Skip to main content Accessibility help
×
Hostname: page-component-f554764f5-c4bhq Total loading time: 0 Render date: 2025-04-21T13:33:59.464Z Has data issue: false hasContentIssue false

Part II - Connectivity in Process Domains

Published online by Cambridge University Press:  10 April 2025

Ronald Pöppl
Affiliation:
BOKU University Vienna
Anthony Parsons
Affiliation:
University of Sheffield
Saskia Keesstra
Affiliation:
Wageningen Universiteit, The Netherlands
Get access
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2025

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.)

Book purchase

Temporarily unavailable

References

References

Antonelli, M., Glaser, B., Teuling, A. J., Klaus, J. & Pfister, L. (2020). Saturated areas through the lens: 2. Spatio-temporal variability of streamflow generation and its relationship with surface saturation. Hydrological Processes, 34, 13331349.CrossRefGoogle Scholar
Barkle, G., Stenger, R., Moorhead, B. & Clague, J. (2021). The importance of the hydrological pathways in exporting nitrogen from grazed artificially drained land. Journal of Hydrology, 597. doi:10.1016/j.hydrol.2021.126218.Google Scholar
Batista, P. V. G., Fiener, P., Scheper, S. & Alewell, C. (2021). A conceptual model-based sediment connectivity assessment for patchy agricultural catchments. Hydrology and Earth System Sciences, Discussions. doi:10.5194/hess-2021-231.Google Scholar
Belter, D., Wwller, M. & Blume, T. (2020). Characterising hillslope-stream connectivity with a joint analysis of stream and groundwater levels. Hydrology and Earth System Sciences, 24, 57135744.Google Scholar
Bennett, G. L., Molnar, P., McArdell, B. W. & Burlando, P. (2014). A probabilistic sediment cascade model of sediment transfer in the Illgraben. Water Resources Research, 50, 12251244.CrossRefGoogle Scholar
Bennett, J. P. (1974). Concepts of mathematical modelling of sediment yield. Water Resources Research 10, 485492.CrossRefGoogle Scholar
Berkowitz, B. & Zehe, E. (2020). Surface water and groundwater: Unifying conceptualization and quantification of two ‘water worlds’. Hydrology and Earth System Sciences, 24, 18311858.CrossRefGoogle Scholar
Betson, R. P. & Marius, J. B. (1969). Source areas of storm runoff. Water Resources Research, 5, 574582.CrossRefGoogle Scholar
Blackburn, W. H. (1975). Factors influencing infiltration and sediment production of semiarid rangelands in Nevada. Water Resources Research, 11, 929937.CrossRefGoogle Scholar
Blampied, J., Carozza, J. M. & Antoine, J. M. (2018). Sediment connectivity in the high Pyrenees mountain range by 2013 flood analysis: role of surficial sediment storages. Geomorphologies-Relief Proccessus Environnement. 24, 389402.Google Scholar
Blume, T & van Meerveld, H. J. (2015). From hillslopes to stream: methods to investigate subsurface connectivity. Wiley Interdisciplinary Reviews – Water 2015, 2: 177198.CrossRefGoogle Scholar
Boardman, J., Vandaele, K., Evans, R. & Foster, I. D. L. (2019). Off-site impacts of soil erosion and runoff: Why connectivity is more important than erosion rates. Soil Use and Management, 35, 245256.CrossRefGoogle Scholar
Brouwer, J., Barker, A. P., Gaze, S. R., Valentin, C., Bromley, J. & Valentine, C. (1997). The role of surface water redistribution in an area of patterned vegetation in a semi-arid environment, south-west Niger. Journal of Hydrology, 198, 129.Google Scholar
Calvo-Cases, A., Arnau-Rosalen, E., Boix-Fayos, C., Estrany, J., Roxo, M. J. & Symeonakis. (2021). Eco-geomorphological connectivity and coupling interactions at hillslope scale in drylands: concepts and critical examples. Journal of Arid Environments, 186, doi:10.1016/jaridenv.2020.104418.CrossRefGoogle Scholar
Cawson, J. G., Sheridan, G. J., Smith, H. G. & Lane, P. N. J. (2013). Effects of fire severity and burn patchiness on hillslope-scale surface runoff, erosion and hydrologic connectivity in a prescribed burn. Forest Ecology and Management, 310, 219233.CrossRefGoogle Scholar
Cislaghi, A. & Bischetti, G. B. (2019). Source areas, connectivity, and delivery rate of sediments in mountainous forested hillslopes: A probabilistic approach. Science of the Total Environment, 652, 11681186.CrossRefGoogle ScholarPubMed
Clapuyt, F., Vanacker, V., Christi, M., Van Oost, K. & Schlunegger, F. (2019). Spatio-temporal dynamics of sediment transfer systems in landslide prone Alpine catchments. Solid Earth, 10, 14801503.CrossRefGoogle Scholar
Coulthard, T. J., Macklin, M. G. & Kirkby, M. J. (2002). A cellular model of Holocene upland river basin and alluvial fan evolution. Earth Surface Processes and Landforms 27, 269288.CrossRefGoogle Scholar
Deb, K., Miyazaki, T., Mizoguchi, M. & Seki, K. (2006). Return flow generating point on a hillslope layered with traffic pan. Transactions of Japanese Society of Irrigation, Drainage and Reclamation Engineering, 241, 111, doi:10.11408/jsidre1965.2006.1.Google Scholar
Di Stefano, C. & Ferro, V. (2017). Testing sediment connectivity at the experimental SPA2 basin, Sicily (Italy). Land Degradation & Development, 28, 15571567.CrossRefGoogle Scholar
Di Stefano, C. & Ferro, V. (2018). Modelling sediment delivery using connectivity components at the experimental SPA2 basin, Sicily (Italy). Journal of Mountain Science, 15, 18681880.CrossRefGoogle Scholar
Favis-Mortlock, D. T., Boardman, J., Parsons, A. J. & Lascelles, B. (2000). Emergence and erosion: a model for rill initiation and development. Hydrological Processes, 14, 21732205.3.0.CO;2-6>CrossRefGoogle Scholar
Ferguson, R. I. (1981). Channel form and channel changes. In Lewin, J. (ed.) British Rivers. London: Allen and Unwin, pp. 90125.Google Scholar
Germain, D., Gavril, I-G., Elizbarashvili, M. & Pop, O. M. (2020). Multidisciplinary approach to sediment connectivity between debris-flow channel network and the Doira River, Mazeri Valley, Southern Caucasus, Georgia. Geomorphology, 371, doi:10.1016/j/geomorph.2020.107455.CrossRefGoogle Scholar
Green, W. H. & Ampt, G. A. (1911). Studies on soil physics. 1. The flow of air and water through soils. Journal of Agricultural Soils, 4, 124.Google Scholar
Hancock, G. & Lowry, J. (2021). Quantifying the influence of rainfall, vegetation and animals on soil erosion and hillslope connectivity in the monsoonal tropics of northern Australia. Earth Surface Processes and Landforms, 46, 21102123.CrossRefGoogle Scholar
Hawkins, R. H. (1982). Interpretations of Source-area Variability in Rainfall-runoff Relations. Littleton, CO: Water Resources Publications, pp. 303324.Google Scholar
Hewett, C. J. M., Wainwright, J., Parsons, A. J. Cooper, J. R., Kitchener, B., Hargrave, G. K., Long, E. J., Onda, Y. & Patin, J. (2014). The importance of simulating changes in topography in process-based soil erosion modelling: implications for landscape evolution modelling. Geophysical Research Abstracts, 16, EGU2014-5422.Google Scholar
Hewlett, J. D. & Hibbert, A. R. (1967). Factors affecting the response of small watersheds to precipitationin humid areas. In 1965 International Symposium on Forest Hydrology, Pennsylvania State University, New York: Pergamon, pp. 275–90.Google Scholar
Horton, R. E. (1933). The role of infiltration in the hydrological cycle. Transactions of the American Geophysical Union, 14, 446460.Google Scholar
Howard, A. D. (1994). A detachment limited model for drainage basin evolution. Water Resources Research, 30, 22612285.CrossRefGoogle Scholar
Iwasaki, K., Katsuyama, M. & Tani, M. (2020). Factors affecting dominant peak-flow runoff generation mechanisms among five neighbouring granitic headwater catchments. Hydrological Processes, 34, 11541166.CrossRefGoogle Scholar
Jautzy, T., Maltaid, M. & Buffin-Belanger, T. (2021). Interannual evolution and hydrosedimenary connectivity induced by forest cover change in a snow-dominated mountainous catchment. Land Degradation and Development, 32, doi:10.1002/ldr.3902.CrossRefGoogle Scholar
Klaus, J. & Jackson, C. R. (2018). Interflow is not binary: a continuous perched layer does not imply continuous connectivity. Water Resources Research, 54, 59215932.CrossRefGoogle Scholar
Laine-Kaulio, H. & Koivusalo, H. (2018). Model-based exploration of hydrological connectivity and solute transport in a forested hillslope. Land Degradation and Development, 29, 11761189.CrossRefGoogle Scholar
Lopez-Vicente, M., Nadal-Romero, E. & Cammeraat, E. L. H. (2017). Hydrological connectivity does change over 70 years of abandonment and afforestation in the Spanish Pyrenees. Land Degradation and Development, 28, 12981310.CrossRefGoogle Scholar
Luk, S.-H., Abrahams, A. D. & Parsons, A. J. (1993). Sediment sources and sediment transport by rill flow and interrill flow on a semi-arid piedmont slope, Southern Arizona. Catena, 20, 93111.CrossRefGoogle Scholar
Madeiros, P. H. A., de Araujo, J. C., Marmede, G. L., Creutzfeldt, B., Guntner, A. & Bronstert, A. (2014). Connectivity of sediment transport in a semiarid environment: a synthesis for the Uppeer Jaguaribe Basin, Brazil. Journal of Soils and Sediments, 14, 19381948.CrossRefGoogle Scholar
Masselink, R. J. H., Temme, A. J. A. M., Gimenez, R., Casali, J. & Keesstra, S. D. (2017). Assessing hillslope-channel connectivity in an agricultural catchment using rare-earth oxide tracers and random forests models. Cuadernos de Investigacion Geografica, 43, 1939.Google Scholar
Michaelides, K., Lister, D., Wainwright, J., and Parsons, A. J. (2012), Linking runoff and erosion dynamics to nutrient fluxes in a degrading dryland landscape, J. Geophys. Res., 117, G00N15, doi:10.1029/2012JG002071.Google Scholar
Moreno-de-las-Heras, M., Merino-Martin, L., Saco, P. M., Espigares, T., Gallart, F. & Nicolau, J. M. (2020). Structural and functional control of surface-patch to hillslope runoff and sediment connectivity in Mediterranean dry reclaimed slope systems. Hydrology and Erath System Sciences, 24, doi:10.5194/hess-24-2855-2020. Southern Arizona.Google Scholar
Mueller, E. N., Wainwright, J. & Parsons, A. J. (2007). Impact of connectivity on the modelling of overland flow within semi-arid shrubland environments. Water Resources Research, 43, W09412.CrossRefGoogle Scholar
Murabayashi, E. T. & Fox, Y.-S. (1979). Urbanization-induced impacts on infiltration capacity and rainfall-runoff relation in Hawaiian urban area. Technical Report 27, Water Resources Research Center, Honolulu, Hawaii.Google Scholar
Nanda, A., Sen, S. & McNamara, J. P. (2019). How spatiotemporal variation in soil moisture can explain hydrological connectivity of infiltration-excess dominated hillslope: observations from lesser Himalyan landscapes. Journal of Hydrology, 579, doi:10.1016/j.hydrol,2019.124146.Google Scholar
Nunes, J. P., Wainwright, J., Bielders, C. L., Darboux, F., Fiener, P., Finger, D. & Turnbull, L. (2018). Better models are more effectively connected models. Earth Surface Processes and Landforms, 43, 13551360. doi:10.1002/esp.4323.Google Scholar
Parsons, A. J., Wainwright, J. & Abrahams, A. D. (1996). Runoff and erosion on semi-arid hillslopes. In Anderson, M. G. & Brooks, S. M. (eds.) Advances in Hillslope Processes. Chichester: John Wiley & Sons, pp. 10611078.Google Scholar
Parsons, A. J., Stromberg, S. G. L. & Greener, M. (1998). Sediment-transport competence of rain-impacted interrill overland flow. Earth Surface Processes and Landforms, 23, 365375.3.0.CO;2-6>CrossRefGoogle Scholar
Parsons, A. J. & Wainwright, J. (2006). Depth distribution of interrill overland flow and the formation of rills. Hydrological Processes, 20, 15111523.CrossRefGoogle Scholar
Parsons, A. J., Wainwright, J., Powell, D. M., Brazier, R. E. (2006). Is sediment delivery a fallacy? Earth Surface Processes and Landforms 31, 13251328.CrossRefGoogle Scholar
Parsons, A. J. (2012). How useful are catchment sediment budgets? Progress in Physical Geography, 36, 6071.CrossRefGoogle Scholar
Parsons, A., Cooper, J., Onda, Y., Patin, J. & Wainwright, J. (2015). Conceptualizing hillslope sediment connectivity as virtual velocity. Geophysical Research Abstracts, 17, EGU2015-1444.Google Scholar
Parsons, A. J., Wainwright, J., Abrahams, A. D. & Simanton, J. R. (2016) Distributed dynamic modelling of interrill overland flow. Hydrological Processes, 11, 18331859.3.0.CO;2-7>CrossRefGoogle Scholar
Parsons, A. J., Cooper, J. R., Wainwright, J. & Sekiguchi, T. (2018). Virtual velocity of sand transport in water. Earth Surface Processes and Landforms, 43, 755761.CrossRefGoogle Scholar
Phillips, R. W., Spence, C. & Pomeroy, J. W. (2011). Connectivity and runoff dynamics in heterogeneous basins. Hydrological Processes, 25, 30613075.CrossRefGoogle Scholar
Pechenick, A. M., Rizzo, D. M., Morrissey, L. A., Garvey, K. M., Underwood, K. L. & Wemple, B. C. (2014). A multi-scale statistical approach to assess the effects of connectivity of road and stream networks on geomorphic channel condition. Earth Surface Processes and Landforms, 39, 15381549.CrossRefGoogle Scholar
Philip, J. R. (1957/8). The theory of infiltration. Soil Science, 83, 345357 and 435–458.CrossRefGoogle Scholar
Puigdefabregas, J. (2005). The role of vegetation patterns in structuring runoff and sediment fluxes in drylands. Earth Surface Processes and Landforms, 30, 133147.CrossRefGoogle Scholar
Reaney, S. M. 2003. Modelling runoff generation and connectivity for semiarid hillslopes and small catchments. PhD thesis, University of Leeds.Google Scholar
Reaney, S. M., Bracken, L. J. & Kirkby, M. J. (2007). Use of the Connectivity of Runoff Model (CRUM) to investigate the influence of storm characteristics on runoff generation and connectivity in semi-arid areas. Hydrological Processes, 21, 894906.CrossRefGoogle Scholar
Reulier, R., Delahaye, D., Viel, V. & Davidson, R. (2017). Hydro-sedimentary connectivity in a small agricultural watershed in French northwest: from field expertise to multi-agent system modelling. Geomorphologie-Relief Processus Environnement, 23, 327340.CrossRefGoogle Scholar
Sharma, M. L., Gander, G. A. & Hunt, C. G. (1980). Spatial variability of infiltration in a watershed. Journal of Hydrology, 45, 101122.CrossRefGoogle Scholar
Shook, K., Papalexiou, S. & Pomeroy, J. W. (2021). Quantifying the effects of Prairie depressional storage complexes on drainage basin connectivity. Journal of Hydrology 593, doi:10.1016/j.jhydrol.2020.125846.CrossRefGoogle Scholar
Sun, W. Y., Mu, X. M., Gao, P., Zhao, G. J., Li, J. Y., Zhang, Y. Q. & Francis, C. (2019). Landscape patches influencing hillslope erosion processes and flow dynamics. Geoderma, 353, 391400.CrossRefGoogle Scholar
Wainwright, J & Parsons, A. J. (2002). The effect of temporal variations in rainfall on scale dependency in runoff coefficients. Water Resources Research, 38(12), 7-1–7-10.CrossRefGoogle Scholar
Wainwright, J., Parsons, A. J., Mueller, E., Brazier, R., Powell, D. M. & Fenti, B. (2008). A transport-distance approach to scaling erosion rates: 1. Background and model development. Earth Surface Processes and Landforms, 33, 813826.CrossRefGoogle Scholar
Wainwright, J., Turnbull, L., Ibrahim, T. G., Lexartza-Artza, I., Thornton, S. F. & Brazier, R. E. (2011). Linking environmental régimes, space and time: interpretations of structural and functional connectivity. Geomorphology, 126, 387404.CrossRefGoogle Scholar
Weihrauch, C., Weber, C. J. & von Sperber, C. (2021). A Soilscaple network approach (SNAp) to investigate subsurface phosphorus translocation along slopes. Science of the Total Environment, 784, doi:10.1016/j.scitotenv.2021.147131.Google ScholarPubMed
Willgoose, G., Bras, R. L. & Rodriguez-Iturbe, I. (1991). A coupled channel network growth and hillslope evolution model 1. Theory. Water Resources. Research, 27, 16711684.CrossRefGoogle Scholar
Wilson, G., Rigby, J. R., Ursic, M. & Dabney, S. M. (2016). Soil pipe flow tracer experiments: 1. Connectivity and transport characteristics. Hydrological Processes, 30, 12651279.CrossRefGoogle Scholar
Wolman, M. G. & Miller, J. P. (1960). Magnitude and frequency of forces in geomorphic processes. Journal of Geology, 68, 5474.CrossRefGoogle Scholar
Wolstenholme, J. M., Smith, M. W., Baird, A. J. & Sim, T. G. (2020). A new approach for measuring surface hydrological connectivity. Hydrological Processes, 34, 538552.CrossRefGoogle Scholar
Young, A. (1972). Slopes. Edinburgh: Oliver & Boyd.Google Scholar
Zimmermann, B. Zimmermann, A., Turner, B. L., Francke, T. & Elsenbeer, H. (2014). Connectivity of overland flow by drainage network expansion in a rain forest catchment. Water Resources Research, 50, 14571473.CrossRefGoogle Scholar
Zuecco, G., Rinderer, M., Penna, D., Borga, M. & van Meerveld, H. J. (2019). Quantification of subsurface hydrologic connectivity in four headwater catchments using graph theory. Science of the Total Environment, 646, 12651280.CrossRefGoogle ScholarPubMed

