Hostname: page-component-848d4c4894-nr4z6 Total loading time: 0 Render date: 2024-05-22T05:09:59.434Z Has data issue: false hasContentIssue false

Combining quantitative field and modelling approaches towards understanding landscape dynamics: an evolution of ideas spanning Jef Vandenberghe's research career

Published online by Cambridge University Press:  24 March 2014

G. Verstraeten*
Affiliation:
Division of Geography, Department of Earth and Environmental Sciences, KU Leuven, Celestijnenlaan 200 E, P.O. Box 2409, B-3001 Heverlee, Belgium. Email: gert.verstraeten@ees.kuleuven.be.
Rights & Permissions [Opens in a new window]

Abstract

Core share and HTML view are not available for this content. However, as you have access to this content, a full PDF is available via the ‘Save PDF’ action button.

Geomorphology as a scientific discipline has underwent major developments since the mid 20th century. From its original descriptive nature aiming to understand landscape evolution, it developed towards a more process-based oriented discipline. To a large extent this evolution followed a quantitative approach whereby modelling becomes more and more important. A schism between applied or engineering geomorphology and system-based geomorphology aiming at understanding landscape change emerges in the 1950-1960's. Only at the end of the 20th century – early 21st century, integration of quantitative field-based approaches on longer term issues of landscape evolution with numerical modelling emerges. This is particularly true for the Holocene for which the importance of human impact on geomorphic processes and landforms became acknowledged. With respect to landscape evolution on much longer timescales, the development of tectonic geomorphology becomes apparent. In this paper, some evolution of ideas and trends within geomorphology with respect to understanding landscape dynamics are summarised and put into the career perspective of Jef Vandenberghe.

Type
Research Article
Copyright
Copyright © Stichting Netherlands Journal of Geosciences 2012

References

Abrahams, A.D. & Marston, R.A., 1993. Drainage-basin sediment budgets – an introduction – A collection of papers presented at the 3rd International Geomorphology Conference, Hamilton, Ontario, August 1993. Physical Geography, 14(3): 221224.CrossRefGoogle Scholar
Ahnert, F., 1970. A comparison of theoretical slope models with slopes in the field. Zeitschrift Fur Geomorphologie, supplement band, 9: 88101.Google Scholar
Bell, M. & Walker, M.J.C., 2005. Late Quaternary Environmental Change. Physical and Human Perspectives. Pearson Eduction (Harlow), England.Google Scholar
Beuselinck, L., Steegen, A., Govers, G., Nachtergaele, J., Takken, I. & Poesen, J., 2000. Characteristics of sediment deposits formed by intense rainfall events in small catchments in the Belgian Loam Belt. Geomorphology, 32(1–2): 6982.CrossRefGoogle Scholar
Bishop, P., 2007. Long-term landscape evolution: linking tectonics and surface processes. Earth Surface Processes and Landforms, 32(3): 329365.CrossRefGoogle Scholar
Bogaart, P.W., Tucker, G.E. & De Vries, J.J., 2003a. Channel network morphology and sediment dynamics under alternating periglacial and temperate regimes: a numerical simulation study. Geomorphology 54(3–4): 257277.CrossRefGoogle Scholar
Bogaart, P.W., Van Balen, R.T., Kasse, C. & Vandenberghe, J., 2003b. Processbased modelling of fluvial system response to rapid climate change – I: model formulation and generic applications. Quaternary Science Reviews 22(20): 20772095.CrossRefGoogle Scholar
Bogaart, P.W., Van Balen, R.T., Kasse, C. & Vandenberghe, J., 2003c. Processbased modelling of fluvial system response to rapid climate change II. Application to the river Maas (the Netherlands) during the last glacialinterglacial transition. Quaternary Science Reviews, 22(20): 20972110.CrossRefGoogle Scholar
