Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-18T20:40:35.140Z Has data issue: false hasContentIssue false

A soil Chronosequence on Lake Mega-Frome Beach Ridges and its Implications for Late Quaternary Pedogenesis and Paleoenvironmental Conditions in the Drylands of Southern Australia

Published online by Cambridge University Press:  20 January 2017

Jan-Hendrik May
Affiliation:
GeoQuest Research Centre, School of Earth and Environmental Sciences, University of Wollongong, Wollongong, 2522 NSW, Australia
Stephen G. Wells
Affiliation:
Desert Research Institute, Nevada System of Higher Education, Reno, NV 89512, USA
Timothy J. Cohen
Affiliation:
GeoQuest Research Centre, School of Earth and Environmental Sciences, University of Wollongong, Wollongong, 2522 NSW, Australia
Samuel K. Marx
Affiliation:
GeoQuest Research Centre, School of Earth and Environmental Sciences, University of Wollongong, Wollongong, 2522 NSW, Australia
Gerald C. Nanson
Affiliation:
GeoQuest Research Centre, School of Earth and Environmental Sciences, University of Wollongong, Wollongong, 2522 NSW, Australia
Sophie E. Baker
Affiliation:
Desert Research Institute, Nevada System of Higher Education, Reno, NV 89512, USA

Abstract

The terminal lake systems of central Australia are key sites for the reconstruction of late Quaternary paleoenvironments. Paleoshoreline deposits around these lakes reflect repeated lake filling episodes and such landforms have enabled the establishment of a luminescence-based chronology for filling events in previous studies. Here we present a detailed documentation of the morphology and chemistry of soils developed in four well-preserved beach ridges of late Pleistocene and mid-to-late Holocene age at Lake Callabonna to assess changes in dominant pedogenic processes. All soil profiles contain evidence for the incorporation of eolian-derived material, likely via the formation of desert pavements and vesicular horizons, and limited illuviation due to generally shallow wetting fronts. Even though soil properties in the four studied profiles also provide examples of parent material influence or site-specific processes related to the geomorphic setting, there is an overall trend of increasing enrichment of eolian-derived material since at least ~ 33 ka. Compared to the Holocene profiles, the derived average accumulation rates for the late Pleistocene profiles are significantly lower and may suggest that soils record important regional changes in paleoenvironments and dust dynamics related to shifts in the Southern Hemisphere westerlies.

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

Al-Farraj, A. (2008). Desert pavement development on the lake shorelines of Lake Eyre (South), South Australia.. Geomorphology 100, 154163.Google Scholar
Anderson, K., Wells, S., and Graham, R. (2002). Pedogenesis of vesicular horizons, Cima volcanic field, Mojave Desert, California.. Soil Science Society of America Journal 66, 878887.Google Scholar
Barrett, L.R. (2001). A strand plain soil development sequence in Northern Michigan, USA.. Catena 44, 163186.CrossRefGoogle Scholar
Birkeland, P.W. (1992). Quaternary soil chronosequences in various environments – extremely arid to humid tropical.. In: Martini, I.P., Chesworth, W. (Eds.), Soils & Paleosols, Weathering., pp. 261281.Google Scholar
Bowler, J.M. (1998). Willandra Lakes revisited: environmental framework for human occupation.. Archaeology in Oceania 33, 120155.Google Scholar
Bowler, J.M., Qi, H., Kezao, C., Head, M.J., and Baoyin, Y. (1986). Radiocarbon dating of playa-lake hydrologic changes: examples from northwestern China and central Australia.. Palaeogeography, Palaeoclimatology, Palaeoecology 54, 241260.CrossRefGoogle Scholar
Bullard, J.E., and White, K. (2005). Dust production and the release of iron oxides resulting from the aeolian abrasion of natural dune sands.. Earth Surface Processes and Landforms 30, 95106.Google Scholar