References

Al Farraj, A., & Harvey, A. (2010). Influence of hillslope-to-channel and tributary-junction coupling on channel morphology and sediments: Bowderdale Beck, Howgill Fells, NW England. Zeitschrift Fur Geomorphologie, 54(2), 203224. doi:10.1127/0372-8854/2010/0054-0018CrossRefGoogle Scholar
Arnold, J. G., White, M. J., Allen, P. M., Gassman, P. W., & Bieger, K. (2021). Conceptual framework of connectivity for a national agroecosystem model based on transport processes and management practices. Journal of the American Water Resources Association, 57(1), 154169.CrossRefGoogle Scholar
Blanpied, J., Carozza, J. M., & Antoine, J. M. (2018). Sediment connectivity in the high Pyrenees mountain range by 2013 flood analysis: role of surficial sediment storages. Geomorphologie-Relief Processus Environnement, 24(4), 389402. doi:10.4000/geomorphologie.12718CrossRefGoogle Scholar
Bombino, G., Boix-Fayos, C., Cataldo, M. F., D’Agostino, D., Denisi, P., de Vente, J., Labate, A., & Zema, D. A. (2020). A modified catchment connectivity index for applications in semi-arid torrents of the Mediterranean environment. River Research and Applications, 36 (5). https://doi.org/10.1002/rra.3606CrossRefGoogle Scholar
Borges, A. V., Darchambeau, F., Lambert, T., Morana, C., Allen, G. H., Tambwe, E., … Bouillon, S. (2019). Variations in dissolved greenhouse gases (CO2, CH4, N2O) in the Congo River network overwhelmingly driven by fluvial-wetland connectivity. Biogeosciences, 16(19), 38013834.CrossRefGoogle Scholar
Borselli, L., Cassi, P., & Torri, D. (2008). Prolegomena to sediment and flow connectivity in the landscape: a GIS and field numerical assessment. Catena, 75(3), 268277.CrossRefGoogle Scholar
Bracken, L. J., Turnbull, L., Wainwright, J., & Bogaart, P. (2015). Sediment connectivity: a framework for understanding sediment transfer at multiple scales. Earth Surface Processes and Landforms, 40(2), 177188. doi:10.1002/esp.3635CrossRefGoogle Scholar
Brierley, G. J., & Fryirs, K. A. (2005). Geomorphology and River Management. Blackwell.Google Scholar
Brierley, G., & Fryirs, K. (2009). Don’t fight the site: three geomorphic considerations in catchment-scale river rehabilitation planning. Environmental Management, 43(6), 12011218. doi:10.1007/s00267-008-9266-4CrossRefGoogle ScholarPubMed
Brierley, G., Fryirs, K., & Jain, V. (2006). Landscape connectivity: the geographic basis of geomorphic applications. Area, 38(2), 165174. doi:10.1111/j.1475-4762.2006.00671.xCrossRefGoogle Scholar
Brierley, G. J., & Fryirs, K. (1999). Tributary-trunk stream relations in a cut-and-fill landscape: a case study from Wolumla catchment, New South Wales, Australia. Geomorphology, 28(1–2), 6173. doi:10.1016/s0169-555x(98)00103-2CrossRefGoogle Scholar
Brierley, G., Tunnicliffe, J., Bizzi, S., Lee, F. Perry, G., Pöppl, R., & Fryirs, K. (2022). Quantifying sediment (Dis)connectivity in the modeling of river systems. Treatise on Geomorphology (2nd ed.) 10, 206224.CrossRefGoogle Scholar
Brunsden, D, & Thornes, J. B. (1979). Landscape sensitivity and change. Transactions of the Institute of British Geographers, NS4, 463484.CrossRefGoogle Scholar
Brunsden, D. (1993). Barriers to geomorphological change. In Thomas, D. S. G, & Allison, R. J. (eds), Landscape Sensitivity. Chichester: John Wiley & Sons, pp. 712.Google Scholar
Cabre, A., Remy, D., Aguilar, G., Carretier, S., & Riquelme, R. (2020). Mapping rainstorm erosion associated with an individual storm from InSAR coherence loss validated by field evidence for the Atacama Desert. Earth Surface Processes and Landforms, 45(9), 20912106. doi:10.1002/esp.4868CrossRefGoogle Scholar
Cammeraat, L. H. (2002). A review of two strongly contrasting geomorphological systems within the context of scale. Earth Surface Processes and Landforms, 27, 12011222.CrossRefGoogle Scholar
Cavalli, M., Trevisani, S., Comiti, F., & Marchi, L. (2013). Geomorphometric assessment of spatial sediment connectivity in small Alpine catchments. Geomorphology, 188, 3141.CrossRefGoogle Scholar
Chartin, C., Evrard, O., Laceby, J. P., Onda, Y., Ottl’e, C., Lef‘evre, I., & Cerdan, O. (2017). The impact of typhoons on sediment connectivity: lessons learnt from contaminated coastal catchments of the Fukushima Prefecture (Japan). Earth Surface Processes and Landforms, 42 (2). https://doi.org/10.1002/esp.4056CrossRefGoogle Scholar
Chen, Y., & Wang, Y. G. (2019). Changes in river connectivity indexes in the lower Yellow River between 1960 and 2015. River Research and Applications, 35(9), 13771386.CrossRefGoogle Scholar
Chiverrell, R. C., Foster, G. C., Marshall, P., Harvey, A. M., & Thomas, G. S. P. (2009). Coupling relationships: Hillslope-fluvial linkages in the Hodder catchment, NW England. Geomorphology, 109 (3–4), 222235. doi:10.1016/j.geomorph.2009.03.004CrossRefGoogle Scholar
Chiverrell, R. C., Foster, G. C., Thomas, G. S. P., & Marshall, P. (2010). Sediment transmission and storage: the implications for reconstructing landform development. Earth Surface Processes and Landforms, 35(1), 415. doi:10.1002/esp.1806CrossRefGoogle Scholar
Cienciala, P., Nelson, A. D., Haas, A. D., & Xu, Z. W. (2020). Lateral geomorphic connectivity in a fluvial landscape system: unraveling the role of confinement, biogeomorphic interactions, and glacial legacies. Geomorphology, 354. doi:10.1016/j.geomorph.2020.107036CrossRefGoogle Scholar
Cislaghi, A., Bischetti, G. B., 2019. Source areas, connectivity, and delivery rate of sediments in mountainous-forested hillslopes: a probabilistic approach. Science of the Total Environment, 652. https://doi.org/10.1016/j.scitotenv.2018.10.318Google ScholarPubMed
Citterio, A., & Piegay, H. (2009). Overbank sedimentation rates in former channel lakes: characterization and control factors. Sedimentology, 56(2): 461482.CrossRefGoogle Scholar
Cossart, É., & Fressard, M. (2017). Assessment of structural sediment connectivity within catchments: Insights from graph theory. Earth Surface Dynamics, 5(2). https://doi.org/10.5194/esurf-5-253-2017CrossRefGoogle Scholar
Coulthard, T. J., & Van De Wiel, M. J. (2017). Modelling long term basin scale sediment connectivity, driven by spatial land use changes. Geomorphology, 277, 265281. doi:10.1016/j.geomorph.2016.05.027CrossRefGoogle Scholar
Covino, T. (2017). Hydrologic connectivity as a framework for understanding biogeochemical flux through watersheds and along fluvial networks. Geomorphology, 277, 133144. doi:10.1016/j.geomorph.2016.09.030CrossRefGoogle Scholar
Crema, S., & Cavalli, M. (2018). SedInConnect: a stand-alone, free and open source tool for the assessment of sediment connectivity. Computers and Geosciences, 111, 3945.CrossRefGoogle Scholar
Croke, J., Fryirs, K., & Thompson, C. (2013). Channel-floodplain connectivity during an extreme flood event: implications for sediment erosion, deposition, and delivery. Earth Surface Processes and Landforms, 38(12), 14441456. doi:10.1002/esp.3430CrossRefGoogle Scholar
Czuba, J. A., & Foufoula-Georgiou, E. (2015). Dynamic connectivity in a fluvial network for identifying hotspots of geomorphic change. Water Resources Research, 51(3): 14011421.CrossRefGoogle Scholar
de Souza, J. O. P., & Correa, A. C. D. (2020). Evolution scenarios of landscape connectivity in semiarid environment – Saco Creek watershed, Serra Talhada/pe – Brazil. Revista Brasileira De Geomorfologia, 21(1), 6377. doi:10.20502/rbg.v21i1.1529Google Scholar
Dezso, J., Loczy, D., Salem, A. M., & Nagy, G. (2019). Floodplain connectivity. In Loczy, D. (ed.), Drava River: Environmental Problems and Solutions, pp. 215230.CrossRefGoogle Scholar
D’Haen, K., Dusar, B., Verstraeten, G., Degryse, P., & De Brue, H. (2013). A sediment fingerprinting approach to understand the geomorphic coupling in an eastern Mediterranean mountainous river catchment. Geomorphology, 197, 6475.CrossRefGoogle Scholar
Diaz-Redondo, M., Egger, G., Marchamalo, M., Damm, C., de Oliveira, R. P., & Schmitt, L. (2018). Targeting lateral connectivity and morphodynamics in a large river-floodplain system: the upper Rhine River. River Research and Applications, 34(7), 734744. doi:10.1002/rra.3287CrossRefGoogle Scholar
Ferguson, R. I. (1981). Channel forms and channel changes. In Lewin, J. (ed.), British Rivers. London: Allen and Unwin, pp. 90125.Google Scholar
Fressard, M., Cossart, E., 2019. A graph theory tool for assessing structural sediment connectivity: Development and application in the Mercurey vineyards (France). The Science of the Total Environment, 651, 25662584.CrossRefGoogle ScholarPubMed
Fritz, K. M., Schofield, K. A., Alexander, L. C., McManus, M. G., Golden, H. E., Lane, C. R., … Pollard, A. I. (2018). Physical and chemical connectivity of streams and riparian wetlands to downstream waters: a synthesis. Journal of the American Water Resources Association, 54(2), 323345. doi:10.1111/1752-1688.12632CrossRefGoogle ScholarPubMed
Fryirs, K. (2013). (Dis)Connectivity in catchment sediment cascades: a fresh look at the sediment delivery problem. Earth Surface Processes and Landforms, 38(1), 3046. doi:10.1002/esp.3242CrossRefGoogle Scholar
Fryirs, K., & Brierley, G. J. (1999). Slope-channel decoupling in Wolumla catchment, New South Wales, Australia: the changing nature of sediment sources following European settlement. Catena, 35(1), 4163. doi:10.1016/s0341-8162(98)00119-2CrossRefGoogle Scholar
Fryirs, K., & Brierley, G. J. (2001). Variability in sediment delivery and storage along river courses in Bega catchment, NSW, Australia: implications for geomorphic river recovery. Geomorphology, 38(3–4), 237265.CrossRefGoogle Scholar
Fryirs, K and Gore, D., 2013. Sediment tracing in the upper Hunter catchment using elemental and mineralogical compositions: implications for catchment-scale suspended sediment (dis)connectivity and management. Geomorphology, 193, 112121.CrossRefGoogle Scholar
Fryirs, K. A., Brierley, G. J., Preston, N. J., & Kasai, M. (2007a). Buffers, barriers and blankets: the (dis)connectivity of catchment-scale sediment cascades. Catena, 70(1), 4967. doi:10.1016/j.catena.2006.07.007CrossRefGoogle Scholar
Fryirs, K. A., Brierley, G. J., Preston, N. J., & Spencer, J. (2007b). Catchment-scale (dis)connectivity in sediment flux in the upper Hunter catchment, New South Wales, Australia. Geomorphology, 84(3–4), 297316. doi:10.1016/j.geomorph.2006.01.044CrossRefGoogle Scholar
Fuller, I. C., & Death, R. G. (2018). The science of connected ecosystems: What is the role of catchment-scale connectivity for healthy river ecology? Land Degradation & Development, 29(5), 14131426. doi:10.1002/ldr.2903CrossRefGoogle Scholar
Gao, P., & Zhang, Z. R. (2016). Spatial patterns of sediment dynamics within a medium-sized watershed over an extreme storm event. Geomorphology, 267, 2536. doi:10.1016/j.geomorph.2016.05.025CrossRefGoogle Scholar
Gilbert, J. T., & Wilcox, A. C. (2020). Sediment routing and floodplain exchange (SeRFE): a spatially explicit model of sediment balance and connectivity through river networks. Journal of Advances in Modeling Earth Systems, 12(9). doi:10.1029/2020ms002048CrossRefGoogle Scholar
Godfrey, A. E., Everitt, B. L., & Duque, J. F. M. (2008). Episodic sediment delivery and landscape connectivity in the Mancos Shale badlands and Fremont River system, Utah, USA. Geomorphology, 102(2), 242251. doi:10.1016/j.geomorph.2008.05.002CrossRefGoogle Scholar
Gootman, K. S., Gonzalez-Pinzon, R., Knapp, J. L. A., Garayburu-Caruso, V., & Cable, J. E. (2020). Spatiotemporal variability in transport and reactive processes across a first- to fifth-order fluvial network. Water Resources Research, 56(5). doi:10.1029/2019wr026303CrossRefGoogle Scholar
Gran, K. B., & Czuba, J. A. (2017). Sediment pulse evolution and the role of network structure. Geomorphology, 277, 1730. doi:10.1016/j.geomorph.2015.12.015CrossRefGoogle Scholar
Grauso, S., Pasanisi, F., & Tebano, C. (2018). Assessment of a simplified connectivity index and specific sediment potential in river basins by means of geomorphometric tools. Geosciences, 8(2). doi:10.3390/geosciences8020048CrossRefGoogle Scholar
Gumiero, B., Mant, J., Hein, T., Elso, J., & Boz, B. (2013). Linking the restoration of rivers and riparian zones/wetlands in Europe: sharing knowledge through case studies. Ecological Engineering, 56, 3650. doi:10.1016/j.ecoleng.2012.12.103CrossRefGoogle Scholar
Harries, R. M., Gailleton, B., Kirstein, L. A., Attal, M., Whittaker, A. C., & Mudd, S. M. (2021). Impact of climate on landscape form, sediment transfer and the sedimentary record. Earth Surface Processes and Landforms, 46(5), 9901006. doi:10.1002/esp.5075CrossRefGoogle Scholar
Harvey, A. M. (1996). The role of alluvial fans in the mountain fluvial systems of southeast Spain: implications of climatic change. Earth Surface Processes and Landforms, 21(6), 543553.3.0.CO;2-F>CrossRefGoogle Scholar
Harvey, A. M. (1997). Coupling between hillslope gully systems and stream channels in the Howgill Fells, northwest England: temporal implications. Geomorphologie: Relief, Processus, Environnement, 1, 320.CrossRefGoogle Scholar
Harvey, A. M. (2002). Effective timescales of coupling within fluvial systems. Geomorphology, 44(3–4), 175201. doi:10.1016/s0169-555x(01)00174-xCrossRefGoogle Scholar
Harvey, A. M. (2007). Differential recovery from the effects of a 100-year storm: Significance of long-term hillslope-channel coupling; Howgill Fells, northwest England. Geomorphology, 84(3–4), 192208. doi:10.1016/j.geomorph.2006.03.009CrossRefGoogle Scholar
Harvey, AM. (2012). The coupling status of alluvial fans and debris cones: a review and synthesis. Earth Surface Processes and Landforms, 37, 6476.CrossRefGoogle Scholar
Heckmann, T., Cavalli, M., Cerdan, O., Foerster, S., Javaux, M., Lode, E., … Brardinoni, F. (2018). Indices of sediment connectivity: opportunities, challenges and limitations. Earth-Science Reviews, 187, 77108. doi:10.1016/j.earscirev.2018.08.004CrossRefGoogle Scholar
Heckmann, T., & Schwanghart, W. (2013). Geomorphic coupling and sediment connectivity in an alpine catchment – exploring sediment cascades using graph theory. Geomorphology, 182, 89103. doi:10.1016/j.geomorph.2012.10.033CrossRefGoogle Scholar
Heritage, G., & Entwistle, N. (2020). Impacts of river engineering on river channel behaviour: implications for managing downstream flood risk. Water, 12(5). doi:10.3390/w12051355CrossRefGoogle Scholar
Hohensinner, S., Jungwirth, M., Muhar, S., & Schmutz, S. (2014). Importance of multi-dimensional morphodynamics for habitat evolution: Danube River 1715–2006. Geomorphology, 215, 319. doi:10.1016/j.geomorph.2013.08.001CrossRefGoogle Scholar
Hooke, J. (2003). Coarse sediment connectivity in river channel systems: a conceptual framework and methodology. Geomorphology, 56(1–2), 7994. doi:10.1016/s0169-555x(03)00047-3CrossRefGoogle Scholar
Hooke, J. M., (2004). Analysis of coarse sediment connectivity in semi-arid river channels. In IAHS Publn 288, Sediment Transfer through the Fluvial System, Moscow Conference August 2004, pp. 269–275.Google Scholar
Hooke, J. M. (2006). Human impacts on fluvial systems in the Mediterranean region. Geomorphology, 79(3–4), 311335. doi:10.1016/j.geomorph.2006.06.036CrossRefGoogle Scholar
Hooke, J., & Sandercock, P. (2012). Use of vegetation to combat desertification and land degradation: recommendations and guidelines for spatial strategies in Mediterranean lands. Landscape and Urban Planning, 107, 389400.CrossRefGoogle Scholar
Hooke, J. M., & Sandercock, P. J. (eds.) (2017). Combating Desertification and Land Degradation: Spatial Strategies Using Vegetation. Springer Briefing, 110 pp. Cham: Springer.CrossRefGoogle Scholar
Hooke, J., & Souza, J. (2021). Challenges of mapping, modelling and quantifying sediment connectivity. Earth-Science Reviews, 223. doi:10.1016/j.earscirev.2021.103847CrossRefGoogle Scholar
Hudson, P. F., Sounny-Slittine, M. A., & LaFevor, M. (2013). A new longitudinal approach to assess hydrologic connectivity: Embanked floodplain inundation along the lower Mississippi River. Hydrological Processes, 27(15), 21872196. doi:10.1002/hyp.9838CrossRefGoogle Scholar
Jacobson, R. B., Janke, T. P., & Skold, J. J. (2011). Hydrologic and geomorphic considerations in restoration of river-floodplain connectivity in a highly altered river system, Lower Missouri River, USA. Wetlands Ecology and Management, 19(4), 295316.CrossRefGoogle Scholar
Jain, V., & Tandon, S. K. (2010). Conceptual assessment of (dis)connectivity and its application to the Ganga River dispersal system. Geomorphology, 118(3–4), 349358. doi:10.1016/j.geomorph.2010.02.002CrossRefGoogle Scholar
James, L. A., Monohan, C., & Ertis, B. (2019). Long-term hydraulic mining sediment budgets: Connectivity as a management tool. Science of the Total Environment, 651, 20242035. doi:10.1016/j.scitotenv.2018.09.358CrossRefGoogle ScholarPubMed
Joyce, H. M., Hardy, R. J., Warburton, J., & Large, A. R. G. (2018). Sediment continuity through the upland sediment cascade: geomorphic response of an upland river to an extreme flood event. Geomorphology, 317, 4561. doi:10.1016/j.geomorph.2018.05.002CrossRefGoogle Scholar
Kalantari, Z., Ferreira, C. S. S., Koutsouris, A. J., Ahmer, A. K., Cerd‘a, A., Destouni, G. (2019). Assessing flood probability for transportation infrastructure based on catchment characteristics, sediment connectivity and remotely sensed soil moisture. Science of the Total Environment, 661. https://doi.org/10.1016/j.scitotenv.2019.01.009.CrossRefGoogle ScholarPubMed
Khan, S., Fryirs, K., & Bizzi, S. (2021). Modelling sediment (dis)connectivity across a river network to understand locational-transmission-filter sensitivity for identifying hotspots of potential geomorphic adjustment. Earth Surface Processes and Landforms. doi:10.1002/esp.5213CrossRefGoogle Scholar
Knighton, D. (1998). Fluvial Forms and Processes: A New Perspective. Arnold.Google Scholar
Kuo, C. W., & Brierley, G. J. (2013). The influence of landscape configuration upon patterns of sediment storage in a highly connected river system. Geomorphology, 180, 255266. doi:10.1016/j.geomorph.2012.10.015CrossRefGoogle Scholar
Lane, S. N., Bakker, M., Gabbud, C., Micheletti, N., & Saugy, J. N. (2017). Sediment export, transient landscape response and catchment-scale connectivity following rapid climate warming and Alpine glacier recession. Geomorphology, 277, 210227. doi:10.1016/j.geomorph.2016.02.015CrossRefGoogle Scholar
Lane, S. N., Reid, S. C., Tayefi, V., Yu, D., & Hardy, R. J. (2008). Reconceptualising coarse sediment delivery problems in rivers as catchment-scale and diffuse. Geomorphology, 98(3–4), 227249. doi:10.1016/j.geomorph.2006.12.028CrossRefGoogle Scholar
Lehotsky, M., Rusnak, M., Kidova, A., & Dudzak, J. (2018). Multitemporal assessment of coarse sediment connectivity along a braided-wandering river. Land Degradation & Development, 29(4), 12491261. doi:10.1002/ldr.2870CrossRefGoogle Scholar