Bohncke, S., Vandenberghe, J. & Huijzer, A.S., 1993. Periglacial environments during the Weichselian late-glacial in the Maas valley, the Netherlands. Geologie en Mijnbouw 72(2): 193210.Google Scholar
Brocklehurst, S.H., 2010. Tectonics and geomorphology. Progress in Physical Geography 34(3): 357383.CrossRefGoogle Scholar
Broothaerts, N., Verstraeten, G., Notebaert, B., Kasse, C., Bohncke, S., Assendelft, R. & Vandenberghe, J., 2012. Humans reshaped the floodplain geoecology in NW Europe through intense agricultural impact, EGU General Assembly, Vienna 2012, Geophysical Research Abstracts, Vol. 14, EGU2012-7961-1.Google Scholar
Burbank, D.W. & Anderson, R.S., 2001. Tectonic Geomorphology. Blackwell Science, Malden, MA, USA.Google Scholar
Chiverrell, R.C., Thorndycraft, V.R. & Hoffmann, T.O., 2011. Cumulative probability functions and their role in evaluating the chronology of geomorphological events during the Holocene. Journal of Quaternary Science 26(1): 7685.CrossRefGoogle Scholar
Church, M., 2005. Continental drift. Earth Surface Processes and Landforms 30(1): 129130.CrossRefGoogle Scholar
Church, M., 2010. The trajectory of geomorphology. Progress in Physical Geography 34(3): 265286.CrossRefGoogle Scholar
De Moor, J.J.W., Kasse, C., Van Balen, R., Vandenberghe, J. & Wallinga, J., 2008. Human and climate impact on catchment development during the Holocene – Geul River, the Netherlands. Geomorphology 98(3–4): 316339.CrossRefGoogle Scholar
De Moor, J.J.W. & Verstraeten, G., 2008. Alluvial and colluvial sediment storage in the Geul River catchment (the Netherlands) – Combining field and modelling data to construct a Late Holocene sediment budget. Geomorphology 95(3–4): 487503.CrossRefGoogle Scholar
De Ploey, J., 1971. Liquefaction and rainwash erosion. Zeitschrift für Geomorphologie 15: 491496.CrossRefGoogle Scholar
De Ploey, J., 1974. Mechanical properties of hillslopes and their relation to gullying in Central semi-arid Tunesia. Zeitschrift für Geomorphologie, supplement band 21: 177190.Google Scholar
De Ploey, J., 1977. Some experimental-data on slopewash and wind action with reference to Quaternary morphogenesis in Belgium. Earth Surface Processes and Landforms 2(2–3): 101115.CrossRefGoogle Scholar
De Ploey, J., 1988. No-tillage experiments in the central Belgian Loess Belt. Soil Technology 1: 181184.CrossRefGoogle Scholar
De Ploey, J., 1990. Modeling the erosional susceptibility of catchments in terms of energy. Catena 17(2): 175183.CrossRefGoogle Scholar
De Ploey, J. & Moeyersons, J., 1975. Runoff creep of coarse debris: experimental data and some field observations. Catena 2: 275288.CrossRefGoogle Scholar
De Ploey, J., Moeyersons, J. & Goossens, D., 1995. The De Ploey erosional susceptibility model for catchments, E(s). Catena 25(1–4): 269314.CrossRefGoogle Scholar
De Ploey, J. & Savat, J., 1968. Contribution a l'étude de l'érosion par le splash. Zeitschrift Fur Geomorphologie 12: 174193.Google Scholar
De Ploey, J., Savat, J. & Moeyersons, J., 1976. Differential impact of some soil loss factors on flow, runoff creep and rainwash. Earth Surface Processes and Landforms 1(2): 151161.CrossRefGoogle Scholar
De Smedt, P., 1973. Paleogeografie en Kwartair-Geologie van het confluentiegebied Dijle-Demer. Acta Geographica Lovaniensia 11.Google Scholar
De Vente, J., Poesen, J., Bazzoffi, P., Van Rompaey, A. & Verstraeten, G., 2006. Predicting catchment sediment yield in Mediterranean environments: the importance of sediment sources and connectivity in Italian drainage basins. Earth Surface Processes and Landforms 31(8): 10171034.CrossRefGoogle Scholar