Bullard, J., Baddock, M., McTainsh, G., and Leys, J. (2008). Sub-basin scale dust source geomorphology detected using MODIS.. Geophysical Research Letters 35, 16.Google Scholar
Burrough, S., and Thomas, D. (2009). Geomorphological contributions to palaeolimnology on the African continent.. Geomorphology 103, 285298.Google Scholar
Callen, R.A., and Tedford, R.H. (1976). New Late Cenozoic rock units and depositional environments, Lake Frome Area, South Australia.. Transactions of the Royal Society of South Australia 100, 125167.Google Scholar
Cattle, S.R., McTainsh, G.H., and Wagner, S. (2002). Æolian dust contributions to soil of the Namoi Valley, northern NSW, Australia.. Catena 47, 245264.Google Scholar
Cohen, T.J., Nanson, G.C., Jansen, J.D., Jones, B.G., Jacobs, Z., Treble, P., Price, D.M., May, J.-H., Smith, a.M., Ayliffe, L.K., and Hellstrom, J.C. (2011). Continental aridification and the vanishing of Australia's megalakes.. Geology 39, 167170.Google Scholar
Cohen, T.J., Nanson, G.C., Jansen, J.D., Gliganic, L.A., May, J.H., Larsen, J.R., Goodwin, I.D., Browning, S., and Price, D.M. (2012a). A pluvial episode identified in arid Australia during the Medieval Climatic Anomaly.. Quaternary Science Reviews 56, 167171.Google Scholar
Cohen, T.J., Nanson, G.C., Jansen, J.D., Jones, B.G., Jacobs, Z., Larsen, J.R., May, J.-H., Treble, P., Price, D.M., and Smith, A.M. (2012b). Late Quaternary mega-lakes fed by the northern and southern river systems of central Australia: varying moisture sources and increased continental aridity.. Palaeogeography, Palaeoclimatology, Palaeoecology 356–357, 89108.Google Scholar
Cornell, R.M., and Schwertmann, U. (2006). The Iron Oxides: Structure, Properties, Reactions, Occurrences and Uses. John Wiley & Sons, .Google Scholar
De Deckker, P., Magee, J.W., and Shelley, J.M.G. (2011). Late Quaternary palaeohydrological changes in the large playa Lake Frome in central Australia, recorded from the Mg/Ca and Sr/Ca in ostracod valves and biotic remains.. Journal of Arid Environments 75, 3850.CrossRefGoogle Scholar
DeVogel, S.B., Magee, J.W., Manley, W.F., and Miller, G.H. (2004). A GIS-based reconstruction of late Quaternary paleohydrology: Lake Eyre, arid central Australia.. Palaeogeography, Palaeoclimatology, Palaeoecology 204, 113.Google Scholar
Dietze, M., and Kleber, A. (2012). Contribution of lateral processes to stone pavement formation in deserts inferred from clast orientation patterns.. Geomorphology 139, 172187.Google Scholar
Dietze, M., Bartel, S., Lindner, M., and Kleber, A. (2012). Formation mechanisms and control factors of vesicular soil structure.. Catena 99, 8396.Google Scholar
Dixon, J.C. (2013). Pedogenesis with respect to geomorphology.. In: Shroder, J.F. (Ed.), Treatise on Geomorphology. Academic Press, San Diego., 2743.Google Scholar
Draper, J.J., and Jensen, A.R. (1976). The geochemistry of Lake Frome, a playa lake in South Australia.. BMR Journal of Geology and Geophysics 1, 83104.Google Scholar
Fedoroff, N., and Courty, M.A. (1999). Soil and soil forming processes under increasing aridity.. In: Singhvi, A.K., Derbyshire, E. (Eds.), Paleoenvironmental Reconstruction in Arid Lands. Oxford & IBH Publishing Co. Pvt. Ltd., New Delhi, Calcutta., pp. 73108.Google Scholar
Fitzsimmons, K.E. (2007). Morphological variability in the linear dunefields of the Strzelecki and Tirari Deserts, Australia.. Geomorphology 91, 146160.Google Scholar
Fitzsimmons, K.E., Magee, J.W., and Amos, K.J. (2009). Characterisation of aeolian sediments from the Strzelecki and Tirari Deserts, Australia: implications for reconstructing palaeoenvironmental conditions.. Sedimentary Geology 218, 6173.Google Scholar
Fitzsimmons, K.E., Cohen, T.J., Hesse, P.P., Jansen, J., Nanson, G.C., May, J.-H., Barrows, T.T., Haberlah, D., Hilgers, A., and Kelly, T. (2013). Late Quaternary palaeoenvironmental change in the Australian drylands.. Quaternary Science Reviews 74, 7896.Google Scholar
Gee, G.W., Bauder, J.W., and Klute, A. (1986). Particle-size analysis. Methods of soil analysis.. Part 1. Physical and Mineralogical, Methods 383411.Google Scholar