Lewin, J., & Ashworth, P. J. (2014). Defining large river channel patterns: Alluvial exchange and plurality. Geomorphology, 215, 8398. doi:10.1016/j.geomorph.2013.02.024CrossRefGoogle Scholar
Lexartza-Artza, I., & Wainwright, J. (2011). Making connections: changing sediment sources and sinks in an upland catchment. Earth Surface Processes and Landforms, 36(8). https://doi.org/10.1002/esp.2134.CrossRefGoogle Scholar
Li, B. W., Yang, Z. F., Cai, Y. P., & Li, B. (2021). The frontier evolution and emerging trends of hydrological connectivity in river systems: a scientometric review. Frontiers of Earth Science, 15(1), 8193. doi:10.1007/s11707-020-0852-yCrossRefGoogle Scholar
Lisenby, P. E., & Fryirs, K. A. (2017). Sedimentologically significant tributaries: catchment-scale controls on sediment (dis) connectivity in the Lockyer Valley, SEQ, Australia. Earth Surface Processes and Landforms, 42(10), 14931504. doi:10.1002/esp.4130CrossRefGoogle Scholar
Lisenby, P. E., Fryirs, K. A., & Thompson, C. J. (2020). River sensitivity and sediment connectivity as tools for assessing future geomorphic channel behavior. International Journal of River Basin Management, 18(3), 279293. doi:10.1080/15715124.2019.1672705CrossRefGoogle Scholar
Liu, Y. L., Cui, B. S., Du, J. Z., Wang, Q., Yu, S. L., & Yang, W. (2020). A method for evaluating the longitudinal functional connectivity of a river-lake-marsh system and its application in China. Hydrological Processes, 34(26), 52785297. doi:10.1002/hyp.13946CrossRefGoogle Scholar
López-Vicente, M., Ben-Salem, N., 2019. Computing structural and functional flow and sediment connectivity with a new aggregated index: a case study in a large Mediterranean catchment. Science of the Total Environment, 651. https://doi.org/10.1016/j.scitotenv.2018.09.170CrossRefGoogle Scholar
Magilligan, F. J., Roberts, M. O., Marti, M., & Renshaw, C. E. (2021). The impact of run-of-river dams on sediment longitudinal connectivity and downstream channel equilibrium. Geomorphology, 376. doi:10.1016/j.geomorph.2020.107568CrossRefGoogle Scholar
Mahoney, D. T., Fox, J. F., al Aamery, N. (2018). Watershed erosion modeling using the probability of sediment connectivity in a gently rolling system. Journal of Hydrology, 561. https://doi.org/10.1016/j.jhydrol.2018.04.034CrossRefGoogle Scholar
Mahoney, D., Blandford, B., & Fox, J. (2021). Coupling the probability of connectivity and RUSLE reveals pathways of sediment transport and soil loss rates for forest and reclaimed mine landscapes. Journal of Hydrology, 594. doi:10.1016/j.jhydrol.2021.125963CrossRefGoogle Scholar
Mahoney, D. T., Fox, J., Al-Aamery, N., & Clare, E. (2020). Integrating connectivity theory within watershed modelling part I: Model formulation and investigating the timing of sediment connectivity. Science of the Total Environment, 740. doi:10.1016/j.scitotenv.2020.140385Google ScholarPubMed
Marcal, M., Brierley, G., & Lima, R. (2017). Using geomorphic understanding of catchment-scale process relationships to support the management of river futures: Macae Basin, Brazil. Applied Geography, 84, 2341. doi:10.1016/j.apgeog.2017.04.008CrossRefGoogle Scholar
Marchamalo, M., Hooke, J. M., & Sandercock, P. J. (2016). Flow and sediment connectivity in semi-arid landscapes in SE Spain: patterns and controls. Land Degradation and Development, 27(4), 10321044.CrossRefGoogle Scholar
Marteau, B., Gibbins, C., Vericat, D., & Batalla, R. J. (2020). Geomorphological response to system-scale river rehabilitation I: sediment supply from a reconnected tributary. River Research and Applications, 36(8), 14881503. doi:10.1002/rra.3683CrossRefGoogle Scholar
Mather, A. E., & Stokes, M. (2018). Bedrock structural control on catchment-scale connectivity and alluvial fan processes, High Atlas Mountains, Morocco. In Ventra, D., & Clarke, L. E. (eds.), Geology and Geomorphology of Alluvial and Fluvial Fans: Terrestrial and Planetary Perspectives, vol. 440, pp. 103128.CrossRefGoogle Scholar
Maxwell, C. M., Fernald, A. G., Cadol, D., Faist, A. M., & King, J. P. (2021). Managing flood flow connectivity to landscapes to build buffering capacity to disturbances: An ecohydrologic modeling framework for drylands. Journal of Environmental Management, 278. doi:10.1016/j.jenvman.2020.111486CrossRefGoogle ScholarPubMed
Mayor, Á. G., Bautista, S., Small, E. E., Dixon, M., & Bellot, J. (2008). Measurement of the connectivity of runoff source areas as determined by vegetation pattern and topography: a tool for assessing potential water and soil losses in drylands. Water Resources Research, 44, W10423.CrossRefGoogle Scholar
Messenzehl, K., Hoffmann, T., & Dikau, R. (2014). Sediment connectivity in the high-alpine valley of Val Müschauns, Swiss National Park – linking geomorphic field mapping with geomorphometric modelling. Geomorphology, 221. https://doi.org/10.1016/j.geomorph.2014.05.033CrossRefGoogle Scholar
Miller, J. R., Lord, M. L., Villarroel, L. F., Germanoski, D., & Chambers, J. C. (2012). Structural organization of process zones in upland watersheds of central Nevada and its influence on basin connectivity, dynamics, and wet meadow complexes. Geomorphology, 139, 384402. doi:10.1016/j.geomorph.2011.11.004CrossRefGoogle Scholar
Miller, J. R., Mackin, G., Lechler, P., Lord, M., & Lorentz, S. (2013). Influence of basin connectivity on sediment source, transport, and storage within the Mkabela Basin, South Africa. Hydrology and Earth System Sciences, 17(2), 761781. doi:10.5194/hess-17-761-2013CrossRefGoogle Scholar
Mishra, K., Sinha, R., Jain, V., Nepal, S., & Uddin, K. (2019). Towards the assessment of sediment connectivity in a large Himalayan river basin. Science of the Total Environment, 661, 251265. doi:10.1016/j.scitotenv.2019.01.118CrossRefGoogle Scholar
Najafi, S., Dragovich, D., Heckmann, T., & Sadeghi, S. H. (2021). Sediment connectivity concepts and approaches. Catena, 196. doi:10.1016/j.catena.2020.104880CrossRefGoogle Scholar
Nicholas, A.P., Ashworth, P.G., Kirkby, M., Macklin, M.G. & Murray, T. (1995). Sediment slugs: large-scale fluctuations in fluvial sediment transport rates and storage volumes. Progress in Physical Geography Earth and Environment, 19, 500519. doi:10.1177/030913339501900404CrossRefGoogle Scholar
Nicoll, T., & Brierley, G. (2017). Within-catchment variability in landscape connectivity measures in the Garang catchment, upper Yellow River. Geomorphology, 277, 197209. doi:10.1016/j.geomorph.2016.03.014CrossRefGoogle Scholar
Oldknow, C. J., & Hooke, J. M. (2017). Alluvial terrace development and changing landscape connectivity in the Great Karoo, South Africa. Insights from the Wilgerbosch River catchment, Sneeuberg. Geomorphology, 288, 1238. doi:10.1016/j.geomorph.2017.03.009CrossRefGoogle Scholar
Ondrackova, L., & Macka, Z. (2019). Geomorphic (dis)connectivity in a middle-mountain context: human interventions in the landscape modify catchment-scale sediment cascades. Area, 51(1), 113125. doi:10.1111/area.12424CrossRefGoogle Scholar
Parida, S., Singh, V., & Tandon, S. K. (2019). Sediment connectivity and evolution of gravel size composition in Dehra Dun – an Intermontane Valley in the frontal zone of NW Himalaya. Zeitschrift Fur Geomorphologie, 62(2), 83105. doi:10.1127/zfg/2019/0568CrossRefGoogle Scholar
Park, E. (2020). Characterizing channel-floodplain connectivity using satellite altimetry: mechanism, hydrogeomorphic control, and sediment budget. Remote Sensing of Environment, 243. doi:10.1016/j.rse.2020.111783CrossRefGoogle Scholar
Pechenick, A. M., Rizzo, D. M., Morrissey, L. A., Garvey, K. M., Underwood, K. L., & Wemple, B. C. (2014). A multi-scale statistical approach to assess the effects of connectivity of road and stream networks on geomorphic channel condition. Earth Surface Processes and Landforms, 39(11), 15381549. doi:10.1002/esp.3611CrossRefGoogle Scholar
Persichillo, M. G., Bordoni, M., Cavalli, M., Crema, S., & Meisina, C. (2018). The role of human activities on sediment connectivity of shallow landslides. Catena, 160. https://doi.org/10.1016/j.catena.2017.09.025CrossRefGoogle Scholar
Pöppl, R. E., Dilly, L. A., Haselberger, S., Renschler, C. S., & Baartman, J. E. M. (2019). Combining soil erosion modeling with connectivity analyses to assess lateral fine sediment input into agricultural streams. Water, 11(9). doi:10.3390/w11091793Google Scholar
Pöppl, R. E., Fryirs, K. A., Tunnicliffe, J., & Brierley, G. J. (2020). Managing sediment (dis)connectivity in fluvial systems. Science of the Total Environment, 736. doi:10.1016/j.scitotenv.2020.139627Google Scholar
Pöppl, R. E., Keiler, M., von Elverfeldt, K., Zweimueller, I., & Glade, T. (2012). The influence of riparian vegetation cover on diffuse lateral sediment connectivity and biogeomorphic processes in a medium-sized agricultural catchment, Austria. Geografiska Annaler Series a-Physical Geography, 94A(4), 511529. doi:10.1111/j.1468-0459.2012.00476.xCrossRefGoogle Scholar
Pöppl, R. E., Keesstra, S. D., & Maroulis, J. (2017). A conceptual connectivity framework for understanding geomorphic change in human-impacted fluvial systems. Geomorphology, 277, 237250. doi:10.1016/j.geomorph.2016.07.033CrossRefGoogle Scholar
Pöppl, R. E., Coulthard, T., Keesstra, S. D., & Keiler, M. (2019). Modeling the impact of dam removal on channel evolution and sediment delivery in a multiple dam setting. International Journal of Sediment Research, 34(6), 537549. doi:10.1016/j.ijsrc.2019.06.001CrossRefGoogle Scholar
Poesen, J. W. A., & Hooke, J. M. (1997). Erosion, flooding and channel management in Mediterranean environments of southern Europe. Progress in Physical Geography-Earth and Environment, 21(2), 157199. doi:10.1177/030913339702100201CrossRefGoogle Scholar
Puttock, A., Macleod, C. J. A., Bol, R., Sessford, P., Dungait, J., & Brazier, R. E. (2013). Changes in ecosystem structure, function and hydrological connectivity control water, soil and carbon losses in semi-arid grass to woody vegetation transitions. Earth Surface Processes and Landforms, 38(13). https://doi.org/10.1002/esp.3455CrossRefGoogle Scholar
Rhoads, B. L. (2020). River dynamics: Geomorphology to Support Management. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Roehl, J. W. (1962). Sediment source areas, delivery ratios and influencing morphological factors. In Symposium of Bari, 59 (59).Google Scholar
Rowntree, K., & Foster, I. (2012). A reconstruction of historical changes in sediment sources, sediment transfer and sediment yield in a small, semi-arid Karoo catchment, South Africa. Zeitschrift Fur Geomorphologie, 56, 87100. doi:10.1127/0372-8854/2012/s-00074CrossRefGoogle Scholar
Saco, P. M., Rodriguez, J. F., Moreno-de las Heras, M., Keesstra, S., Azadi, S., Sandi, S., … Rossi, M. J. (2020). Using hydrological connectivity to detect transitions and degradation thresholds: Applications to dryland systems. Catena, 186. doi:10.1016/j.catena.2019.104354CrossRefGoogle Scholar
Sandercock, P., & Hooke, J. (2006). Strategies for reducing sediment connectivity and land degradation in desertified areas using vegetation: The RECONDES project. IAHS-AISH Publication, 306, 127135.Google Scholar
Sandercock, P. J., & Hooke, J. M. (2011). Vegetation effects on sediment connectivity and processes in an ephemeral channel in SE Spain. Journal of Arid Environments, 75 (3), 239254.CrossRefGoogle Scholar
Schmitt, R. J. P., Bizzi, S., & Castelletti, A. (2016). Tracking multiple sediment cascades at the river network scale identifies controls and emerging patterns of sediment connectivity. Water Resources Research, 52(5), 39413965. doi:10.1002/2015wr018097CrossRefGoogle Scholar
Schmitt, R. J. P., Bizzi, S., Castelletti, A. F., & Kondolf, G. M. (2018). Stochastic modeling of sediment connectivity for reconstructing sand fluxes and origins in the unmonitored Se Kong, Se San, and Sre Pok tributaries of the Mekong river. Journal of Geophysical Research-Earth Surface, 123(1), 225. doi:10.1002/2016jf004105CrossRefGoogle Scholar
Schumm, S. A. (1977). The Fluvial System. New York: Wiley, 338 pp.Google Scholar
Sidle, R. C., Gomi, T., Usuga, J. C. L., & Jarihani, B. (2017). Hydrogeomorphic processes and scaling issues in the continuum from soil pedons to catchments. Earth-Science Reviews, 175, 7596.CrossRefGoogle Scholar
Singh, M, & Sinha, R. (2019). Evaluating dynamic hydrological connectivity of a floodplain wetland in North Bihar, India using geostatistical methods. Science of the Total Environment, 651, 24732488.CrossRefGoogle ScholarPubMed
Singh, M., & Sinha, R. (2020). Distribution, diversity, and geomorphic evolution of floodplain wetlands and wetland complexes in the Ganga plains of north Bihar, India. Geomorphology, 351. doi:10.1016/j.geomorph.2019.106960CrossRefGoogle Scholar
Singh, M., Sinha, R., & Tandon, S. K. (2021). Geomorphic connectivity and its application for understanding landscape complexities: a focus on the hydro-geomorphic systems of India. Earth Surface Processes and Landforms, 46(1), 110130. doi:10.1002/esp.4945CrossRefGoogle Scholar
Singh, M., Tandon, S. K., & Sinha, R. (2017). Assessment of connectivity in a water-stressed wetland (Kaabar Tal) of Kosi-Gandak interfan, north Bihar Plains, India. Earth Surface Processes and Landforms, 42(13), 19821996. doi:10.1002/esp.4156CrossRefGoogle Scholar
Skarpich, V., Galia, T., Ruman, S., & Macka, Z. (2019). Variations in bar material grain-size and hydraulic conditions of managed and re-naturalized reaches of the gravel-bed Becva River (Czech Republic). Science of the Total Environment, 649, 672685. doi:10.1016/j.scitotenv.2018.08.329CrossRefGoogle ScholarPubMed
Souza, J. O. P., Correa, A. C. B., & Brierley, G. J. (2016). An approach to assess the impact of landscape connectivity and effective catchment area upon bedload sediment flux in Saco Creek Watershed, Semiarid Brazil. Catena, 138, 1329. doi:10.1016/j.catena.2015.11.006CrossRefGoogle Scholar
Tan, Z. Q., Li, Y. L., Zhang, Q., Liu, X. G., Song, Y. Y., Xue, C. Y., & Lu, J. Z. (2021). Assessing effective hydrological connectivity for floodplains with a framework integrating habitat suitability and sediment suspension behavior. Water Research, 201. doi:10.1016/j.watres.2021.117253CrossRefGoogle ScholarPubMed
Tangi, M., Schmitt, R., Bizzi, S., & Castelletti, A. (2019). The CASCADE toolbox for analyzing river sediment connectivity and management. Environmental Modelling & Software, 119, 400406. doi:10.1016/j.envsoft.2019.07.008CrossRefGoogle Scholar
Tena, A., Piegay, H., Seignemartin, G., Barra, A., Berger, J. F., Mourier, B., & Winiarski, T. (2020). Cumulative effects of channel correction and regulation on floodplain terrestrialisation patterns and connectivity. Geomorphology, 354. doi:10.1016/j.geomorph.2020.107034CrossRefGoogle Scholar
Thompson, C. J., Fryirs, K., & Croke, J. (2016). The disconnected sediment conveyor belt: patterns of longitudinal and lateral erosion and deposition during a catastrophic flood in the Lockyer valley, south east Queensland, Australia. River Research and Applications, 32(4), 540551. doi:10.1002/rra.2897CrossRefGoogle Scholar
Toone, J., Rice, S. P., & Piegay, H. (2014). Spatial discontinuity and temporal evolution of channel morphology along a mixed bedrock-alluvial river, upper Drome River, southeast France: Contingent responses to external and internal controls. Geomorphology, 205, 516. doi:10.1016/j.geomorph.2012.05.033CrossRefGoogle Scholar
Trimble, S. W. (1983). A sediment budget for Coon Creek Basin in the Driftless Area, Wisconsin, 1853–1977. American Journal of Science, 283, 454474.CrossRefGoogle Scholar
Upadhayay, H. R., Lamichhane, S., Bajracharya, R. M., Cornelis, W., Collins, A. L., & Boeckx, P. (2020). Sensitivity of source apportionment predicted by a Bayesian tracer mixing model to the inclusion of a sediment connectivity index as an informative prior: Illustration using the Kharka catchment (Nepal). Science of the Total Environment, 713. doi:10.1016/j.scitotenv.2020.136703CrossRefGoogle Scholar
van der Waal, B., & Rowntree, K. (2018). Landscape connectivity in the upper Mzimvubu River catchment: an assessment of anthropogenic influences on sediment connectivity. Land Degradation & Development, 29(3), 713723. doi:10.1002/ldr.2766CrossRefGoogle Scholar
Vanacker, V., Molina, A., Govers, G., Poesen, J., Dercon, G., & Deckers, S. (2005). River channel response to short-term human-induced change in landscape connectivity in Andean ecosystems. Geomorphology, 72(1–4), 340353. doi:10.1016/j.geomorph.2005.05.013CrossRefGoogle Scholar
Wang, D. C., Wang, X., Huang, Y., Zhang, X., Zhang, W., Xin, Y., … Cao, Z. J. (2021). Impact analysis of small hydropower construction on river connectivity on the upper reaches of the great rivers in the Tibetan Plateau. Global Ecology and Conservation, 26. doi:10.1016/j.gecco.2021.e01496CrossRefGoogle Scholar
Walley, Y., Tunnicliffe, J., & Brierley, G. (2018). The influence of network structure upon sediment routing in two disturbed catchments, East Cape, New Zealand. Geomorphology, 307, 3849. doi:10.1016/j.geomorph.2017.10.029CrossRefGoogle Scholar
Walling, D. E. (1983). The sediment delivery problem. Journal of Hydrology, 65, 209237.CrossRefGoogle Scholar
Warner, R. F. (2006). Natural and artificial linkages and discontinuities in a Mediterranean landscape: Some case studies from the Durance Valley, France. Catena, 66(3), 236250. doi:10.1016/j.catena.2006.02.004CrossRefGoogle Scholar
Wohl, E., Brierley, G., Cadol, D., Coulthard, T. J., Covino, T., Fryirs, K. A., … Sklar, L. S. (2019). Connectivity as an emergent property of geomorphic systems. Earth Surface Processes and Landforms, 44(1), 426. doi:10.1002/esp.4434CrossRefGoogle Scholar
Zanandrea, F., Michel, G. P., Kobiyama, M., Censi, G., & Abatti, B. H. (2021). Spatial-temporal assessment of water and sediment connectivity through a modified connectivity index in a subtropical mountainous catchment. Catena, 204. doi:10.1016/j.catena.2021.105380CrossRefGoogle Scholar
Zhao, L., Liu, Y., & Luo, Y. (2020). Assessing hydrological connectivity mitigated by reservoirs, vegetation cover, and climate in Yan River Watershed on the Loess Plateau, China: the network approach. Water, 12(6). doi:10.3390/w12061742CrossRefGoogle Scholar
Zingaro, M., Refice, A., Giachetta, E., D’Addabbo, A., Lovergine, F., de Pasquale, V., Pepe, G., Brandolini, P., Cevasco, A., & Capolongo, D. (2019). Sediment mobility and connectivity in a catchment: a new mapping approach. Science of the Total Environment, 672. doi:10.1016/j.scitotenv.2019.03.461CrossRefGoogle Scholar
Zingaro, M., Refice, A., D’Addabbo, A., Hostache, R., Chini, M., & Capolongo, D. (2020). Experimental application of sediment flow connectivity index (SCI) in flood monitoring. Water, 12(7). doi:10.3390/w12071857CrossRefGoogle Scholar