Desmet, P.J.J. & Govers, G., 1995. GIS-based simulation of erosion and deposition patterns in an agricultural landscape – a comparison of model results with soil map information. Catena 25(1–4): 389401.CrossRefGoogle Scholar
Desmet, P.J.J. & Govers, G., 1996. A GIS procedure for automatically calculating the USLE LS factor on topographically complex landscape units. Journal of Soil and Water Conservation 51(5): 427433.Google Scholar
Desmet, P.J.J., Poesen, J., Govers, G. & Vandaele, K., 1999. Importance of slope gradient and contributing area for optimal prediction of the initiation and trajectory of ephemeral gullies. Catena 37(3–4): 377392.CrossRefGoogle Scholar
Doyle, M.W. & Julian, J.P., 2005. The most-cited works in Geomorphology. Geomorphology 72(1–4): 238249.CrossRefGoogle Scholar
Duley, F.L. & Miller, M.F., 1923. Erosion and surface runoff under different soil conditions, University of Missouri Agricultural Experiment Station.Google Scholar
Duller, G.A.T., 2004. Luminescence dating of Quaternary sediments: recent advances. Journal of Quaternary Science 19(2): 183192.CrossRefGoogle Scholar
Dusar, B., Verstraeten, G., D'fHaen, K., Bakker, J., Kaptijn, E. & Waelkens, M., 2012. Sensitivity of the Eastern Mediterranean geomorphic system towards environmental change during the Late Holocene: a chronological perspective. Journal of Quaternary Science.CrossRefGoogle Scholar
Fuchs, M., Will, M., Kunert, E., Kreutzer, S., Fischer, M. & Reverman, R., 2011. The temporal and spatial quantification of Holocene sediment dynamics in a mesoscale catchment in northern Bavaria, Germany. Holocene 21(7): 10931104.CrossRefGoogle Scholar
Gosse, J.C. & Phillips, F.M., 2001. Terrestrial in situ cosmogenic nuclides: theory and application. Quaternary Science Reviews 20(14): 14751560.CrossRefGoogle Scholar
Govers, G., 1991. Rill erosion on arable land in central belgium – rates, controls and predictability. Catena 18(2): 133155.CrossRefGoogle Scholar
Govers, G. & Poesen, J., 1988. Assessment of the interrill and rill contributions to total soil loss from an upland field plot Geomorphology 1: 343354.Google Scholar
Govers, G., Quine, T.A., Desmet, P.J.J. & Walling, D.E., 1996. The relative contribution of soil tillage and overland flow erosion to soil redistribution on agricultural land. Earth Surface Processes and Landforms 21(10): 929946.3.0.CO;2-C>CrossRefGoogle Scholar
Graveleau, F. & Dominguez, S., 2008. Analogue modelling of the interaction between tectonics, erosion and sedimentation in foreland thrust belts. Comptes Rendus Geoscience 340(5): 324333.CrossRefGoogle Scholar
Gregory, K.J., Benito, G., Dikau, R., Golosov, V., Jones, A.J.J., Macklin, M.G., Parsons, A.J., Passmore, D.G., Poesen, J., Starkel, L. & Walling, D.E., 2006. Past hydrological events related to understanding global change: An ICSU research project. Catena 66(1–2): 213.CrossRefGoogle Scholar
Gullentops, F., 1957. Quelques phénomènes géomorphologiques depuis le Pléni-Wurm. Bulletin de la Société Belge de Géologie 61(1): 8695.Google Scholar
Gullentops, F., 1960. Quelques indices de cycles climatiques au Pleistocène inferieur et moyen en Belgique. Bulletin Peryglacjalny 9: 9195.Google Scholar
Gullentops, F., Mullenders, W., Schaillee, L., Gilot, E. & Bastin-Servais, Y., 1966. Observations géomorpologiques et palynologiques dans la vallée de la Lienne. Acta Geographica Lovaniensia 4: 192204.Google Scholar
Hancock, G.R. & Willgoose, G.R., 2002. The use of a landscape simulator in the validation of the Siberia landscape evolution model: Transient landforms. Earth Surface Processes and Landforms 27(12): 13211334.CrossRefGoogle Scholar
Hasbargen, L.E. & Paola, C., 2000. Landscape instability in an experimental drainage basin. Geology 28(12), 10671070.2.0.CO;2>CrossRefGoogle Scholar