Gell, P.A., and Bickford, S. (1996). Vegetation.. In: Davies, M., Twidale, C.R., Tyler, M.J. (Eds.), Natural History of the Flinders Ranges, pp. 86101.Google Scholar
Gile, L.H., and Grossman, R.B. (1979). The desert project soil monograph: soils and landscapes of a desert region astride the Rio Grande Valley near Las Cruces, New Mexico. US Department of Agriculture, Soil Conservation Service, Washington, DC.Google Scholar
Gile, L.H., Peterson, F.F., and Grossman, R.B. (1966). Morphological and genetic sequences of carbonate accumulation in desert soils.. Soils Science 101, 347360.Google Scholar
Gliganic, L.A., Cohen, T.J., May, J.H., Jansen, J.D., Nanson, G.C., Dosseto, a., Larsen, J.R., and Aubert, M. (2014). Late-Holocene climatic variability indicated by three natural archives in arid southern Australia.. The Holocene 24, 104117.Google Scholar
Haberlah, D., Glasby, P., Williams, M.A.J., Hill, S.M., Williams, F., Rhodes, E.J., Gostin, V., O'Flaherty, A., and Jacobsen, G.E. (2010a). ‘Of droughts and flooding rains’: an alluvial loess record from central South Australia spanning the last glacial cycle.. Geological Society, London, Special Publications 346, 185223.CrossRefGoogle Scholar
Haberlah, D., Williams, M.A.J., Halverson, G., McTainsh, G.H., Hill, S.M., Hrstka, T., Jaime, P., Butcher, A.R., and Glasby, P. (2010b). Loess and floods: high-resolution multi-proxy data of Last Glacial Maximum (LGM) slackwater deposition in the Flinders Ranges, semi-arid South Australia.. Quaternary Science Reviews 29, 26732693.Google Scholar
Harden, J.W. (1982). A quantitative index of soil development from field descriptions: examples from a chronosequence in central California.. Geoderma 28, 128.Google Scholar
Hesse, P.P. (1994). The record of continental dust from Australia in Tasman Sea Sediments.. Quaternary Science Reviews 13, 257272.Google Scholar
Hesse, P.P., and McTainsh, G.H. (2003). Australian dust deposits: modern processes and the Quaternary record.. Quaternary Science Reviews 22, 20072035.Google Scholar
Huggett, R.J. (1998). Soil chronosequences, soil development, and soil evolution: a critical review.. Catena 32, 155172.CrossRefGoogle Scholar
Hurst, V.J. (1977). Visual estimation of iron in saprolite.. Geological Society of America Bulletin 88, 174176.Google Scholar
Johnson, D.L., Keller, E.A., and Rockwell, T.K. (1990). Dynamic pedogenesis: new views on some key soil concepts, and a model for interpreting Quaternary soils.. Quaternary Research 33, 306319.Google Scholar
Kendrick, K.J. (2007). Pedogenic silica accumulation.. Encyclopedia of Soil Science Second edition Taylor & Francis, 12511253.Google Scholar
Lebrón, I., Herrero, J., and Robinson, D. (2009). Determination of gypsum content in dryland soils exploiting the gypsum–bassanite phase change.. Soil Science Society of America Journal 73, 403411.Google Scholar
Leon, J.X., and Cohen, T.J. (2012). An improved bathymetric model for the modern and palaeo Lake Eyre.. Geomorphology 173–174, 6979.Google Scholar
Luly, J.G. (2001). On the equivocal fate of Late Pleistocene Callitris Vent. (Cupressaceae) woodlands in arid South Australia.. Quaternary International 83–85, 155168.Google Scholar
Machette, M.N. (1985). Calcic soils of the Southwestern United States.. GSA Special Paper 203, .Google Scholar
Magee, J.W., Bowler, J.M., Miller, G.H., and Williams, D.L.G. (1995). Stratigraphy, sedimentology, chronology and palaeohydrology of Quaternary lacustrine deposits at Madigan Gulf, Lake Eyre, South Australia.. Palaeogeography, Palaeoclimatology, Palaeoecology 113, .Google Scholar
Marx, S.K., McGowan, H.A., and Kamber, B.S. (2009). Long-range dust transport from eastern Australia: a proxy for Holocene aridity and ENSO-type climate variability.. Earth and Planetary Science Letters 282, 167177.Google Scholar