References

Al-Masrahy, M. A., & Mountney, N. P.. 2015. A classification scheme for fluvial–aeolian system interaction in desert-margin settings. Aeolian Research 17: 6788.CrossRefGoogle Scholar
Anderson, R. S. 1987. A theoretical model for aeolian impact ripples. Sedimentology 34: 943956.CrossRefGoogle Scholar
Anthony, E. J., & Aagaard, T.. 2020. The lower shoreface: Morphodynamics and sediment connectivity with the upper shoreface and beach. Earth-Science Reviews 210: 103334.CrossRefGoogle Scholar
Armbrust, D. V., & Retta, A.. 2000. Wind and sandblast damage to growing vegetation. Annals of Arid Zone 39: 273284.Google Scholar
Bagnold, R. A. 1941. The Physics of Blown Sand and Desert Dunes. Methuen, New York.Google Scholar
Baker, A. R., Kelly, S. D., Biswas, K. F., Witt, M., & Jickells, T. D.. 2003. Atmospheric deposition of nutrients to the Atlantic Ocean. Geophysical Research Letters 30: 2296.CrossRefGoogle Scholar
Baker, A. R., Weston, K., Kelly, S. D., Voss, M., Streu, P., & Cape, J. N.. 2007. Dry and wet deposition of nutrients from the tropical Atlantic atmosphere: Links to primary productivity and nitrogen fixation. Deep-Sea Research Part I-Oceanographic Research Papers 54: 17041720.CrossRefGoogle Scholar
Ballantyne, A. P., Brahney, J., Fernandez, D., Lawrence, C. L., Saros, J., & Neff, J. C.. 2011. Biogeochemical response of alpine lakes to a recent increase in dust deposition in the Southwestern, US. Biogeosciences 8: 26892706.CrossRefGoogle Scholar
Bauer, B. O., Davidson-Arnott, R. G. D., Hesp, P. A., Namikas, S. L., Ollerhead, J., & Walker, I. J.. 2009. Aeolian sediment transport on a beach: Surface moisture, wind fetch, and mean transport. Geomorphology 105: 106116.CrossRefGoogle Scholar
Belly, P. Y. 1964. Sand Movement by Wind. Technical Memorandum No. 1. U.S. Army Coastal Engineering Research Center, Washington, DC.Google Scholar
Belnap, J., Phillips, S. L., Herrick, J. E., & Johansen, J. R.. 2007. Wind erodibility of soils at Fort Irwin, California (Mojave Desert), USA, before and after trampling disturbance: Implications for land management. Earth Surface Processes and Landforms 32: 7584.CrossRefGoogle Scholar
Bogle, R., Redsteer, M. H., & Vogel, J.. 2015. Field measurement and analysis of climatic factors affecting dune mobility near Grand Falls on the Navajo Nation, southwestern United States. Geomorphology 228: 4151.CrossRefGoogle Scholar
Bradley, E. F., & Mulhearn, P. J.. 1983. Development of velocity and shear-stress distributions in the wake of a porous shelter fence. Journal of Wind Engineering and Industrial Aerodynamics 15: 145156.CrossRefGoogle Scholar
Brandle, J. R., Hodges, L., & Zhou, X. H.. 2004. Windbreaks in North American agricultural systems. Pages 6578 in Nair, P. K. R., Rao, M. R., and Buck, L. E., editors. New Vistas in Agroforestry: A Compendium for 1st World Congress of Agroforestry, 2004. Springer Netherlands, Dordrecht.CrossRefGoogle Scholar
Chadwick, O. A., Derry, L. A., Vitousek, P. M., Huebert, B. J., & Hedin, L. O.. 1999. Changing sources of nutrients during four million years of ecosystem development. Nature 397: 491497.CrossRefGoogle Scholar
Charlson, R. J., & Heintzenberg, J.. 1995. Aerosol Forcing of Climate. Wiley, Chichester, New York.Google Scholar
Cornelis, W. M., & Gabriels, D.. 2005. Optimal windbreak design for wind-erosion control. Journal of Arid Environments 61: 315332.CrossRefGoogle Scholar
Davidson-Arnott, R. G. D., & Law, M. N.. 1990. Seasonal patterns and controls on sediment supply to coastal foredunes, Long Point, Lake Erie. Pages 177200 in Nordstrom, K. F., Psuty, N., & Carter, B., editors. Coastal Dunes: Form and Process. Wiley, New York.Google Scholar
de Vries, S., van Thiel de Vries, J. S. M., van Rijn, L. C., Arens, S. M., & Ranasinghe, R.. 2014. Aeolian sediment transport in supply limited situations. Aeolian Research 12: 7585.CrossRefGoogle Scholar
Dong, Z., Lv, P., Zhang, Z., & Lu, J.. 2014. Aeolian transport over a developing transverse dune. Journal of Arid Land 6: 243254.CrossRefGoogle Scholar
Draut, A. E. 2012. Effects of river regulation on aeolian landscapes, Colorado River, southwestern USA. Journal of Geophysical Research: Earth Surface 117: F02022, doi:10.1029/2011JF002329CrossRefGoogle Scholar
Duce, R. A., & Tindale, N. W.. 1991. Atmospheric transport of iron and its deposition in the ocean. Limnology and Oceanography 36: 17151726.CrossRefGoogle Scholar
Edwards, S. R. 1993. Luminescence dating of sand from the Kelso Dunes, California. Geological Society, London, Special Publications 72: 59.CrossRefGoogle Scholar
Fecan, F., Marticorena, B., & Bergametti, G.. 1999. Parameterization of the increase of the aeolian erosion threshold wind friction velocity due to soil moisture for arid and semi-arid areas. Annales Geophysicae-Atmospheres Hydrospheres and Space Sciences 17: 149157.Google Scholar
Feng, G., & Sharratt, B.. 2007. Validation of WEPS for soil and PM10 loss from agricultural fields within the Columbia Plateau of the United States. Earth Surface Processes and Landforms 32: 743753.CrossRefGoogle Scholar
Foken, T. 2006. 50 years of the Monin–Obukhov similarity theory. Boundary-Layer Meteorology 119: 431447.CrossRefGoogle Scholar
Fryberger, S. G., & Dean, G.. 1979. Dune forms and wind regime. A study of global sand seas 1052: 137169.Google Scholar
Fryrear, D. W. 1984. Soil Ridges-Clods and Wind Erosion. Transactions of the Asae 27: 445448.CrossRefGoogle Scholar
Gillette, D. A. 1974. On the production of soil wind erosion aerosols having the potential for long range transport. Journal des Recherches Atmospheriques 8: 735744.Google Scholar
Gillette, D. A. 1977. Fine particulate emissions due to wind erosion. Transactions of the American Society of Agricultural Engineers 20: 890897.CrossRefGoogle Scholar
Gillette, D. A. 1988. Threshold friction velocities for dust production for agricultural soils. Journal of Geophysical Research 93: 12645–12662.CrossRefGoogle Scholar
Gillette, D. A., & Pitchford, A. M.. 2004. Sand flux in the northern Chihuahuan desert, New Mexico, USA, and the influence of mesquite-dominated landscapes. Journal of Geophysical Research-Earth Surface 109: F04003.CrossRefGoogle Scholar
Gillette, D. A., Herrick, J. E., & Herbert, G. A.. 2006. Wind characteristics of mesquite streets in the northern Chihuahuan Desert, New Mexico, USA. Environmental Fluid Mechanics 6: 241275.CrossRefGoogle Scholar
Gillette, D. A., Niemeyer, T. C., & Helm, P. J.. 2001. Supply-limited horizontal sand drift at an ephemerally crusted, unvegetated saline playa. Journal of Geophysical Research-Atmospheres 106: 18085–18098.CrossRefGoogle Scholar
Gillette, D. A., Adams, J., Endo, A., Smith, D., & Kihl, R.. 1980. Threshold velocities for input of soil particles into the air by desert soils. Journal of Geographical Research 85: 56215630.Google Scholar
Gillies, J. A., Nield, J. M., & Nickling, W. G.. 2014. Wind speed and sediment transport recovery in the lee of a vegetated and denuded nebkha within a nebkha dune field. Aeolian Research 12: 135141.CrossRefGoogle Scholar
Gillies, J. A., Green, H., McCarley-Holder, G., Grimm, S., Howard, C., Barbieri, N., Ono, D., & Schade, T.. 2015. Using solid element roughness to control sand movement: Keeler Dunes, Keeler, California. Aeolian Research 18: 3546.CrossRefGoogle Scholar
Greeley, R., & Iversen, J. D.. 1985. Wind as a Geological Process on Earth, Mars, Venus and Titan. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Hagen, L. J. 2004. Evaluation of the Wind Erosion Prediction System (WEPS) erosion submodel on cropland fields. Environmental Modelling & Software 19: 171176.CrossRefGoogle Scholar
Han, G., Zhang, G., & Dong, Y.. 2007. A model for the active origin and development of source-bordering dunefields on a semiarid fluvial plain: A case study from the Xiliaohe Plain, Northeast China. Geomorphology 86: 512524.CrossRefGoogle Scholar
Haywood, J., & Boucher, O.. 2000. Estimates of the direct and indirect radiative forcing due to tropospheric aerosols: A review. Reviews of Geophysics 38: 513543.CrossRefGoogle Scholar
Hernández-Calvento, L., Jackson, D. W. T., Medina, R., Hernández-Cordero, A. I., Cruz, N., & Requejo, S.. 2014. Downwind effects on an arid dunefield from an evolving urbanised area. Aeolian Research 15: 301309.CrossRefGoogle Scholar
Herrick, J. E., Van Zee, J. W., Havstad, K. M., Burkett, L. M., & Whitford, W. G.. 2005. Monitoring Manual for Grasslands Shrublands and Savana Ecosysts. Volume 1: Quick Start. USDA-ARS Jornada Experimental Range, Las Cruces, New Mexico.Google Scholar
Hesp, P. 2002. Foredunes and blowouts: Initiation, geomorphology and dynamics. Geomorphology 48: 245268.CrossRefGoogle Scholar
Houser, C. 2009. Synchronization of transport and supply in beach-dune interaction. Progress in Physical Geography: Earth and Environment 33: 733746.CrossRefGoogle Scholar
Houser, C., & Mathew, S.. 2011. Alongshore variation in foredune height in response to transport potential and sediment supply: South Padre Island, Texas. Geomorphology 125: 6272.CrossRefGoogle Scholar
Iversen, J. D., & White, B. R.. 1982. Saltation threshold on Earth, Mars and Venus. Sedimentology 29: 111119.CrossRefGoogle Scholar
Iversen, J. D., Pollack, J. B., Greeley, R., & White, B. R.. 1976. Saltation threshold on Mars – effect of interparticle force, surface-roughness, and low atmospheric density. Icarus 29: 381393.CrossRefGoogle Scholar
Ivester, A. H., & Leigh, D. S.. 2003. Riverine dunes on the coastal plain of Georgia, USA. Geomorphology 51: 289311.CrossRefGoogle Scholar
Karl, J. W., Duniway, M. C., & Schrader, T. S.. 2011. A technique for estimating rangeland canopy-gap size distributions from high-resolution digital imagery. Rangeland Ecology & Management 65: 196207.CrossRefGoogle Scholar
Kawamura, R. 1951. Study of Sand Movement by Wind. University of California, Berkeley, CA.Google Scholar
Kocurek, G., Carr, M., Ewing, R., Havholm, K. G., Nagar, Y. C., & Singhvi, A. K.. 2007. White sands dune field, New Mexico: age, dune dynamics and recent accumulations. Sedimentary Geology 197: 313331.CrossRefGoogle Scholar
Kocurek, G., & Havholm, K. G.. 1993. Chapter 16. Eolian sequence stratigraphy-a conceptual framework. Pages 393409 in Posamentier, H. W. and Allen, G. P., editors. Recent Developments in Siliciclastic Sequence Stratigraphy, AAPG Memoir 58. The American Association of Petroleum Geologists, Tulsa, OK.Google Scholar
Kocurek, G., & Lancaster, N.. 1999. Aeolian system sediment state: Theory and Mojave Desert Kelso dune field example. Sedimentology 46: 505515.CrossRefGoogle Scholar
Lancaster, N. 1997. Response of eolian geomorphic systems to minor climate change: Examples from the southern Californian deserts. Geomorphology 19: 333347.CrossRefGoogle Scholar
Lawrence, C. R., Neff, J. C., & Farmer, G. L.. 2011. The accretion of aeolian dust in soils of the San Juan Mountains, Colorado, USA. Journal of Geophysical Research 116: F02013.CrossRefGoogle Scholar
Li, J. R., Okin, G. S., & Epstein, H. E.. 2009b. Effects of enhanced wind erosion on surface soil texture and characteristics of windblown sediments. Journal of Geophysical Research-Biogeosciences 114. G02003. doi:10.1029/2008JG000903CrossRefGoogle Scholar
Li, J., Okin, G. S., Alvarez, L. J., & Epstein, H. E.. 2009a. Sediment deposition and soil nutrient heterogeneity in two desert grassland ecosystems, southern New Mexico. Plant and Soil 319: 6784.CrossRefGoogle Scholar
Li, J., Okin, G. S., Hartman, L. J., & Epstein, H. E.. 2007. Quantitative assessment of wind erosion and soil nutrient loss in desert grasslands of southern New Mexico, USA. Biogeochemistry 85: 317332.CrossRefGoogle Scholar
Li, J., Okin, G. S., Tatarko, J., Webb, N. P., & Herrick, J. E.. 2014. Consistency of wind erosion assessments across land use and land cover types: A critical analysis. Aeolian Research 15: 253260.CrossRefGoogle Scholar
Li, J., Okin, G. S., Herrick, J. E., Belnap, J., Miller, M. E., Vest, K., & Draut, A. E.. 2013. Evaluation of a new model of aeolian transport in the presence of vegetation. Journal of Geophysical Research: Earth Surface 118: 288306.CrossRefGoogle Scholar
Litaor, M. I. 1987. The influence of Eolian dust on the genesis of alpine soils in the front range, Colorado. Soil Science Society of America Journal 51: 142147.CrossRefGoogle Scholar
Ludwig, J. A., Tongway, D. J., & Marsden, S. G.. 1999. Stripes, strands or stipples: Modelling the influence of three landscape banding patterns on resource capture and productivity in semi-arid woodlands, Australia. Catena 37: 257273.CrossRefGoogle Scholar
Ludwig, J. A., Bastin, G. N., Chewings, V. H., Eager, R. W., & Liedloff, A. C.. 2007. Leakiness: A new index for monitoring the health of arid and semiarid landscapes using remotely sensed vegetation cover and elevation data. Ecological Indicators 7: 442454.CrossRefGoogle Scholar
Ludwig, J. A., Eager, R. W., Bastin, G. N., Chewings, V. H., and Liedloff, A. C.. 2002. A leakiness index for assessing landscape function using remote sensing. Landscape Ecology 17: 157171.CrossRefGoogle Scholar
Ludwig, J. A., Eager, R. W., Liedloff, A. C., Bastin, G. N., & Chewings, V. H.. 2006. A new landscape leakiness index based on remotely sensed ground-cover data. Ecological Indicators 6: 327336.CrossRefGoogle Scholar
Ma, R., Li, J., Ma, Y., Wei, L., & Zhang, Y.. 2020. A wind tunnel study of the seasonal shelter efficiency of deciduous windbreaks. Transactions of the ASABE 63: 913922.CrossRefGoogle Scholar