HEC, 1968. HEC-1: Flood hydrograph package, US Army Corps of Engineers, Hydraulic Engineering Center.Google Scholar
Hoffmann, T., Erkens, G., Cohen, K.M., Houben, P., Seidel, J. & Dikau, R., 2007. Holocene floodplain sediment storage and hillslope erosion within the Rhine catchment. Holocene 17(1): 105118.CrossRefGoogle Scholar
Hoffmann, T., Erkens, G., Gerlach, R., Klostermann, J. & Lang, A., 2009. Trends and controls of Holocene floodplain sedimentation in the Rhine catchment. Catena 77(2), 96106.CrossRefGoogle Scholar
Hoffmann, T., Lang, A. & Dikau, R., 2008. Holocene river activity: analysing C-14-dated fluvial and colluvial sediments from Germany. Quaternary Science Reviews 27(21–22): 20312040.CrossRefGoogle Scholar
Hooke, R.L., 2000. On the history of humans as geomorphic agents. Geology 28(9): 843846.2.0.CO;2>CrossRefGoogle Scholar
Horton, R.E., 1945. Erosional development of streams and their drainage basins, hydrophysical approach to quantitative morphology. Bulletin of the Geological Society of America 56(3): 275370.CrossRefGoogle Scholar
Howard, A.D., Dietrich, W.E. & Seidl, M.A., 1994. Modeling fluvial erosion on regional to continental scales. Journal of Geophysical Research-Solid Earth 99(B7): 1397113986.CrossRefGoogle Scholar
Kasse, C., 1995. Younger Dryas cooling and fluvial response (Maas River, the Netherlands). Geologie en Mijnbouw 74(3): 251256.Google Scholar
Keesstra, S.D., Van Dam, O., Verstraeten, G. & Van Huissteden, J., 2009. Changing sediment dynamics due to natural reforestation in the Dragonja catchment, SW Slovenia. Catena 78(1): 6071.CrossRefGoogle Scholar
Keesstra, S.D., Van Huissteden, J., Vandenberghe, J., Van Dam, O., De Gier, J. & Pleizier, I.D., 2005. Evolution of the morphology of the river Dragonja (SW Slovenia) due to land-use changes. Geomorphology 69(1–4): 191207.CrossRefGoogle Scholar
Keylock, C.J., 2010. Introduction to special issue: The future of geomorphology. Progress in Physical Geography 34(3): 261264.CrossRefGoogle Scholar
Kirkby, M.J., 1971. Hillslope process-response models based on the continuity equation. Transactions of the Institute of British Geographers, special publication 3: 1530.Google Scholar
Kooi, H. & Beaumont, C., 1996. Large-scale geomorphology: Classical concepts reconciled and integrated with contemporary ideas via a surface processes model. Journal of Geophysical Research – Solid Earth 101(B2): 33613386.CrossRefGoogle Scholar
Lentz, G.H., Sinclair, J.D. & Meginnis, G.H., 1930. Soil erosion in the silt loam uplands of Mississippi, Southern Forest Experiment Station. Journal of Forestry: 971977.Google Scholar
Leopold, L.B. & Maddock, T.J., 1953. The hydraulic geometry of stream channels and some physiographic implications. US Geol. Survey Prof. Paper 252.CrossRefGoogle Scholar
Leopold, L.B. & Wolman, M.G., 1957. River channel patterns: braided, meandering and straight. US Geol. Survey Prof. Paper 282-B.CrossRefGoogle Scholar
Macklin, M.G., Benito, G., Gregory, K.J., Johnstone, E., Lewin, J., Michczyńska, D.J., Soja, R., Starkel, L. & Thorndycraft, V.R., 2006. Past hydrological events reflected in the Holocene fluvial record of Europe. Catena 66(1–2): 145154.CrossRefGoogle Scholar
Macklin, M.G. & Lewin, J., 2003. River sediments, great floods and centennialscale Holocene climate change. Journal of Quaternary Science 18(2): 101105.CrossRefGoogle Scholar
Messerli, B., Grosjean, M., Hofer, T., Nunez, L. & Pfister, C., 2000. From naturedominated to human-dominated environmental changes. Quaternary Science Reviews 19(1–5): 459479.CrossRefGoogle Scholar
Meyer-Peter, E. & Müller, R., 1948. Formulas for Bed-Load transport. Proceedings of the 2nd IAHSR meeting, 7 06 1948, Stockholm.Google Scholar