Marx, S.K., Kamber, B.S., McGowan, H.A., and Denholm, J. (2011). Holocene dust deposition rates in Australia's Murray–Darling Basin record the interplay between aridity and the position of the mid-latitude westerlies.. Quaternary Science Reviews 30, 32903305.CrossRefGoogle Scholar
McFadden, L.D., and Wells, S.G. (1987). Influences of eolian and pedogenic processes on the origin and evolution of desert pavements.. Geology 15, 504508.Google Scholar
McFadden, L.D., Wells, S.G., Brown, W.J., and Enzel, Y. (1992). Soil genesis on beach ridges of pluvial Lake Mojave: implications for Holocene lacustrine and eolian events in the Mojave Desert, Southern California.. Catena 19, 7797.Google Scholar
McFadden, L.D., McDonald, E.V., Wells, S.G., Anderson, K., Quade, J., and Forman, S.L. (1998). The vesicular layer and carbonate collars of desert soils and pavements: formation, age and relation to climate change.. Geomorphology 24, 101145.Google Scholar
McTainsh, G. (1989). Quaternary aeolian dust processes and sediments in the Australian region.. Quaternary Science Reviews 8, 235253.Google Scholar
McTainsh, G., and Lynch, A. (1996). Quantitative estimates of the effect of climate change on dust storm activity in Australia during the Last Glacial Maximum.. Geomorphology 17, 263271.Google Scholar
McTainsh, G.H., Love, B.M., Leys, J.F., and Strong, C. (2002). Wind erodibility of arid lands in the Channel Country of western Queensland, Australia, a sequel (1994–2000).. In: Lee, J.A., Zobeck, T.M. (Eds.), ICAR5/GCTE-SEN Joint conference. International Center for Arid and Semiarid Lands Studies, Texas Tech University, Texas, USA., pp. 179183.Google Scholar
Meadows, D.G., Young, M.H., and McDonald, E.V. (2008). Influence of relative surface age on hydraulic properties and infiltration on soils associated with desert pavements.. Catena 72, 169178.Google Scholar
Monger, H.C. (2006). Soil development in the Jornada Basin.. In: Havstad, K.M., Huenneke, L.F., Schlesinger, W.H. (Eds.), Structure and function of a Chihuahuan Desert ecosystem: The Jornada Basin Long Term Ecological Research site. Oxford Univ. Press, New York., pp. 81106.Google Scholar
Nanson, G.C., and Price, D.W. (1998). Quaternary change in the Lake Eyre basin of Australia: an introduction.. Palaeogeography, Palaeoclimatology, Palaeoecology 144, 235237.Google Scholar
Nanson, G.C., Callen, R.A., and Price, D.M. (1998). Hydroclimatic interpretation of Quaternary shorelines on South Australian playas.. Palaeogeography, Palaeoclimatology, Palaeoecology 144, 281305.Google Scholar
Quade, J., Chivas, A., and McCulloch, M. (1995). Strontium and carbon isotope tracers and the origins of soil carbonate in South Australia and Victoria.. Palaeogeography, Palaeoclimatology, Palaeoecology 113, 103117.Google Scholar
Reheis, M.C., Goodmacher, J.C., Harden, J.W., McFadden, L.D., Thomas, K., Shroba, R.R., Sowers, J.M., and Taylor, E.M. (1995). Quaternary soils and dust deposition in southern Nevada and California Quaternary soils and dust deposition in southern Nevada and California.. Geological Society of America Bulletin 107, 10031022.2.3.CO;2>CrossRefGoogle Scholar
Rhoades, J., Sparks, D., Page, A., Helmke, P., Loeppert, R., Soltanpour, P., Tabatabai, M., Johnston, C., and Sumner, M. (1996). Salinity: electrical conductivity and total dissolved solids.. Methods of Soil Analysis. Part 3–Chemical Methods 417435.Google Scholar
Rose, D.A., Kanukcu, F., and Gowing, J.W. (2005). Effect of watertable depth on evaporation and salt accumulation from saline groundwater.. Australian Journal of Soil Research 43, 565573.Google Scholar
Ross, G., and Wang, C. (1993). Extractable Al, Fe, Mn, and Si.. Soil Sampling and Methods of Analysis 1993 239246.Google Scholar
Saxton, K.E., and Rawls, W.J. (2006). Soil water characteristic estimates by texture and organic matter for hydrologic solutions.. Soil Science Society of America Journal 70, 15691578.Google Scholar
Schaetzl, R.J., and Anderson, S. (2005). Soils: Genesis and Geomorphology. Cambridge University Press, .CrossRefGoogle Scholar