Mahowald, N. M., Baker, A. R., Bergametti, G., Brooks, N., Duce, R. A., Jickells, T. D., Kubilay, N., Prospero, J. M., & Tegen, I.. 2005. Atmospheric global dust cycle and iron inputs to the ocean. Global Biogeochemical Cycles 19. GB4025, doi:10.1029/2004GB002402CrossRefGoogle Scholar
Mahowald, N., Jickells, T. D., Baker, A. R., Artaxo, P., Benitez-Nelson, C. R., Bergametti, G., Bond, T. C., Chen, Y., Cohen, D. D., Herut, B., Kubilay, N., Losno, R., Luo, C., Maenhaut, W., McGee, K. A., Okin, G. S., Siefert, R. L., & Tsukuda, S.. 2008. The global distribution of atmospheric phosphorus deposition and anthropogenic impacts. Global Biogeochemical Cycles 22: GB4026.CrossRefGoogle Scholar
Mahowald, N. M., Engelstaedter, S., Luo, C., Sealy, A., Artaxo, P., Benitez-Nelson, C. R., Bonnet, S., Chen, Y., Chuang, P. Y., Cohen, D. D., Dulac, F., Herut, B., Johansen, A. M., Kubilay, N., Losno, R., Maenhaut, W., Paytan, A., Prospero, J. M., Shank, L. M., & Siefert, R. L.. 2009. Atmospheric iron deposition: Global distribution, variability and human perturbations. Annual Reviews of Marine Sciences 1: 245278.CrossRefGoogle ScholarPubMed
Marqués, M. A., Psuty, N. P., & Rodriguez, R.. 2001. Neglected effects of Eolian dynamics on artificial beach nourishment: The case of Riells, Spain. Journal of Coastal Research 17: 694704.Google Scholar
Marsham, J. H., Knippertz, P., Dixon, N. S., Parker, D. J., & Lister, G. M. S.. 2011. The importance of the representation of deep convection for modeled dust-generating winds over West Africa during summer. Geophysical Research Letters 38: L16803.CrossRefGoogle Scholar
Marticorena, B., & Bergametti, G.. 1995. Modeling the atmospheric dust cycle: 1. Design of a soil-derived dust emission scheme. Journal of Geophysical Research 100: 16415–16430.CrossRefGoogle Scholar
Mayaud, J. R., Wiggs, G. F. S., & Bailey, R. M.. 2016. Characterizing turbulent wind flow around dryland vegetation. Earth Surface Processes and Landforms 41: 14211436.CrossRefGoogle Scholar
McGlynn, I. O., & Okin, G. S.. 2006. Characterization of shrub distribution using high spatial resolution remote sensing: Ecosystem implications for a former Chihuahuan Desert grassland. Remote Sensing of Environment 101: 554566.CrossRefGoogle Scholar
Mladenov, N., Sommaruga, R., Morales-Baquero, R., Laurion, I., Camarero, L., Diéguez, M. C., Camacho, A., Delgado, A., Torres, O., Chen, Z., Felip, M., & Reche, I.. 2011. Dust inputs and bacteria influence dissolved organic matter in clear alpine lakes. Nature Communications 2: 405.CrossRefGoogle ScholarPubMed
Moreno-de las Heras, M., Saco, P. M., Willgoose, G. R., & Tongway, D. J.. 2011. Assessing landscape structure and pattern fragmentation in semiarid ecosystems using patch-size distributions. Ecological Applications 21: 27932805.CrossRefGoogle ScholarPubMed
Muhs, D. R., Roskin, J., Tsoar, H., Skipp, G., Budahn, J. R., Sneh, A., Porat, N., Stanley, J.-D., Katra, I., & Blumberg, D. G.. 2013. Origin of the Sinai–Negev erg, Egypt and Israel: Mineralogical and geochemical evidence for the importance of the Nile and sea level history. Quaternary Science Reviews 69: 2848.CrossRefGoogle Scholar
Munroe, J. S. 2014. Properties of modern dust accumulating in the Uinta Mountains, Utah, USA, and implications for the regional dust system of the Rocky Mountains. Earth Surface Processes and Landforms 39: 19791988.CrossRefGoogle Scholar
Munroe, J. S., McElroy, R., O’Keefe, S., Peters, A., & Wasson, L.. 2021. Holocene records of eolian dust deposition from high-elevation lakes in the Uinta Mountains, Utah, USA. Journal of Quaternary Science 36: 6675.CrossRefGoogle Scholar
Nordstrom, K. F., & Jackson, N. L.. 1992. Effect of source width and tidal elevation changes on aeolian transport on an estuarine beach. Sedimentology 39: 769778.CrossRefGoogle Scholar
Nordstrom, K. F., & Hotta, S.. 2004. Wind erosion from cropland in the USA: A review of problems, solutions and prospects. Geoderma 121: 157167.CrossRefGoogle Scholar
Okin, G. S. 2008. A new model for wind erosion in the presence of vegetation. Journal of Geophysical Research-Earth Surface 113: F02S10.CrossRefGoogle Scholar
Okin, G. S., & Gillette, D. A.. 2001. Distribution of vegetation in wind-dominated landscapes: Implications for wind erosion modeling and landscape processes. Journal of Geophysical Research 106: 96739683.CrossRefGoogle Scholar
Okin, G. S., & Painter, T. H.. 2004. Effect of grain size on remotely sensed spectral reflectance of sandy desert surfaces. Remote Sensing of Environment 89: 272280.CrossRefGoogle Scholar
Okin, G. S., Murray, B., & Schlesinger, W. H.. 2001a. Degradation of sandy arid shrubland environments: Observations, process modelling, and management implications. Journal of Arid Environments 47: 123144.CrossRefGoogle Scholar
Okin, G. S., Herrick, J. E., & Gillette, D. A.. 2006. Multiscale controls on and consequences of aeolian processes in landscape change in arid and semiarid environments. Journal of Arid Environments 65: 253275.CrossRefGoogle Scholar
Okin, G. S., Murray, B., & Schlesinger, W. H.. 2001b. Desertification in an arid shrubland in the southwestern United States: Process modeling and validation. Pages 5370 in Conacher, A., editor. Land Degradation: Papers Selected from Contributions to the Sixth meeting of the International Geographical Union’s Commission on Land Degradation and Desertification, Perth, Western Australia, 20–28 September 1999. Kluwer Academic Publishers, Dordrecht.CrossRefGoogle Scholar
Okin, G. S., Mahowald, N., Chadwick, O. A., & Artaxo, P.. 2004. The impact of desert dust on the biogeochemistry of phosphorus in terrestrial ecosystems. Global Biogeochemical Cycles 18: 10.1029/2003GB002145.CrossRefGoogle Scholar
Okin, G. S., Sala, O. E., Vivoni, E. R., Zhang, J., & Bhattachan, A.. 2018. The interactive role of wind in water in functioning of drylands: What does the future hold? Bioscience 68: 670677.CrossRefGoogle Scholar
Okin, G. S., Parsons, A. J., Wainwright, J., Herrick, J. E., Bestelmeyer, B. T., Peters, D. P. C., & Fredrickson, E. L.. 2009. Do changes in connectivity explain desertification? Bioscience 59: 237244.CrossRefGoogle Scholar
Okin, G. S., Baker, A. R., Tegen, I., Mahowald, N. M., Dentener, F. J., Duce, R. A., Galloway, J. N., Hunter, K. A., Kanakidou, M., Kubilay, N., Prospero, J. M., Sarin, M. M., Surapipith, V., Uematsu, M., & Zhuo, T.. 2011. Impacts of atmospheric nutrient deposition on marine productivity: Roles of nitrogen, phosphorus, and iron Global Biogeochemical Cycles 25: GB2022.CrossRefGoogle Scholar
Page, K. 1971. Riverine source bordering sand dune. Australian Geographer 11: 603605.CrossRefGoogle Scholar
Page, K. J., Dare-Edwards, A. J., Owens, J. W., Frazier, P. S., Kellett, J., & Price, D. M.. 2001. TL chronology and stratigraphy of riverine source bordering sand dunes near Wagga Wagga, New South Wales, Australia. Quaternary International 83–85: 187193.CrossRefGoogle Scholar
Painter, T. H., Deems, J., Belnap, J., Hamlet, A. F., Landry, C. C., & Udall, B.. 2010. Response of Colorado River runoff to dust radiative forcing in snow. Proceedings of the National Academy of Science 107: 17125–17130.CrossRefGoogle ScholarPubMed
Painter, T. H., Barrett, A. P., Landry, C., Neff, J., Cassidy, M. P., Lawrence, C., McBride, K. E., & Farmer, G. L.. 2007. Impact of disturbed desert soils on duration of mountain snowcover. Geophysical Research Letters 34: L12502.CrossRefGoogle Scholar
Prospero, J. M. 1999. Long-term measurements of the transport of African mineral dust to the southeastern United States: Implications for regional air quality. Journal of Geophysical Research-Atmospheres 104: 15917–15927.CrossRefGoogle Scholar
Psuty, N. 1996. Coastal foredune development and vertical displacement. Zeitschrift für Geomorphologie. Supplement band 102: 211221.Google Scholar
Raupach, M. R. 1992. Drag and drag partition on rough surfaces. Boundary-Layer Meteorology 60: 375395.CrossRefGoogle Scholar
Raupach, M. R., Gillette, D. A., & Leys, J. F.. 1993. The effect of roughness elements on wind erosion threshold. Journal of Geophysical Research 98: 30233029.CrossRefGoogle Scholar
Raupach, M. R., Woods, N., Dorr, G., Leys, J. F., & Cleugh, H. A.. 2001. The entrapment of particles by windbreaks. Atmospheric Environment 35: 33733383.CrossRefGoogle Scholar
Ravi, S., D’Odorico, P., Breshears, D. D., Field, J. P., Goudie, A. S., Huxman, T. E., Li, J., Okin, G. S., Swap, R. J., Thomas, A. D., Van Pelt, R. S., Whicker, J. J., & Zobeck, T. M.. 2011. Aeolian processes and the biosphere. Reviews of Geophysics 49: RG3001.CrossRefGoogle Scholar
Rendell, H. M., Clarke, M. L., Warren, A., & Chappell, A.. 2003. The timing of climbing dune formation in southwestern Niger: Fluvio-aeolian interactions and the rôle of sand supply. Quaternary Science Reviews 22: 10591065.CrossRefGoogle Scholar
Sankey, J. B., Caster, J., Kasprak, A., & East, A. E.. 2018a. The response of source-bordering aeolian dunefields to sediment-supply changes 2: Controlled floods of the Colorado River in Grand Canyon, Arizona, USA. Aeolian Research 32: 154169.CrossRefGoogle Scholar
Sankey, J. B., Kasprak, A., Caster, J., East, A. E., & Fairley, H. C.. 2018b. The response of source-bordering aeolian dunefields to sediment-supply changes 1: Effects of wind variability and river-valley morphodynamics. Aeolian Research 32: 228245.CrossRefGoogle Scholar
Schlesinger, W. H., Reynolds, J. F., Cunningham, G. L., Huenneke, L. F., Jarrell, W. M., Virginia, R. A., & Whitford, W. G.. 1990. Biological feedbacks in global desertification. Science 247: 10431048.CrossRefGoogle ScholarPubMed
Shao, Y., Raupach, M. R., & Findlater, P. J.. 1993. Effect of saltation bombardment on the entrainment of dust by wind. Journal of Geophysical Research 98: 12719–12726.CrossRefGoogle Scholar
Sherman, D. J., & Lyons, W.. 1994. Beach-state controls on aeolian sand delivery to coastal dunes. Physical Geography 15: 381395.CrossRefGoogle Scholar
Sokolik, I. N., & Toon, O. B.. 1996. Direct radiative forcing by anthropogenic airborne mineral aerosols. Nature 381: 681683.CrossRefGoogle Scholar
Sterk, G., Parigiani, J., Cittadini, E., Peters, P., Scholberg, J., & Peri, P.. 2012. Aeolian sediment mass fluxes on a sandy soil in Central Patagonia. Catena 95: 112123.CrossRefGoogle Scholar
Swap, R., Garstang, M., Greco, S., Talbot, R., & Kallberg, P.. 1992. Saharan dust in the Amazon Basin. Tellus Series B-Chemical and Physical Meteorology 44: 133149.CrossRefGoogle Scholar
Thomas, D. S. G., & Wiggs, G. F. S.. 2008. Aeolian system responses to global change: Challenges of scale, process and temporal integration. Earth Surface Processes and Landforms 33: 13961418.CrossRefGoogle Scholar
Torshizi, M. R., Miri, A., Shahriari, A., Dong, Z., & Davidson-Arnott, R.. 2020. The effectiveness of a multi-row Tamarix windbreak in reducing aeolian erosion and sediment flux, Niatak area, Iran. Journal of Environmental Management 265: 110486.CrossRefGoogle ScholarPubMed
Von Kármán, T. 1931. Mechanical Similitude and Turbulence. National Advisory Committee for Aeronautics. https://books.google.co.uk/books?id=ONBCAQAAIAAJ&dq=von+karman&lr=&source=gbs_navlinks_sGoogle Scholar
Walter, B., Gromke, C., Leonard, K. C., Manes, C., and Lehning, M.. 2012. Spatio-temporal surface shear-stress variability in live plant canopies and cube arrays. Boundary-Layer Meteorology 143: 337356.CrossRefGoogle Scholar
Webb, N. P., Okin, G. S., & Brown, S.. 2014. The effect of roughness elements on wind erosion: The importance of surface shear stress distribution. Journal of Geophysical Research-Atmospheres 119: 60666084.CrossRefGoogle Scholar
Webb, N. P., Galloza, M. S., Zobeck, T. M., & Herrick, J. E.. 2016. Threshold wind velocity dynamics as a driver of aeolian sediment mass flux. Aeolian Research 20: 4558.CrossRefGoogle Scholar
Webb, N. P., Kachergis, E., Miller, S. W., McCord, S. E., Bestelmeyer, B. T., Brown, J. R., Chappell, A., Edwards, B. L., Herrick, J. E., Karl, J. W., Leys, J. F., Metz, L. J., Smarik, S., Tatarko, J., Van Zee, J. W., & Zwicke, G.. 2020. Indicators and benchmarks for wind erosion monitoring, assessment and management. Ecological Indicators 110: 105881.CrossRefGoogle Scholar
Wolfe, S. A., & Nickling, W. G.. 1993. The protective role of sparse vegetation in wind erosion. Progress in Physical Geography: Earth and Environment 17: 5068.CrossRefGoogle Scholar
Wolfe, S. A., & Nickling, W. G.. 1996. Shear stress partitioning in sparsely vegetated desert canopies. Earth Surface Processes and Landforms 21: 607619.3.0.CO;2-1>CrossRefGoogle Scholar
Wopfner, H., & Twidale, C. R.. 1988. Formation and age of desert dunes in the Lake Eyre depocentres in central Australia. Geologische Rundschau 77: 815834.CrossRefGoogle Scholar
Yu, L., Lai, Z., & An, P.. 2013. OSL chronology and paleoclimatic implications of paleodunes in the middle and southwestern Qaidam Basin, Qinghai–Tibetan Plateau. Sciences in Cold and Arid Regions 5: 211219.Google Scholar
Zhang, J., Okin, G. S., Zhou, B., & Karl, J. W.. 2021b. UAV-derived imagery for vegetation structure estimation in rangelands: Validation and application. Ecosphere 12: e03830.CrossRefGoogle Scholar
Zhang, J., Guo, W., Zhou, B., & Okin, G. S.. 2021a. Drone-based remote sensing for research on wind erosion in drylands: Possible applications. Remote Sensing 13: 283.CrossRefGoogle Scholar