Montgomery, D.R., 2007. Dirt. The Erosion of Civilizations. University of California Press (Berkeley).CrossRefGoogle Scholar
Montgomery, D.R. & Dietrich, W.E., 1992. Channel initiation and the problem of landscape scale. Science 255(5046): 826830.CrossRefGoogle ScholarPubMed
Mucher, H.J. & De Ploey, J., 1977. Experimental and micromorphological investigation of erosion and redeposition of loess by water. Earth Surface Processes and Landforms 2(2–3): 117124.CrossRefGoogle Scholar
Mullenders, W. & Gullentops, F., 1956. Evolution de la végétation et de la plaine alluviale de la Dyle, à Louvain, depuis le Pleni-Wurm. Mededelingen van de Klasse der Wetenschappen, Koninklijke Academie van België 42: 11231137.Google Scholar
Mullenders, W. & Gullentops, F., 1957. Palynologisch en geologisch onderzoek in de alluviale vlakte van de Dijle te Heverlee-Leuven. Agricultura 5: 5764.Google Scholar
Mullenders, W., Gullentops, F., Lorent, J., Coremans, M. & Gilot, E., 1966. Le remblaiement de la vallée de la Néthen. Acta Geographica Lovaniensia 4: 169181.Google Scholar
Notebaert, B. & Verstraeten, G., 2010. Sensitivity of West and Central European river systems to environmental changes during the Holocene A review. Earth-Science Reviews 103(3–4): 163182.CrossRefGoogle Scholar
Notebaert, B., Verstraeten, G., Govers, G. & Poesen, J., 2010. Quantification of alluvial sediment storage in contrasting environments: Methodology and error estimation. Catena, 82(3): 169182.CrossRefGoogle Scholar
Notebaert, B., Verstraeten, G., Rommens, T., Vanmontfort, B., Govers, G. & Poesen, J., 2009. Establishing a Holocene sediment budget for the river Dijle. Catena 77(2): 150163.CrossRefGoogle Scholar
Notebaert, B., Verstraeten, G., Vandenberghe, D., Marinova, E., Poesen, J. & Govers, G., 2011a. Changing hillslope and fluvial Holocene sediment dynamics in a Belgian loess catchment. Journal of Quaternary Science 26(1): 4458.CrossRefGoogle Scholar
Notebaert, B., Verstraeten, G., Ward, P., Renssen, H. & Van Rompaey, A., 2011b. Modeling the sensitivity of sediment and water runoff dynamics to Holocene climate and land use changes at the catchment scale. Geomorphology 126(1–2): 1831.CrossRefGoogle Scholar
Peeters, I., Rommens, T., Verstraeten, G., Govers, G., Van Rompaey, A., Poesen, J., & Van Oost, K., 2006. Reconstructing ancient topography through erosion modelling. Geomorphology 78(3–4): 250264.CrossRefGoogle Scholar
Peeters, I., Van Oost, K., Govers, G., Verstraeten, G., Rommens, T. & Poesen, J., 2008. The compatibility of erosion data at different temporal scales. Earth and Planetary Science Letters 265(1–2): 138152.CrossRefGoogle Scholar
Phillips, J.D., 1991. Fluvial sediment budgets in the north-carolina piedmont. Geomorphology 4(3–4): 231241.CrossRefGoogle Scholar
Poesen, J., 1986. Field measurements of splash erosion to validate a splash transport model. Zeitschrift Fur Geomorphologie, supplement band 58: 8191.Google Scholar
Poesen, J.W., Vandaele, K. & Van Wesemael, B., 1996. Contribution of gully erosion to sediment production on cultivated lands and rangelands. In: Walling, D.E.W.B.W. (ed.): Erosion and Sediment Yield: Global and Regional Perspectives. Iahs Publications: 251266.Google Scholar
Reiners, P.W. & Brandon, M.T., 2006. Using thermochronology to understand orogenic erosion, Annual Review of Earth and Planetary Sciences. Annual Review of Earth and Planetary Sciences: 419466.Google Scholar
Reynolds, O., 1883. An experimental investigation of the circumstances which determine whether the motion of water shall be direct or sinuous, and the law of resistance in parallel channels. Philosophical Transactions of the Royal Society of London 174: 935982.Google Scholar
Richards, K. & Clifford, N., 2008. Science, systems and geomorphologies: why LESS may be more. Earth Surface Processes and Landforms 33(9): 13231340.CrossRefGoogle Scholar