Schlesinger, W.H. (1985). The formation of caliche in soils of the Mojave Desert, California.. Geochimica et Cosmochimica Acta 49, 5766.Google Scholar
Schmid, G.L. (2013). Soil chronosequences. In: Shroder, J.F. (Ed.), Treatise on Geomorphology. Academic Press, San Diego., 277283.Google Scholar
Schoeneberger, P.J., Wysocki, D.A., Benham, E.C., and Broderson, W.D. (2002). Field Book for Describing and Sampling Soils, Version 2.0. Natural Resources Conservation Service, National Soil Survey Center, Lincoln, NE.Google Scholar
Schwerdtfeger, P., and Curran, E. (1996). Climate of the Flinders Ranges.. In: Davies, M., Twidale, C.R., Tyler, M.J. (Eds.), Natural History of the Flinders Ranges, pp. 6375.Google Scholar
Sheard, M.J. (2009). Explanatory notes for CALLABONNA 1:250.000 Geological Map, sheet SH 54-6.. The Journal of the Royal College of General Practitioners..Google Scholar
Sherrod, L., Dunn, G., Peterson, G., and Kolberg, R. (2002). Inorganic carbon analysis by modified pressure-calcimeter method.. Soil Science Society of America Journal 66, 299305.Google Scholar
Shulmeister, J., Goodwin, I., Renwick, J., Harle, K., Armand, L., McGlone, M.S., Cook, E., Dodson, J., Hesse, P.P., Mayewski, P., and Curran, M. (2004). The Southern Hemisphere westerlies in the Australasian sector over the last glacial cycle: a synthesis.. Quaternary International 118–119, 2353.Google Scholar
Singh, G. (1981). Late Quaternary pollen records and seasonal palaeoclimates of Lake Frome, South Australia.. Hydrobiologia 82, 419430.Google Scholar
Singh, G., and Luly, J. (1991). Changes in vegetation and seasonal climate since the last full glacial at Lake Frome, South Australia.. Palaeogeography, Palaeoclimatology, Palaeoecology 84, 7586.Google Scholar
Sprigg, R.C. (1982). Alternating wind cycles of the Quaternary era and their influences on aeolian sedimentation in and around the dune deserts of South Australia.. In: Wasson, R.J. (Ed.), Australian National University, Canberra., 211240.Google Scholar
Summerfield, M.A. (1982). Distribution, nature and probable genesis of silcrete in arid and semi-arid southern Africa.. Catena (Suppl. 1), 3766.Google Scholar
Thomas, G., Sparks, D., Page, A., Helmke, P., Loeppert, R., Soltanpour, P., Tabatabai, M., Johnston, C., and Sumner, M. (1996). Soil pH and soil acidity.. Methods of Soil Analysis. Part 3–Chemical, Methods 475490.Google Scholar
Torrent, J., Schwertmann, U., and Schulze, D.G. (1980). Iron oxide mineralogy of some soils of two river terrace sequences in Spain.. Geoderma 23, 191208.Google Scholar
Torrent, J., Schwertmann, U., Fechter, H., and Alferez, F. (1983). Quantitative relationships between soil color and hematite content.. Soil Science 136, 354358.Google Scholar
Ullman, W.J., and Collerson, K.D. (1994). The Sr-isotope record of late quaternary hydrologic changes around Lake Frome, South Australia.. Australian Journal of Earth Sciences 41, 3745.Google Scholar
Wells, S.G., and Dohrenwend, J.C. (1985). Relict sheetflood bed forms on late Quaternary alluvial-fan surfaces in the southwestern United States.. Geology 13, 512516.Google Scholar
Wells, S.G., Dohrenwend, J.C., McFadden, L.D., Turrin, B.D., and Mahrer, K.D. (1985). Late Cenozoic landscape evolution on lava flow surfaces of the Cima volcanic field, Mojave Desert, California.. Geological Society of America Bulletin 96, 15181529.2.0.CO;2>CrossRefGoogle Scholar
Williams, G.E. (1973). Late Quaternary piedmont sedimentation, soil formation and paleoclimates in arid Australia.. Zeitschrift für Geomorphologie N. F. 17, 102125.Google Scholar
Yaalon, D.H. (1979). Soils in the Mediterranean region: what makes them different?.. Catena 28, 157169.Google Scholar
Yaalon, D., and Ganor, E. (1973). The influence of dust on soils during the Quaternary.. Soil Science 116, 146155.Google Scholar
Zielhofer, C., Recio Espejo, J.M., Núnez Granados, M.À., and Faust, D. (2009). Durations of soil formation and soil development indices in a Holocene Mediterranean floodplain.. Quaternary International 209, 4465.Google Scholar