References

Ali, G. A., & Roy, A. G. (2009). Revisiting hydrologic sampling strategies for an accurate assessment of hydrologic connectivity in humid temperate systems. Geography Compass, 3(1), 350374. https://doi.org/10.1111/j.1749-8198.2008.00180.xCrossRefGoogle Scholar
Alley, R. B. (2000). The Two-Mile Time Machine. Princeton: Princeton University Press.Google Scholar
Altmann, M., Piermattei, L., Haas, F., Heckmann, T., Fleischer, F., Rom, J., Betz-Nutz, S., Knoflach, B., Müller, S., Ramskogler, K., Pfeiffer, M., Hofmeister, F., Ressl, C., & Becht, M. (2020). Long-term changes of morphodynamics on little ice age lateral moraines and the resulting sediment transfer into mountain streams in the Upper Kauner Valley, Austria. Water, 12(12):3375. https://doi.org/10.3390/w12123375CrossRefGoogle Scholar
Bernhardt, A., Schwanghart, W., Hebbeln, D., Stuut, J. W., & Strecker, M. R. (2017). Immediate propagation of deglacial environmental change to deep-marine turbidite systems along the Chile convergent margin. Earth and Planetary Science Letters, 473, 190204. https://doi.org/10.1016/j.epsl.2017.05.017CrossRefGoogle Scholar
Bitanja, R. (1999). On the glaciological, meteorological and climatological significance of Antarctic blue ice areas. Reviews of Geophysics, 37(3), 337359.CrossRefGoogle Scholar
Boulton, G. S. (2006). Glaciers and their coupling with hydraulic and sedimentary processes. In Knight, P. G., ed., Glacier Science and Environmental Change. Oxford: Blackwell, pp. 322.Google Scholar
Bracken, L. J., Wainwright, J., Ali, G. A., Tetzlaff, D., Smith, M. W., Reaney, S. M., & Roy, A. G. (2013). Concepts of hydrological connectivity: Research approaches, pathways and future agendas. Earth-Science Reviews, 119, 17-34. https://doi.org/10.1016/j.earscirev.2013.02.001CrossRefGoogle Scholar
Bracken, L. J., Turnbull, L., Wainwright, J., & Bogaart, P. (2015). Sediment connectivity: A framework for understanding sediment transfer at multiple scales. Earth Surface Processes and Landforms, 40(2), 177188.CrossRefGoogle Scholar
Caine, N. (1986). Sediment movement and storage on alpine slopes in the Colorado Rocky Mountains. In Abrahams, A. D., ed., Hillslope Processes. London: Allen & Unwin, pp. 115137.Google Scholar
Collins, D. (1979). Quantitative Determination of the Subglacial Hydrology of Two Alpine Glaciers. Journal of Glaciology, 23(89), 347362. https://doi.org/10.3189/S0022143000029956CrossRefGoogle Scholar
Cook, S. J., & Swift, D. A. (2012). Subglacial basins: Their origin and importance in glacial systems and landscapes. Earth Science Reviews, 115(4), 332372. http://doi.org/10.1016/j.earscirev.2012.09.009CrossRefGoogle Scholar
Etzelmüller, B., Ødegard, R. S., Vatne, G., Mysterud, R. S., Tonning, T., and Sollid, J. L. (2000). Glacier characteristics and sediment transfer system of Longyearbreen and Larsbreen, western Spitsbergen. Norsk Geografisk Tidsskrift–Norwegian Journal of Geography, 54, 157168. https://doi.org/10.1080/002919500448530CrossRefGoogle Scholar
Evatt, G. W., Coughlan, M. J., Joy, K. H., Smedley, A. R. D., Connolly, P. J., & Abrahams, I. D. (2016). A potential hidden layer of meteorites below the ice surface of Antarctica. Nature Communications, 7, 10679. https://doi.org/10.1038/ncomms10679CrossRefGoogle ScholarPubMed
Ferguson, R. (1981). Channel forms and channel changes. In Lewin, J., ed., British Rivers. London: Allen & Unwin, pp. 90125.Google Scholar
Hassan, M. A., Bird, S., Reid, D., Ferrer-Boix, C., Hogan, D., Brardinoni, F., & Chartrand, S. (2019). Variable hillslope-channel coupling and channel characteristics of forested mountain streams in glaciated landscapes. Earth Surface Processes and Landforms, 44, 736751. https://doi.org/10.1002/esp.4527CrossRefGoogle Scholar
Jaeger, J. M., & Koppes, M. N. (2016). The role of the cryosphere in source-to-sink systems. Earth-Science Reviews, 153, 4376. https://doi.org/10.1016/j.earscirev.2015.09.011CrossRefGoogle Scholar
Knight, P. G. (1997). The basal ice layer of glaciers and ice sheets. Quaternary Science Reviews, 16, 975993. https://doi.org/10.1016/S0277-3791(97)00033-4CrossRefGoogle Scholar
Knight, P. G. (1999). Glaciers. Cheltenham: Stanley Thornes.Google Scholar
Knight, P. G., Jennings, C. E., Waller, R. I., & Robinson, Z. P. (2007). Changes in ice-margin processes and sediment routing during ice-sheet advance across a marginal moraine. Geografiska Annaler: Series A, Physical Geography, 89(3), 203215. https://doi.org/10.1111/j.1468-0459.2007.00319.xCrossRefGoogle Scholar
Kummert, M., & Delaloye, R. (2018). Mapping and quantifying sediment transfer between the front of rapidly moving rock glaciers and torrential gullies. Geomorphology, 309, 6076. https://doi.org/10.1016/j.geomorph.2018.02.021CrossRefGoogle Scholar
Lane, S. N., Bakker, M., Gabbud, C., Micheletti, N., & Saugi, G. (2017). Sediment export, transient landscape response and catchment-scale connectivity following rapid climate warming and Alpine glacier recession. Geomorphology, 277, 210227. https://doi.org/10.1016/j.geomorph.2016.02.015CrossRefGoogle Scholar
Li, Y., Lu, Y., Zhang, Z., Shi, H., & Xi, H. (2019). Characterizing three-dimensional features of Antarctic subglacial lakes from the inversion of hydraulic potential – Lake Vostok as a case study. Advances in Polar Science, 30, 7075. https://doi.org/10.13679/j.advps.2019.1.00070Google Scholar
MacDonell, S., Sharp, M., & Fitzsimons, S. (2016). Cryoconite hole connectivity on the Wright Lower Glacier, McMurdo Dry Valleys, Antarctica. Journal of Glaciology, 62(234), 714724. https://doi.org/10.1017/jog.2016.62CrossRefGoogle Scholar
Mancini, D., & Lane, S. N. (2020). Changes in sediment connectivity following glacial debuttressing in an Alpine valley system. Geomorphology, 352, 106987. https://doi.org/10.1016/j.geomorph.2019.106987CrossRefGoogle Scholar
Miles, E. S., Steiner, J., Willis, I., Buri, P., Immerzeel, W. W., Chesnokova, A., & Pellicciotti, F. (2017). Pond Dynamics and Supraglacial-Englacial Connectivity on Debris-Covered Lirung Glacier, Nepal. Frontiers in Earth Science, 5(69). https://doi.org/10.3389/feart.2017.00069CrossRefGoogle Scholar
Piotrowski, J. A. (2006). Groundwater under ice sheets and glaciers. In Knight, P. G, ed., Glacier Science and Environmental Change. Oxford: Blackwell.Google Scholar
Pöppl, R. E., & Parsons, A. J. (2018). The geomorphic cell: A basis for studying connectivity. Earth Surface Processes and Landforms, 34, 11551159. https://doi.org/10.1002/esp.4300CrossRefGoogle Scholar
Porter, P., Smart, M., Irvine-Fynn, T. D. L. (2019). Glacial sediment stores and their reworking. In: Heckmann, T. & Morche, D. eds., Geomorphology of Proglacial Systems. Geography of the Physical Environment. Cham: Springer. https://doi.org/10.1007/978-3-319-94184-4_10Google Scholar
Small, R., Beecroft, I., & Stirling, D. (1984). Rates of deposition on Lateral Moraine Embankments, Glacier De Tsidjiore Nouve, Valais, Switzerland. Journal of Glaciology, 30(106), 275281. https://doi.org/10.3189/S0022143000006092CrossRefGoogle Scholar
Small, R. J. (1987). Moraine sediment budgets. In Gurnell, A. M. & Clark, M. J., eds., Glacio-Fluvial Sediment Transfer – An Alpine Perspective. Chichester: John Wiley and Sons, pp. 165197.Google Scholar
Stevens, I., Irvine-Fynn, T., Porter, P. R., Cook, J., Edwards, A., Smart, M., Moorman, B., Hodson, A., & Mitchell, A. (2018). Near-surface hydraulic conductivity of Northern Hemisphere glaciers. Hydrological Processes, 32(7), 850865. https://doi.org/10.1002/hyp.11439CrossRefGoogle Scholar
Stocker-Waldhuber, M., Kuhn, M. (2019). Closing the balances of ice, water and sediment fluxes through the terminus of gepatschferner. In Heckmann, T and Morche, D, eds. Geomorphology of Proglacial Systems. Geography of the Physical Environment. Cham: Springer. https://doi.org/10.1007/978-3-319-94184-4_5Google Scholar
Swift, D. A., Cook, S. J., Graham, D., Midgley, N., Fallick, A. E., Storrar, R., Toubes Rodrigo, M., & Evans, D. (2018). Terminal zone glacial sediment transfer at a temperate overdeepened glacier system. Quaternary Science Reviews, 180, 111–131. https://doi.org/10.1016/j.quascirev.2017.11.027CrossRefGoogle Scholar
Swift, D. A., Tallentire, G. D., Farinotti, D., Cook, S. J., Higson, W. J., & Bryant, R. G. (2021). The hydrology of glacier-bed overdeepenings: Sediment transport mechanics, drainage system morphology, and geomorphological implications. Earth Surface Processes and Landforms, 46, 115. https://doi.org/10.1002/esp.5173CrossRefGoogle Scholar
Toubes-Rodrigo, M., Potgieter-Vermaak, S., Sen, R., Oddsdottir, E. S, Elliott, D., & Cook, S. (2021). Active microbial ecosystem in glacier basal ice fuelled by iron and silicate comminution-derived hydrogen. Microbiology Open, 10(4), e1200. https://doi.org/10.1002/mbo3.1200CrossRefGoogle ScholarPubMed
Warburton, J. (1990). An alpine proglacial fluvial sediment budget. Geografiska Annaler Series A, Physical Geography, 72(3/4), 261272. https://doi.org/10.2307/521154CrossRefGoogle Scholar
Wright, A. P., Siegert, M. J., Le Brocq, A. M., & Gore, D. B. (2008). High sensitivity of subglacial hydrological pathways in Antarctica to small ice-sheet changes. Geophysical Research Letters, 35, L17504. https://doi.org/10.1029/2008GL034937CrossRefGoogle Scholar
Wright, A. P., Young, D. A., Roberts, J. L., Schroeder, D. M., Bamber, J. L., Dowdeswell, J. A., Young, N. W., Le Brocq, A. M., Warner, R. C., Payne, A. J., Blankenship, D. D., van Ommen, T. D., & Siegert, M. J. (2012). Evidence of a hydrological connection between the ice divide and ice sheet margin in the Aurora Subglacial Basin, East Antarctica. Journal of Geophysical Research Earth Surface, 117(F1), F01033. https://doi.org/10.1029/2011JF002066Google Scholar
Wright, A., Young, D., Bamber, J., Dowdeswell, J., Payne, A., Blankenship, D., & Siegert, M. (2014). Subglacial hydrological connectivity within the Byrd Glacier catchment, East Antarctica. Journal of Glaciology, 60(220), 345352. https://doi.org/10.3189/2014JoG13J014CrossRefGoogle Scholar

