oxford.Google Scholar
Penner, J. E. Ghan, S. J., and Walton, J. J. (1991). The role of biomass burning in the budget and cycle of carbonaceous soot aerosols and their climate impact. In “Global Biomass Burning: Atmospheric, Climatic, and Biospheric Implications” (Levine, J. S., Ed.), pp. 387393. MIT Press, Cambridge, MA.CrossRefGoogle Scholar
Radke, L. F. Hegg, D. A. Hobbs, Nance, P. Lyons, J. H. Laursen, K. K. Weiss, R. E. Riggan, P. J., and Ward, D. E. (1991). Particulate and trace gas emissions from large biomass fires in North America. In“Global Biomass Burning: Atmospheric, Climatic, and Biospheric Implications” (Levine, J. S., Ed.), pp. 209224, MIT Press, Cambridge, MA.CrossRefGoogle Scholar
Renberg, I., and Wik, M. (1985). Soot particle counting in recent lake sediments: An indirect dating method. Ecologicaf Bu//emis 37, 5357.Google Scholar
Robinson, J. M. (1989). Phanerozoic 02 variation, fire, and terrestrial ecology. Palaeogeography, Palaeoclimatology. and Palaeoecology 75, 223240.CrossRefGoogle Scholar
Rosen, H. A. Hansen, D. A., and Novakov, T. (1984). Role of graphitic carbon particles in radiative transfer in the Arctic haze. The Science of the Total Environment 36, 103110.CrossRefGoogle Scholar
Sandberg, D. V. Pierovich, J. M. Fox, D. G., and Ross, E. W. (1979). Effects of fire on air. United States Department of Agriculture Forest Service General Technical Report WO-9, Washington, DC.Google Scholar
Sander, P. M., and Gee, C. T. (1990). Fossil charcoal: Techniques and applications. Review of Palaeobotany and Palynology 63, 269279.CrossRefGoogle Scholar
Schweingruber, F. H. (1976). Prahistorisches Holz: Die Bedeutung von Holzfunden aus Mitteleuropa fur die Losung archaologischer und vegetationskundlicher Probleme. Acad. Helv. 2, 1106.Google Scholar
Seiler, W., and Crutzen, P. J. (1980). Estimates of gross and net fluxes of carbon between the biosphere and the atmosphere from biomass burning. Climatic Change 2, 207247.CrossRefGoogle Scholar
Shaw, G. E. (1982). Perturbation to the atmospheric radiation field from carbonaceous aerosols. In “Particulate Carbon: Atmospheric Life Cycle” (Wolff, G. T. and Klimisch, R. L., Eds.), pp. 5374. Plenum, New York.CrossRefGoogle Scholar
Swain, A. M. (1973). A history of fire and vegetation in fortheastem Minnesota as recorded in lake sediment. Quaternary Research 3, 383396.CrossRefGoogle Scholar
Tolonen, K. (1983). The post-glacial fire record. In “The Role of Fire in Northern Circumpolar Ecosystems” (Wein, R. W. and Mac-Lean, D. A., Eds.), pp. 2144. Wiley, New York.Google Scholar
Vemet, J.-L. Bazile, E., and Evin, J. (1979). Coordination des analyses anthracologiques et des datations absolues sur charbon debois. Bulletin de la Societe Prehistorique Franqaise 76, 7679.Google Scholar
Williams, M. (1989). “Americans and Their Forests: A Historical Geography.” Cambridge Univ. Press, Cambridge, England.Google Scholar
Winkler, M. J. (1985). Charcoal analysis for paleoenvironmental interpretation: A chemical assay. Quaternary Research 23, 313326.CrossRefGoogle Scholar