Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-05-31T21:06:55.965Z Has data issue: false hasContentIssue false

Non-linear connections between dune activity and climate in the High Plains, Kansas and Oklahoma, USA

Published online by Cambridge University Press:  20 January 2017

Corey M. Werner*
Affiliation:
Department of Geography, University of Central Missouri, Warrensburg, MO 64093, USA
Joseph A. Mason
Affiliation:
Department of Geography, University of Wisconsin-Madison, Madison, WI 53706, USA
Paul R. Hanson
Affiliation:
School of Natural Resources, University of Nebraska-Lincoln, Lincoln, NE 68583, USA
*
Corresponding author. Department of Geography and Interdisciplinary Studies, Wood 6b, University of Central Missouri, Warrensburg, MO 64093, USA. Fax: + 1 660 543 4048.

Abstract

Discrete dune fields are found throughout much of the Great Plains of North America, and the timing of past dune activity is often used as a proxy for paleoclimate because of the intuitive link between dune activity and a more arid climate. This research suggests that feedbacks in the soil-geomorphic system create a relationship between dune activity and climate that varies both spatially and temporally. Older eolian landforms are more resistant to activation because of the long-term accumulation of finer soil particles in a Bt horizon which retain moisture and anchor the deposit even during more arid times. Conversely, younger deposits lack these fines and are more easily reactivated. This spatially variable relationship is supported by soil stratigraphy, particle size analysis, and optical age control. Additionally, the water retention of the Bt horizons is quantifiably greater than that of the soils found in the younger dunes of the area. This complication in the relationship between eolian activity and climate is important because it suggests that caution is needed when using past dune activity as the lone proxy for paleoclimate.