References

Alley, R. B., Cuffey, K. M. & Zoet, L. K. (2019). Glacial erosion: Status and outlook. Annals of Glaciology 60 (80): 113. https://doi.org/10.1017/aog.2019.38.CrossRefGoogle Scholar
Alley, R. B., Cuffey, K. M., Evenson, E. B., Strasser, J. C., Lawson, D. E. & Larson, G. J. (1997). How glaciers entrain and transport basal sediment: Physical constraints. Quaternary Science Reviews 16 (9): 10171038. https://doi.org/10.1016/S0277-3791(97)00034-6.CrossRefGoogle Scholar
Anselmetti, F. S., Bühler, R., Finger, D., Girardclos, S., Lancini, A., Rellstab, C. & Sturm, M. (2007). Effects of Alpine hydropower dams on particle transport and lacustrine sedimentation. Aquatic Sciences 69 (2): 179198. https://doi.org/10.1007/s00027-007-0875-4.CrossRefGoogle Scholar
Antoniazza, G. & Lane, S. N. (2021). Sediment yield over glacial cycles: A conceptual model. Progress in Physical Geography: Earth and Environment, 45(6), 842865. https://doi.org/10.1177/0309133321997292.CrossRefGoogle Scholar
Arfstrom, J. D. (2003). Protalus ramparts and transverse ridge moraines on Mars: Indicators of surface ice depositional processes. In 34th Annual Lunar and Planetary Science Conference, March 17–21, 2003, League City, TX, abstract no.1050.Google Scholar
Ashmore, P. (1993). Contemporary erosion of the Canadian landscape. Progress in Physical Geography 17: 190204.CrossRefGoogle Scholar
Ashworth, P. J. & Ferguson, R. I. (1986). Interrelationships of channel processes, changes and sediments in a proglacial braided river. Geografiska Annaler: Series A, Physical Geography 68 (4): 361371. https://doi.org/10.1080/04353676.1986.11880186.CrossRefGoogle Scholar
Baewert, H. & Morche, D. (2014). Coarse sediment dynamics in a proglacial fluvial system (Fagge River, Tyrol). Geomorphology 218 (August): 8897. https://doi.org/10.1016/j.geomorph.2013.10.021.CrossRefGoogle Scholar
Bakker, M., Costa, A., Silva, T. A. Stutenbecker, L., Girarclos, S., Loizeau, J.-L., Molnar, P., Schlunegger, F. & Lane, S. N. (2018). Combined flow abstraction and climate change impacts on an aggrading Alpine river. Water Resources Research 54, 223242.CrossRefGoogle Scholar
Bakker, M., Antoniazza, G., Odermatt, E. & Lane, S. N. (2019). Morphological Response of an Alpine Braided Reach to Sediment‐Laden Flow Events. Journal of Geophysical Research: Earth Surface 124 (5): 13101328. https://doi.org/10.1029/2018JF004811.CrossRefGoogle Scholar
Ballantyne, C. K. (2002a). A General Model of Paraglacial Landscape Response. The Holocene 12 (3): 371376. https://doi.org/10.1191/0959683602hl553fa.CrossRefGoogle Scholar
Ballantyne, C. K. (2002b). Paraglacial Geomorphology. Quaternary Science Reviews 21 (18–19): 19352017. https://doi.org/10.1016/S0277-3791(02)00005-7.CrossRefGoogle Scholar
Ballantyne, C. K. & Benn, D. I. (1994). Paraglacial slope adjustment and resedimenfation following recent glacier retreat, Fåbergstølsdalen, Norway. Arctic and Alpine Research 26: 255269.CrossRefGoogle Scholar
Ballantyne, C. K. & Benn, D. I. (1996). Paraglacial slope adjustment during recent deglaciation and its implications for slope evolution in formerly glaciated environments. Anderson, M. G. and Brooks, S., Editors, Advances in Hillslope Processes 2: 11731195.Google Scholar
Ballantyne, Colin K. & Kirkbride, M. P. (1986). The characteristics and significance of some lateglacial protalus ramparts in upland Britain. Earth Surface Processes and Landforms 11 (6): 659671. https://doi.org/10.1002/esp.3290110609.CrossRefGoogle Scholar
Benn, D. I., Bolch, T., Hands, K., Gulley, J., Luckman, A., Nicholson, L. I., Quincey, D., Thompson, S., Toumi, R. & Wiseman, S. (2012). Response of debris-covered glaciers in the mount everest region to recent warming, and implications for outburst flood hazards. Earth-Science Reviews 114 (1–2): 156174. https://doi.org/10.1016/j.earscirev.2012.03.008.CrossRefGoogle Scholar
Bertoldi, W., Zanoni, L. & Tubino, M. (2010). Assessment of morphological changes induced by flow and flood pulses in a gravel bed braided river: The Tagliamento River (Italy). Geomorphology 114 (3): 348360. https://doi.org/10.1016/j.geomorph.2009.07.017.CrossRefGoogle Scholar
Beylich, A. A., Laute, K., Liermann, S., Hansen, L., Burki, V., Vatne, G., Fredin, O., Gintz, D. & Berthling, I (2009). Subrecent sediment dynamics and sediment budget of the braided sandur system at Sandane, Erdalen (Nordfjord, Western Norway). Norsk Geografisk Tidsskrift – Norwegian Journal of Geography 63 (2): 123131. https://doi.org/10.1080/00291950902907934.CrossRefGoogle Scholar
Beylich, A. A., Laute, K. & Storms, J. E. A. (2017). Contemporary suspended sediment dynamics within two partly glacierized mountain drainage basins in western Norway (Erdalen and Bødalen, Inner Nordfjord). Geomorphology 287 (June): 126143. https://doi.org/10.1016/j.geomorph.2015.12.013.CrossRefGoogle Scholar
Blair, R. W. (1994). Moraine and valley wall collapse due to rapid deglaciation in mount cook national park, New Zealand. Mountain Research and Development 14 (4): 347. https://doi.org/10.2307/3673731.CrossRefGoogle Scholar
Bochet, E., Poesen, J. & Rubio, J. L. (2000). Mound development as an interaction of individual plants with soil, water erosion and sedimentation processes on slopes. Earth Surface Processes and Landforms 25: 847867.3.0.CO;2-Q>CrossRefGoogle Scholar
Bogen, J. (2010). Sediment dynamics of glacier-fed rivers. IAHS-AISH Publication, 181–188.Google Scholar
Borselli, L., Cassi, P., & Torri, D. (2008). Prolegomena to Sediment and Flow Connectivity in the Landscape: A GIS and Field Numerical Assessment. CATENA 75 (3): 268–77. https://doi.org/10.1016/j.catena.2008.07.006.CrossRefGoogle Scholar
Bosson, J.-B., Deline, P., Bodin, X., Schoeneich, P., Baron, L., Gardent, M. & Lambiel, C. (2015). The influence of ground ice distribution on geomorphic dynamics since the little ice age in proglacial areas of two cirque glacier systems: Influence of ground ice on geomorphic dynamics in proglacial areas. Earth Surface Processes and Landforms 40 (5): 666680. https://doi.org/10.1002/esp.3666.CrossRefGoogle Scholar
Bratlie, B. (1994). Senkvartære Sedimenter Og Glasialhistorie i Van Keulenfjorden, Svalbard. Master’s Thesis, Universitetet i Oslo, Norway.Google Scholar
Brönnimann, S., Rajczak, J., Fischer, E. M., Raible, C. C., Rohrer, M. & Schär, C. (2018). Changing seasonality of moderate and extreme precipitation events in the alps. Natural Hazards and Earth System Sciences 18 (7): 20472056. https://doi.org/10.5194/nhess-18-2047-2018.CrossRefGoogle Scholar
Brooks, G. R. (1994). The fluvial reworking of late Pleistocene drift, Squamish river drainage basin, southwestern British Colombia. Géographie Physique et Quaternaire 48 (1): 5168. https://doi.org/10.7202/032972ar.CrossRefGoogle Scholar
Cammeraat, L. H. & Imeson, A. C. (1999). The evolution and significance of soil-vegetation patterns following land abandonment and fire in Spain. CATENA 37: 107127.CrossRefGoogle Scholar
Carling, P. A. (2009). Morphology, sedimentology and palaeohydraulic significance of large gravel dunes, Altai Mountains, Siberia. Sedimentology 43 (4): 647664. https://doi.org/10.1111/j.1365-3091.1996.tb02184.x.CrossRefGoogle Scholar
Carling, P. A., Kirkbride, A. D., Parnachov, S., Borodavko, P. S. & Berger, G. W. (2002). Late quaternary catastrophic flooding in the Altai mountains of south–central Siberia: A synoptic overview and an introduction to flood deposit sedimentology. In Flood and Megaflood Processes and Deposits, edited by Martini, I. Peter, Baker, Victor R., and Garzn, Guillermina, 1735. Oxford, UK: Blackwell Publishing Ltd. https://doi.org/10.1002/9781444304299.ch2.CrossRefGoogle Scholar
Carrivick, J. L., Geilhausen, M., Warburton, J., Dickson, N. E., Carver, S. J., Evans, A. J. & Brown, L. E. (2013). Contemporary geomorphological activity throughout the proglacial area of an alpine catchment. Geomorphology 188 (April): 8395. https://doi.org/10.1016/j.geomorph.2012.03.029.CrossRefGoogle Scholar
Carrivick, J. L. & Heckmann, T. (2017). Short-term geomorphological evolution of proglacial systems. Geomorphology 287 (June): 328. https://doi.org/10.1016/j.geomorph.2017.01.037.CrossRefGoogle Scholar
Carrivick, J. L., Pringle, J. K., Russell, A. J. & Cassidy, N. J. (2007). GPR-derived sedimentary architecture and stratigraphy of outburst flood sedimentation within a bedrock valley system, Hraundalur, Iceland. Journal of Environmental and Engineering Geophysics 12 (1): 127143. https://doi.org/10.2113/JEEG12.1.127.CrossRefGoogle Scholar
Carrivick, J. L. & Rushmer, E. L. (2009). Inter- and intra-catchment variations in proglacial geomorphology: An example from Franz Josef Glacier and Fox Glacier, New Zealand. Arctic, Antarctic, and Alpine Research 41 (1): 1836. https://doi.org/10.1657/1523-0430-41.1.18.CrossRefGoogle Scholar
Carrivick, J. L. & Russel, A. J. (2013). Glaciofluvial landforms of deposition. In Elias, S. A. (ed.), The Encyclopedia of Quaternary Science. Amsterdam: Elsevier, 2: 617.CrossRefGoogle Scholar
Carrivick, J. L., Russell, A. J., Tweed, F. S. & Twigg, D. (2004). Palaeohydrology and sedimentary impacts of Jökulhlaups from Kverkfjöll, Iceland. Sedimentary Geology 172 (1–2): 1940. https://doi.org/10.1016/j.sedgeo.2004.07.005.CrossRefGoogle Scholar
Carrivick, J. L. & Tweed, F. S. (2013). Proglacial lakes: Character, behaviour and geological importance. Quaternary Science Reviews 78 (October): 3452. https://doi.org/10.1016/j.quascirev.2013.07.028.CrossRefGoogle Scholar
Cavalli, M., Trevisani, S., Comiti, F. & Marchi, L. (2013). Geomorphometric assessment of spatial sediment connectivity in small alpine catchments. Geomorphology 188 (April): 3141. https://doi.org/10.1016/j.geomorph.2012.05.007.CrossRefGoogle Scholar
Chiarle, M., Iannotti, S., Mortara, G. & Deline, P. (2007). Recent debris flow occurrences associated with glaciers in the Alps. Global and Planetary Change 56 (1–2): 123136. https://doi.org/10.1016/j.gloplacha.2006.07.003.CrossRefGoogle Scholar
Chiu, Y.-F., Tfwala, S. S., Hsu, Y.-C., Chiu, Y.‐Y., Lee, C.‐Y. & Chen, S.‐C. (2021). Upstream morphological effects of a sequential check dam adjustment process. Earth Surface Processes and Landforms 46 (13): 25272539. https://doi.org/10.1002/esp.5178.CrossRefGoogle Scholar
Church, M. & Ryder, J. M. (1972). Paraglacial sedimentation: A consideration of fluvial processes conditioned by glaciation. Geological Society of America Bulletin 83 (10): 3059. https://doi.org/10.1130/0016-7606(1972)83[3059:PSACOF]2.0.CO;2.CrossRefGoogle Scholar
Church, M. & Slaymaker, O. (1989). Disequilibrium of holocene sediment yield in glaciated British Columbia. Nature 337 (6206): 452454. https://doi.org/10.1038/337452a0.CrossRefGoogle Scholar
Clague, J. (2000). A Review of Catastrophic Drainage of Moraine-Dammed Lakes in British Columbia. Quaternary Science Reviews 19 (17–18): 17631783. https://doi.org/10.1016/S0277-3791(00)00090-1.CrossRefGoogle Scholar
Cossart, É. (2008). Landform Connectivity and Waves of Negative Feedbacks during the Paraglacial Period, a Case Study: The Tabuc Subcatchment since the End of the Little Ice Age (Massif Des Écrins, France). Géomorphologie: Relief, Processus, Environnement 14 (4): 249260. https://doi.org/10.4000/geomorphologie.7430.CrossRefGoogle Scholar
Cossart, E. & Fort, M. (2008). Sediment release and storage in early deglaciated areas: Towards an application of the exhaustion model from the case of Massif Des Écrins (French Alps) since the Little Ice Age. Norsk Geografisk Tidsskrift – Norwegian Journal of Geography 62 (2): 115131. https://doi.org/10.1080/00291950802095145.CrossRefGoogle Scholar
Cossart, É. & Fressard, M. (2017). Assessment of structural sediment connectivity within catchments: Insights from graph theory. Earth Surface Dynamics 5 (2): 253268. https://doi.org/10.5194/esurf-5-253-2017.CrossRefGoogle Scholar
Cossart, É., Viel, V., Lissak, C., Reulier, R., Fressard, M. & Delahaye, D. (2018). How might sediment connectivity change in space and time? Land Degradation & Development 29 (8): 25952613. https://doi.org/10.1002/ldr.3022.CrossRefGoogle Scholar
Costa, A., Anghileri, D. & Molnar, P. (2018). Hydroclimatic control on suspended sediment dynamics of a regulated alpine catchment: A conceptual approach. Hydrology and Earth System Sciences 22 (6): 34213434. https://doi.org/10.5194/hess-22-3421-2018.CrossRefGoogle Scholar
Costa, A., Molnar, P., Stutenbecker, L., Bakker, M., Silva, T. A., Schlunegger, F., Lane, S. N., Loizeau, J.-L. & Girardclos, S. (2018). Temperature signal in suspended sediment export from an alpine catchment. Hydrology and Earth System Sciences 22 (1): 509528. https://doi.org/10.5194/hess-22-509-2018.CrossRefGoogle Scholar
Coulthard, T. J. & Van de Wiel, M. J. (2013). Climate, tectonics or morphology: What signals can we see in drainage basin sediment yields? Earth Surface Dynamics 1 (1): 1327. https://doi.org/10.5194/esurf-1-13-2013.CrossRefGoogle Scholar
Cruden, D. M. & Hu, X. Q. (1993). Exhaustion and steady state models for predicting landslide hazards in the canadian rocky mountains. Geomorphology 8 (4): 279285. https://doi.org/10.1016/0169-555X(93)90024-V.CrossRefGoogle Scholar
Curry, A. M., Cleasby, V. & Zukowskyj, P. (2006). Paraglacial response of steep, sediment-mantled slopes to post-“Little Ice Age” glacier recession in the central Swiss Alps. Journal of Quaternary Science 21 (3): 211225. https://doi.org/10.1002/jqs.954.CrossRefGoogle Scholar
Curry, A. M., Sands, T. B. & Porter, P. R. (2009). Geotechnical Controls on a Steep Lateral Moraine Undergoing Paraglacial Slope Adjustment. Geological Society, London, Special Publications 320 (1): 181197. https://doi.org/10.1144/SP320.12.CrossRefGoogle Scholar