Rommens, T., Verstraeten, G., Peeters, I., Poesen, J., Govers, G., Van Rompaey, A., Mauz, B., Packman, S. & Lang, A., 2007. Reconstruction of late-Holocene slope and dry valley sediment dynamics in a Belgian loess environment. Holocene 17(6): 777788.CrossRefGoogle Scholar
Rommens, T., Verstraeten, G., Poesen, J., Govers, G., Van Rompaey, A., Peeters, I. & Lang, A., 2005. Soil erosion and sediment deposition in the Belgian loess belt during the Holocene: establishing a sediment budget for a small agricultural catchment. Holocene 15(7): 10321043.CrossRefGoogle Scholar
Schaller, M., Von Blanckenburg, F., Hovius, N. & Kubik, P.W., 2001. Large-scale erosion rates from in situ-produced cosmogenic nuclides in European river sediments. Earth and Planetary Science Letters 188(3–4): 441458.CrossRefGoogle Scholar
Schumm, S.A., 1977. The Fluvial System. John Wiley & Sons Ltd (New York) USA.Google Scholar
Shields, A., 1936. Anwendung der Aehnlichkeitsmechanik und der Turbulenzforschung auf die Geschiebebewegung. Mitteilungen der Preussischen Versuchsanhalt fur Wasserbau und Schiffbau: 26.Google Scholar
Steegen, A., Govers, G., Nachtergaele, J., Takken, I., Beuselinck, L. & Poesen, J., 2000. Sediment export by water from an agricultural catchment in the Loam Belt of central Belgium. Geomorphology 33(1–2): 2536.CrossRefGoogle Scholar
Stolz, C., 2011. Budgeting soil erosion from floodplain sediments of the central Rhenish Slate Mountains (Westerwald), Germany. Holocene 21(3): 499510.CrossRefGoogle Scholar
Strahler, A.N., 1952. Hypsometric (Area-Altitude) analysis of erosional topography. Bulletin of the Geological Society of America 63: 11171142.CrossRefGoogle Scholar
Summerfield, M.A., 2000. Geomorphology and Global Tectonics. Wiley (Chichester) UK.Google Scholar
Summerfield, M.A., 2005. The changing landscape of geomorphology. Earth Surface Processes and Landforms 30(6): 779781.CrossRefGoogle Scholar
Syvitski, J.P.M. & Milliman, J.D., 2007. Geology, geography, and humans battle for dominance over the delivery of fluvial sediment to the coastal ocean. Journal of Geology 115(1): 119.CrossRefGoogle Scholar
Temme, A., Peeters, I., Buis, E., Veldkamp, A. & Govers, G., 2011. Comparing landscape evolution models with quantitative field data at the millennial time scale in the Belgian loess belt. Earth Surface Processes and Landforms 36(10): 13001312.CrossRefGoogle Scholar
Trimble, S.W., 1999. Decreased rates of alluvial sediment storage in the Coon Creek Basin, Wisconsin, 1975-93. Science 285(5431): 12441246.CrossRefGoogle ScholarPubMed
Tucker, G.E. & Slingerland, R.L., 1994. Erosional dynamics, flexural isostasy, and long-lived escarpments – a numerical modeling study. Journal of Geophysical Research-Solid Earth 99(B6): 1222912243.CrossRefGoogle Scholar
Turowski, J.M., Lague, D., Crave, A. & Hovius, N., 2006. Experimental channel response to tectonic uplift. J. Geophys. Res. 111(F3): F03008.CrossRefGoogle Scholar
Van Kolfschoten, T., Roebroeks, W. & Vandenberghe, J., 1993. The Middle and Late Pleistocene sequence at Maastricht-Belvedere: the type locality of the Belvedere Interglacial. Mededelingen Rijks Geologische Dienst, N.S. 47: 8191.Google Scholar
Van Oost, K., Govers, G. & Desmet, P., 2000. Evaluating the effects of changes in landscape structure on soil erosion by water and tillage. Landscape Ecology 15(6): 577589.CrossRefGoogle Scholar
Van Rompaey, A.J.J., Verstraeten, G., Van Oost, K., Govers, G. & Poesen, J., 2001. Modelling mean annual sediment yield using a distributed approach. Earth Surface Processes and Landforms 26(11): 12211236.CrossRefGoogle Scholar