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

Arbogast, A.F. Stratigraphic evidence for late-Holocene aeolian sand mobilization and soil formation in south-central Kansas, U.S.A.. Journal of Arid Environments 34, (1996). 403414.Google Scholar
Arbogast, A.F. Late Quaternary paleoenvironments and landscape evolution on the Great Bend Sand Prairie. Kansas Geological Survey Bulletin 242, (1998). Google Scholar
Arbogast, A.F., and Johnson, W.C. Late-Quaternary landscape response to environmental change in south-central Kansas. Annals of the Association of American Geographers 88, 1 (1998). 126145.Google Scholar
Arbogast, A.F., and Muhs, D.R. Geochemical and mineralogical evidence from eolian sediments for northwesterly mid-Holocene paleowinds, central Kansas USA. Quaternary International 67, 1 (2000). 107118.CrossRefGoogle Scholar
Beuselinck, L., Grovers, G., Poeson, J., Degraer, G., and Froyen, L. Grain-size analysis by laser diffractometry: comparison with the sieve-pipette method. Catena 32, (1998). 193208.Google Scholar
Birkeland, P.W. Soils and Geomorphology. (1999). Oxford University Press, New York.Google Scholar
Bullard, J.E., Mctainsh, G.H., and Pudmenzky, C. Factors affecting the nature and rate of dust production from natural dune sands. Sedimentology 54, (2007). 169182.Google Scholar
Byrd, T.C. Soil Survey for Stevens County, KS. (2006). National Resource Conservation Service, Google Scholar
Byrne, F.E., and Mclaughlin, T.G. Geology and Groundwater Resources of Seward County, Kansas. Kansas Geological Survey Bulletin 69, (1948). Google Scholar
Christensen, J.H., Hewitson, B., Busuioc, A., Chen, A., Gao, X., Held, I., Jones, R., Kolli, R.K., Kwon, W.-T., Laprise, R., Magana Rueda, V., Mearns, L., Menendez, C.G., Ralsanen, J., Rinke, A., Sarr, A., and Whetton, P. Regional climate projections. Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M., and Miller, H.L. Climate Change 2007: the Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. (2007). Cambridge University Press, Google Scholar
Cordova, C.E., Porter, J.C., Lepper, K., Kalchgruber, R., and Scott, G.F. Preliminary assessment of sand dune stability along a bioclimatic gradient, north-central and northwestern Oklahoma. Great Plains Research 15, (2005). 227249.Google Scholar
Forman, S.L., Marin, L., and Gomez, J. Eolian depositional records from southwestern Kansas and southeastern Colorado, a potential landscape response to droughts in the past 10,000 years. Abstracts with Programs — Geological Society of America 36, 5 (2004). 412 Google Scholar
Forman, S.L., Marin, L., Gomez, J., and Pierson, J. Late Quaternary eolian sand depositional record for southwestern Kansas: landscape sensitivity to droughts. Palaeogeography, Palaeoclimatology, Palaeoecology 265, (2008). 107120.Google Scholar
Forman, S.L., Oglesby, R., and Webb, R.S. Temporal and spatial patterns of Holocene dune activity on the Great Plains of North America: megadroughts and climate links. Global and Planetary Change 29, (2001). 129.CrossRefGoogle Scholar
Gesch, D.B. The national elevation dataset. Maune, D. Digital Elevation Model Technologies and Applications: The DEM Users Manual. 2nd ed (2007). American Society for Photogrammetry and Remote Sensing, Bethesda MD. 99118.Google Scholar
Gesch, D.B., Oimoen, M., Greenlee, S., Nelson, C., Steuck, M., and Tyler, D. The national elevation dataset. Photogrammetric Engineering and Remote Sensing 68, 1 (2002). 511.Google Scholar
Gile, L.H. Holocene soils in eolian sediments of Bailey County, Texas. Soil Science Society of America Journal 43, (1979). 9941003.Google Scholar
Hanson, P.R., Arbogast, A.F., Johnson, W.C., Joeckel, R.M., and Young, A.R. Megadroughts and late Holocene dune activation at the eastern margin of the Great Plains, north-central Kansas, USA. Aeolian Research 1, 3–4 (2010). 101110.Google Scholar
Holliday, V.T. Soils and landscape evolution of eolian plains: the southern High Plains of Texas and New Mexico. Geomorphology 3, (1990). 489515.Google Scholar
Holliday, V.T. Stratigraphy and geochronology of upper Quaternary eolian sand on the southern High Plains of Texas and New Mexico, United States. GSA Bulletin 113, 1 (2001). 88108.Google Scholar
Hugenholtz, C.H., and Wolfe, S.A. Biogeomorphic model of dunefield activation and stabilization on the northern Great Plains. Geomorphology 70, 1–2 (2005). 5370.Google Scholar
Irwin, J.H., and Morton, R.B. Hydrogeologic information on the Glorieta sandstone and the Ogallala formation in the Oklahoma Panhandle and adjoining areas as related to underground waste disposal. U.S. Geological Survey Circular vol. 630, (1969). United States Geological Survey, Google Scholar