Vanacker, V., Von Blanckenburg, F., Govers, G., Molina, A., Poesen, J., Deckers, J. & Kubik, P., 2007. Restoring dense vegetation can slow mountain erosion to near natural benchmark levels. Geology 35(4): 303306.CrossRefGoogle Scholar
Vandaele, K. & Poesen, J., 1995. Spatial and temporal patterns of soil-erosion rates in an agricultural catchment, central Belgium. Catena 25(1–4): 213226.CrossRefGoogle Scholar
Vandenberghe, J., 1969. Enquêtes géoélectriques dans la méandre recoupé d'Annevoie. Acta Geographica Lovaniensia 7, 93103.Google Scholar
Vandenberghe, J., 1973. Geomorfologie van de Zuiderkempen. Phd Geography, KU Leuven, Leuven.Google Scholar
Vandenberghe, J., 1977. Geomorphology of the Zuiderkempen Belgium. Verhandelingen van de Koninklijke Academie voor Wetenschappen, Letteren en Schone Kunsten van België 39(140).Google Scholar
Vandenberghe, J., 1995. Timescales, climate and river development. Quaternary Science Reviews 14(6): 631638.CrossRefGoogle Scholar
Vandenberghe, J., 2008. The fluvial cycle at cold-warm-cold transitions in lowland regions: A refinement of theory. Geomorphology 98(3–4): 275284.CrossRefGoogle Scholar
Vandenberghe, J. & De Smedt, P., 1979. Palaeomorphology in the Eastern Scheldt basin (Central Belgium). The Dijle-Demer-Grote Nete Confluence Area. Catena 6: 73105.CrossRefGoogle Scholar
Vandenberghe, J., Kasse, C., Bohncke, S. & Kozarski, S., 1994. Climate-related river activity at the weichselian holocene transition – a comparative study of the warta and maas rivers. Terra Nova 6(5): 476485.CrossRefGoogle Scholar
Verstraeten, G., Lang, A. & Houben, P., 2009a. Human impact on sediment dynamics – quantification and timing Introduction. Catena 77(2): 7780.CrossRefGoogle Scholar
Verstraeten, G. & Poesen, J., 1999. The nature of small-scale flooding, muddy floods and retention pond sedimentation in central Belgium. Geomorphology 29(3-4): 275292.CrossRefGoogle Scholar
Verstraeten, G., Poesen, J., De Vente, J. & Koninckx, X., 2003. Sediment yield variability in Spain: a quantitative and semiqualitative analysis using reservoir sedimentation rates. Geomorphology 50(4): 327348.CrossRefGoogle Scholar
Verstraeten, G., Rommens, T., Peeters, I., Poesen, J., Govers, G. & Lang, A., 2009b. A temporarily changing Holocene sediment budget for a loess-covered catchment (central Belgium). Geomorphology 108(1–2): 2434.CrossRefGoogle Scholar
Verstraeten, G., Van Oost, K., Van Rompaey, A., Poesen, J. & Govers, G., 2002. Evaluating an integrated approach to catchment management to reduce soil loss and sediment pollution through modelling. Soil Use and Management 18(4): 386394.CrossRefGoogle Scholar
Ward, P.J., Aerts, J.C.J.H., De Moel, H. & Renssen, H., 2007. Verification of a coupled climate-hydrological model against Holocene palaeohydrological records. Global and Planetary Change 57(3–4): 283300.CrossRefGoogle Scholar
Ward, P.J., Van Balen, R.T., Verstraeten, G., Renssen, H. & Vandenberghe, J., 2009. The impact of land use and climate change on late Holocene and future suspended sediment yield of the Meuse catchment. Geomorphology 103(3): 389400.CrossRefGoogle Scholar
Wilkinson, B.H. & McElroy, B.J., 2007. The impact of humans on continental erosion and sedimentation. Geological Society of America Bulletin 119(1–2): 140156.CrossRefGoogle Scholar
Willgoose, G., Bras, R.L. & Rodrigueziturbe, I., 1991. A coupled channel network growth and hillslope evolution model. 1. Theory. Water Resources Research 27(7): 16711684.CrossRefGoogle Scholar
Wischmeier, W.H. & Smith, D.D., 1960. A universal soil-loss equation to guide conservation farm planning. Transactions International Congress Soil Science: 418425.Google Scholar
Wolman, M.G. & Leopold, L.B., 1957. River floodplains: Some observations on their formation. US Geol. Survey Prof. Paper 282-C: C87-C107.CrossRefGoogle Scholar