Kilmer, V.J., and Alexander, L.T. Methods of making mechanical analyses of soils. Soil Science 68, (1949). 1524.CrossRefGoogle Scholar
Kutilek, M., and Nielsen, D.R. Soil Hydrology. (1994). Catena Verlag, Cremlingen-Destedt Germany.Google Scholar
Laird, K.R., Fritz, S.C., Cumming, B.F., and Grimm, E.C. Early-Holocene limnological and climatic variability in the northern Great Plains. The Holocene 8, (1998). 275285.Google Scholar
Laird, K.R., Fritz, S.C., Maasch, K.A., and Cumming, B.F. Greater drought intensity and frequency before AD 1200 in the northern Great Plains, USA. Nature 384, (1996). 552554.Google Scholar
Lancaster, N. Development of linear dunes in the southwestern Kalahari, Southern Africa. Journal of Arid Environments 14, (1988). 233244.Google Scholar
Lancaster, N., and Helm, P. A test of a climatic index of dune mobility using measurements from the southwestern United States. Earth Surface Processes and Landforms 25, (2000). 197207.Google Scholar
Lepper, K., and Scott, G.F. Late Holocene aeolian activity in the Cimarron River valley of west-central Oklahoma. Geomorphology 70, (2005). 4252.Google Scholar
Madole, R.F. Spatial and temporal patterns of Late Quaternary eolian deposition, Eastern Colorado, U.S.A. Quaternary Science Reviews 14, (1995). 155177.Google Scholar
Mason, J.A., Jacobs, P.M., Greene, R.S.B., and Nettleton, W.D. Sedimentary aggregates in the Peoria Loess of Nebraska, USA. Catena 53, (2003). 377397.Google Scholar
Mason, J.A., Swinehart, J.B., Goble, R.J., and Loope, D.B. Late Holocene dune activity linked to hydrological drought, Nebraska Sand Hills, USA. The Holocene 14, 2 (2004). 209217.Google Scholar
Mclaughlin, T.G. Accelerated channel erosion in the Cimarron Valley in southwestern Kansas. Journal of Geology 55, (1947). 7693.Google Scholar
Meko, D.M., (1995). Dendroclimatic evidence from the Great Plains of the United States. Climate since A.D. 1500. Bradley, R. S. and Jones, P. D.. London; New York, Routledge.: 312330.Google Scholar
Miao, X., Mason, J.A., Swinehart, J.B., Loope, D.B., Hanson, P.R., Goble, R.J., and Liu, X. A 10,000 year record of dune activity, dust storms, and severe drought in the central Great Plains. Geology 35, 2 (2007). 119122.Google Scholar
Muhs, D.R., and Holliday, V.T. Evidence of active dune sand on the Great Plains in the 19th century from accounts of early explorers. Quaternary Research 43, (1995). 198208.Google Scholar
Muhs, D.R., and Holliday, V.T. Origin of late Quaternary dune fields on the southern High Plains of Texas and New Mexico. GSA Bulletin 113, 1 (2001). 7587.Google Scholar
Muhs, D.R., and Maat, P.B. The potential response of eolian sands to greenhouse warming and precipitation reduction on the Great Plains of the U.S.A. Journal of Arid Environments 25, 4 (1993). 351361.Google Scholar
Muhs, D.R., Reynolds, R.L., Been, J., and Skipp, G. Eolian sand transport pathways in the southwestern United States: importance of the Colorado River and local sources. Quaternary International 104, (2003). 318.Google Scholar
Muhs, D.R., Stafford, T.W., Cowherd, S.D., Mahan, S.A., Kihl, R., Maat, P.B., Bush, C.A., and Nehring, J. Origin of the late Quaternary dune fields of northeastern Colorado. Geomorphology 17, (1996). 129149.Google Scholar
Muhs, D.R., and Zarate, M. Late Quaternary eolian records of the Americas and their paleoclimatic significance. Markgraf, V. Interhemispheric Climate Linkages. (2001). Academic Press, San Diego. 183216.Google Scholar
Murray, A.S., and Roberts, R.G. Factors controlling the shape of the OSL decay curve in quartz. Radiation Measurements 29, (1998). 503515.CrossRefGoogle Scholar
Murray, A.S., and Wintle, A.G. Luminescence dating of quartz using an improved single-aliquot regenerative-dose protocol. Radiation Measurements 32, (2000). 5773.Google Scholar
Namikas, S.L., and Sherman, D.J. A review of the effects of surface moisture content on aeolian sand transport. Tchakerian, V.P. Desert Aeolian Processes. (1995). Chapman and Hall, London. 269293.Google Scholar
National Climatic Data Center Monthly station normals of temperature, precipitation, and heating and cooling degree days 1971–2000: Kansas, Climatography of the United States, No. 81. National Oceanic and Atmospheric Administration. (2002). Google Scholar
National Climatic Data Center Monthly station normals of temperature, precipitation, and heating and cooling degree days 1971–2000: Oklahoma, Climatography of the United States, No. 81. National Oceanic and Atmospheric Administration. (2002). Google Scholar
Noy-Meir, I. Desert ecosystems: environment and producers. Annual Review of Ecology and Systematics 4, (1973). 2551.Google Scholar
Olson, C.G., and Porter, D.A. Isotopic and geomorphic evidence for Holocene climate, southwestern Kansas. Quaternary International 87, (2002). 2944.Google Scholar
Porter, D.A., (1997). Soil genesis and landscape evolution within the Cimarron Bend area, southwest Kansas. Unpublished Ph.D. dissertation, Kansas State University, .Google Scholar
Prescott, J.R., and Hutton, J.T. Cosmic ray contributions to dose rates for luminescence and ESR dating: large depths and long-time variations. Radiation Measurements 23, (1994). 497500.Google Scholar
Rawls, W.J., Ahuja, L.R., and Brakensiek, D.L. Estimating soil hydraulic properties from soil data. Van Genuchten, M.T., and Leij, F. Indirect Methods for Estimating the Hydraulic Properties of Unsaturated Soils. (1992). University of California and U.S. Salinity Lab, Riverside CA. 329340.Google Scholar
Reginato, R.J., and Van Bavel, C.H.M. Pressure cell for soil cores. Soil Science Society of America Proceedings 26, (1962). 13.Google Scholar
Sala, O.E., Parton, W.J., Joyce, L.A., and Lauenroth, W.K. Primary production of the central grassland region of the United States. Ecology 69, 1 (1988). 4045.Google Scholar
Schaetzl, R.J., and Anderson, S. Soils; Genesis and Geomorphology. (2005). Cambridge University Press, New York.Google Scholar
Schumm, S.A., and Lichty, R.W. Channel widening and flood-plain construction along Cimarron River in southwestern Kansas. Geological Survey Professional Paper 352-D. (1963). CrossRefGoogle Scholar
Smith, H.T.U., (1940). Geologic studies in southwestern Kansas.Google Scholar
Soil Survey Staff Spatial and tabular data of the Soil Survey for Seward County, KS. (2005). National Resource Conservation Service, Google Scholar
Soil Survey Staff Spatial and tabular data of the Soil Survey for Stevens County, KS. National Resource Conservation Service. (2005). Google Scholar
Soil Survey Staff, N.R.C.S Soil Survey Geographic (SSURGO) Database for Seward County, Ks. United States Department of Agriculture. (2009). Google Scholar
Soil Survey Staff, N.R.C.S Soil Survey Geographic (SSURGO) Database for Stevens County, Ks. United States Department of Agriculture. (2009). Google Scholar
Soil Survey Staff, N.R.C.S Soil Survey Geographic (SSURGO) Database for Texas County, OK. United States Department of Agriculture. (2009). Google Scholar
Soil Survey Staff, N.R.C.S Soil Survey Geographic (SSURGO) Database for Roberts County, Tx. United States Department of Agriculture. (2010). Google Scholar
Sridhar, V., Loope, D.B., Swinehart, J.B., Mason, J.A., Oglesby, R.J., and Rowe, C.M. Large wind shift on the Great Plains during the Medieval Warm Period. Science 313, (2006). 345347.Google Scholar
Stokes, S., and Swinehart, J.B. Middle- and late-Holocene dune reactiviation in the Nebraska Sand Hills, USA. Holocene 7, 3 (1997). 263272.Google Scholar
Taylor, H.M. Moisture relationships of some rangeland soils of the southern Great Plains. Journal of Range Management 13, (1960). 7780.Google Scholar
Thorp, J., Smith, H.T.U., (1952). Pleistocene Eolian Deposits of the United States, Alaska, and parts of Canada. National Research Council Committee for the Study of Eolian Deposits, Geological Society of America, scale 1:2,500,000.Google Scholar
Tsoar, H. Sand dunes mobility and stability in relation to climate. Physica A 357, (2005). 5056.Google Scholar
Vanlooy, J.A., and Martin, C.W. Channel and vegetation change on the Cimarron River, southwestern Kansas, 1953–2001. Annals of the Association of American Geographers 95, 4 (2005). 727739.Google Scholar
Walter, N.F., Hallberg, G.R., and Fenton, T.E. Particle size analysis by the Iowa State University Soil Laboratory. Hallberg, G.R. Standard Procedures for the Evaluation of Quaternary Materials in Iowa. Technical Information Series vol. 8, (1978). Iowa Geological Survey, 6174.Google Scholar
Wolfe, S.A. Impact of increased aridity on sand dune activity in the Canadian Prairies. Journal of Arid Environments 36, (1997). 421432.Google Scholar
Woodhouse, C.A., and Overpeck, J.T. 2000 years of drought variability in the central United States. Bulletin of the American Meteorological Association 79, 12 (1998). 26932714.Google Scholar
Yizhaq, H., Ashkenazy, Y., and Tsoar, H. Why do active and stabilized dunes coexist under the same climatic conditions?. Physical Review Letters 98, (2007). 188001-1188001-4.Google Scholar