Dabney, S.M, Liu, Z., Lane, M., Douglas, J., Zhu, J. & Flanagan, D. C. (1999). Landscape benching from tillage erosion between grass hedges1 paper presented at international symposium on tillage translocation and tillage erosion held in conjunction with the 52nd annual conference of the soil and water conservation society, Toronto, Canada, 24–25 July 1997.1. Soil and Tillage Research 51 (3–4): 219231. https://doi.org/10.1016/S0167-1987(99)00039-2.CrossRefGoogle Scholar
Delaloye, R. & Morard, S. (2019). Géomorphologie Périglaciaire. Géographie Dép. Géosciences, Université de Fribourg – GG.02262–263.Google Scholar
Delaloye, R., Strozzi, T., Lambiel, C., Perruchoud, E. & Raetzo, H. (2007). Landslide-like development of rockglaciers detected with ERS-1/2 SAR Interfero-Metry. Presented at the Proceedings of the Frascati, Italy, FRINGE 2007 Work-shop, 26–30 November 2007 (ESA SP-649, February 2008).Google Scholar
Delmas, M., Calvet, M. & Gunnell, Y. (2009). Variability of quaternary glacial erosion rates – a global perspective with special reference to the eastern Pyrenees. Quaternary Science Reviews 28 (5–6): 484498. https://doi.org/10.1016/j.quascirev.2008.11.006.CrossRefGoogle Scholar
Desloges, J. R. (1994). Varve deposition and sediment yield record at three small lakes of the southern Canadian Cordillera. Arctic and Alpine Research 26: 130140.CrossRefGoogle Scholar
Dirszowsky, R. W. & Desloges, J. R. (1997). Glaciolacustrine sediments and neoglacial history of the Chephren Lake Basin, Banff National Park, Alberta. Geographie Physique et Quaternaire 51: 4153.CrossRefGoogle Scholar
Draut, A. E., Logan, J. B. & Mastin, M. C., (2011). Channel evolution on the dammed Elwha River, Washington, USA. Geomorphology, 127, 7187CrossRefGoogle Scholar
Dunne, T. & Leopold, L. B. (1978). Water in Environmental Planning. San Francisco. CA: W. H. Freeman and Co., 818. https://doi.org/10.1002/esp.3290040322.CrossRefGoogle Scholar
Evans, D. J. A., Hiemstra, J. F., Boston, C. M., Leighton, I., Cofaigh, C. Ó. & Rea, B. R. (2012). Till stratigraphy and sedimentology at the margins of terrestrially terminating ice streams: Case study of the western Canadian prairies and high plains. Quaternary Science Reviews 46 (July): 80125. https://doi.org/10.1016/j.quascirev.2012.04.028.CrossRefGoogle Scholar
Fatichi, S., Rimkus, S., Burlando, P., Bordoy, R. & Molnar, P. (2015). High-resolution distributed analysis of climate and anthropogenic changes on the hydrology of an alpine catchment. Journal of Hydrology 525 (June): 362382. https://doi.org/10.1016/j.jhydrol.2015.03.036.CrossRefGoogle Scholar
French, H. (2015). Periglacial Environments. Oxford University Press. https://doi.org/10.1093/obo/9780199363445-0038.CrossRefGoogle Scholar
Gabbud, C., Robinson, C. & Lane, S. N., (2019). Summer is in winter: disturbance-driven shifts in macroinvertebrate communities following hydroelectric power exploitation. Science of the Total Environment, 650, 21642180.CrossRefGoogle ScholarPubMed
Gabbud, C. & Lane, S. N. (2016). Ecosystem impacts of alpine water intakes for hydropower: The challenge of sediment management. WIREs Water 3 (1): 4161. https://doi.org/10.1002/wat2.1124.CrossRefGoogle Scholar
Galay, V. J. (1983). Causes of river bed degradation. Water Resources Research, 19(5), 10571090.CrossRefGoogle Scholar
Gärtner-Roer, I. & Bast, A. (2019). (Ground) ice in the proglacial zone. In Heckmann, Tobias & Morche, David (eds.), Geomorphology of Proglacial Systems. 8598. Cham: Springer International Publishing, https://doi.org/10.1007/978-3-319-94184-4_6.CrossRefGoogle Scholar
Geilhausen, M., Otto, J.-C., Morche, D. & Schrott, L. (2012). Decadal sediment yield from an alpine proglacial zone inferred from reservoir sedimentation (Pasterze, Hohe Tauern, Austria). In Erosion and Sediment Yields in the Changing Environment, 161172.Google Scholar
Gobiet, A. & Kotlarski, S. (2020). Future Climate Change in the European Alps. Oxford Research Encyclopedia of Climate Science, Oxford University Press. https://doi.org/10.1093/acrefore/9780190228620.013.767.CrossRefGoogle Scholar
Gurnell, A. M. (1983). Downstream channel adjustments in response to water abstraction for hydro-electric power generation from alpine glacial melt-water streams. The Geographical Journal 149 (3): 342. https://doi.org/10.2307/634009.CrossRefGoogle Scholar
Haeberli, W., Schaub, Y. & Huggel, C. (2017). Increasing risks related to landslides from degrading permafrost into new lakes in de-glaciating mountain ranges. Geomorphology 293 (September): 405417. https://doi.org/10.1016/j.geomorph.2016.02.009.CrossRefGoogle Scholar
Harbor, J. & Warburton, J. (1993). Relative rates of glacial and nonglacial erosion in alpine environments. Arctic and Alpine Research 25 (1): 1-7. https://doi.org/10.2307/1551473.CrossRefGoogle Scholar
Harvey, A M. (2001). Coupling between Hillslopes and Channels in Upland Fluvial Systems: Implications for Landscape Sensitivity, Illustrated from the Howgill Fells, Northwest England, 26.CrossRefGoogle Scholar
Heckmann, T., Cavalli, M. & Marchi, L. (2019). Sediment Connectivity in Proglacial Areas. In Heckmann, T. & Morche, D. (eds) Geomorphology of Proglacial Systems – Landform and Sediment Dynamics in Recently Deglaciated Alpine Landscapes, 271–87. (Geography of the Physical Environment), Cham, Schweiz: Springer.CrossRefGoogle Scholar
Heckmann, T. & Morche, D. (2019). Geomorphology of proglacial systems: landform and sediment dynamics in recently deglaciated alpine landscapes. In Heckmann, T. & Morche, D. (eds.), Geography of the Physical Environment. Cham: Springer International Publishing, https://doi.org/10.1007/978-3-319-94184-4.Google Scholar
Heckmann, T. & Schwanghart, W. (2013). Geomorphic Coupling and Sediment Connectivity in an Alpine Catchment – Exploring Sediment Cascades Using Graph Theory, 15.CrossRefGoogle Scholar
Hedding, D. W. (2011). Pronival rampart and protalus rampart: A review of terminology. Journal of Glaciology 57 (206): 11791180. https://doi.org/10.3189/002214311798843241.CrossRefGoogle Scholar
Hirschberg, J., Fatichi, S., Bennett, G. L., McArdell, B. W., Peleg, N., Lane, S. N., Schlunegger, F. & Molnar, P. (2021). Climate change impacts on sediment yield and debris‐flow activity in an alpine catchment. Journal of Geophysical Research: Earth Surface 126 (1). https://doi.org/10.1029/2020JF005739.Google Scholar
Hugenholtz, C. H., Moorman, B. J., Barlow, J. & Wainstein, P. A. (2008). Large-scale moraine deformation at the Athabasca glacier, Jasper National Park, Alberta, Canada. Landslides 5 (3): 251260. https://doi.org/10.1007/s10346-008-0116-5.CrossRefGoogle Scholar
Huss, M., Bookhagen, B., Huggel, C., Jacobsen, D., Bradley, R. S., Clague, J. J., Vuille, M. & al. (2017). Toward mountains without permanent snow and ice: Mountains without permanent snow and ice. Earth’s Future 5 (5): 418435. https://doi.org/10.1002/2016EF000514.CrossRefGoogle Scholar
Huss, M. & Hock, R. (2018). Global-scale hydrological response to future glacier mass loss. Nature Climate Change 8 (2): 135140. https://doi.org/10.1038/s41558-017-0049-x.CrossRefGoogle Scholar
IPCC. (2021). Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change [Masson-Delmotte, V., P. Zhai, A. Pirani, S. L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M. I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J. B. R. Matthews, T. K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou (Eds.)]. Cambridge University Press. In Press.Google Scholar
Jackson, L. E., MacDonald, G. M. & Wilson, M. C. (1982). Paraglacial origin for terraced river sediments in bow valley, Alberta. Canadian Journal of Earth Sciences 19 (12): 22192231. https://doi.org/10.1139/e82-196.CrossRefGoogle Scholar
Kenner, R., Phillips, M., Limpach, P., Beutel, J., & Hiller, M. (2018). Monitoring mass movements using georeferenced time-lapse photography_ Ritigraben Rock Glacier, Western Swiss Alps. Cold Regions Science and Technology, 145, 127134. https://doi.org/10.1016/j.coldregions.2017.10.018.CrossRefGoogle Scholar
Kondolf, G. M. & Swanson, M. L. (1993). Channel adjustments to reservoir construction and instream gravel mining, Stony Creek, California. Environmental Geology and Water Science, 21, 256269.CrossRefGoogle Scholar
Koppes, M. N. & Montgomery, D. R. (2009). The relative efficacy of fluvial and glacial erosion over modern to orogenic timescales. Nature Geoscience 2 (9): 644647. https://doi.org/10.1038/ngeo616.CrossRefGoogle Scholar
Kroes, D. E. & Hupp, C. R. (2010). The effect of channelization on floodplain sediment deposition and subsidence along the Pocomoke River, Maryland. Journal of the American Water Resources Association, 46 686699CrossRefGoogle Scholar
Kummert, M. & Delaloye, R. (2018). Mapping and quantifying sediment transfer between the front of rapidly moving rock glaciers and torrential gullies. Geomorphology 309 (May): 6076. https://doi.org/10.1016/j.geomorph.2018.02.021.CrossRefGoogle Scholar
Kummert, M., Delaloye, R. & Braillard, L. (2018). Erosion and sediment transfer processes at the front of rapidly moving rock glaciers: Systematic observations with automatic cameras in the western Swiss Alps. Permafrost and Periglacial Processes 29 (1): 2133. https://doi.org/10.1002/ppp.1960.CrossRefGoogle Scholar
Lambiel, C., Delaloye, R., Strozzi, T., Lugon, R. & Raetzo, R. (2008). ERS InSAR for Detecting the Rock Glacier Activity. Presented at the Kane, D. L. & K. M. Hinkel (eds), Proceedings of the 9th International Conference on Permafrost, June 29–July 3, 2008, Fairbanks, Alaska 2: 1019–1024.Google Scholar
Lane, S., Bakker, M., Balin, D., Lovis, B. & Regamey, B. (2014). Climate and Human Forcing of Alpine River Flow. In Schleiss, Anton, de Cesare, Giovanni, Franca, Mario, & Pfister, Michael (eds.), River Flow 2014, 715. CRC Press, Lausanne. https://doi.org/10.1201/b17133-5.CrossRefGoogle Scholar
Lane, S. N., Bakker, M., Costa, A., Girardclos, S., Loizeau, J.-L., Molnar, P., Silva, T., Stutenbecker, L. & Schlunegger, F. (2019). Making stratigraphy in the Anthropocene: Climate change impacts and economic conditions controlling the supply of sediment to Lake Geneva. Scientific Reports 9 (1): 8904. https://doi.org/10.1038/s41598-019-44914-9.CrossRefGoogle ScholarPubMed
Lane, S. N., Gaillet, T. & Goldenschue, L. (2022). Restoring morphodynamics downstream from Alpine dams: development of a geomorphological version of the serial discontinuity concept. Geomorphology, 402. https://doi.org/10.1016/j.geomorph.2022.108131.CrossRefGoogle Scholar
Lane, S. N. & Nienow, P. W. (2019). Decadal-scale climate forcing of alpine glacial hydrological systems. Water Resources Research, 55: 24782492. https://doi.org/10.1029/2018WR024206.CrossRefGoogle Scholar
Lane, S. N., Richards, K. S. & Chandler, J. H. (1996). Discharge and sediment supply controls on erosion and deposition in a dynamic alluvial channel. Geomorphology 1C5 (1): 115. https://doi.org/10.1016/0169-555X(95)00113-J.CrossRefGoogle Scholar
Lane, S. N., Bakker, M. Gabbud, C., Micheletti, N. & Saugy, J.-N. (2017). Sediment export, transient landscape response and catchment-scale connectivity following rapid climate warming and alpine glacier recession. Geomorphology 277 (January): 210227. https://doi.org/10.1016/j.geomorph.2016.02.015.CrossRefGoogle Scholar
Lee, K.-H., Isenhart, T. M., Schultz, R. C. & Mickelson, S. K. (2000). Multispecies riparian buffers trap sediment and nutrients during rainfall simulations. Journal of Environmental Quality 29 (4): 12001205. https://doi.org/10.2134/jeq2000.00472425002900040025x.CrossRefGoogle Scholar
Leonard, E. M. (1985). Glaciological and climatic controls on lake sedimentation, canadian rocky mountains. Zeitschrift Fur. Gletscherkunde Und Glazialgeologie 21: 3542.Google Scholar
Liermann, S., Beylich, A. A. & Welden, A. (2012). Contemporary suspended sediment transfer and accumulation processes in the small proglacial Sætrevatnetsub-catchment, Bødalen, western Norway. Geomorphology, 167: 91101.CrossRefGoogle Scholar
Livingstone, S. J., Clark, C. D., Piotrowski, J. A., Tranter, M., Bentley, M. J., Hodson, A., Swift, D. A. & Woodward, J. (2012). Theoretical framework and diagnostic criteria for the identification of Palaeo-subglacial lakes. Quaternary Science Reviews 53 (October): 88110. https://doi.org/10.1016/j.quascirev.2012.08.010.CrossRefGoogle Scholar
Łozinzki, W. von. (1909). Über Die Mechanische Verwitterung Der Sandsteine Im Gemässigten Klima. Bulletin International de l’Academie Des Sciences de Cracovie, 1: 125. Classe des Sciences Mathematiques et Naturelles.Google Scholar
Lugon, R. (2010). Rock-glacier dynamics and magnitude–frequency relations of debris flows in a high-elevation watershed: Ritigraben, Swiss Alps. Global and Planetary Change, 73 (3–4): 202210, ISSN 0921-8181, https://doi.org/10.1016/j.gloplacha.2010.06.004.CrossRefGoogle Scholar
Mahoney, D. T., Fox, J. F. & Al Aamery, N. (2018). Watershed erosion modeling using the probability of sediment connectivity in a gently rolling system. Journal of Hydrology 561 (June): 862883. https://doi.org/10.1016/j.jhydrol.2018.04.034.CrossRefGoogle Scholar
Mancini, D. & Lane, S. N. (2020). Changes in sediment connectivity following glacial debuttressing in an alpine valley system. Geomorphology 352 (March): 106987. https://doi.org/10.1016/j.geomorph.2019.106987.CrossRefGoogle Scholar
Marren, P. M. (2005). Magnitude and frequency in proglacial rivers: A geomorphological and sedimentological perspective. Earth-Science Reviews 70 (3–4): 203251. https://doi.org/10.1016/j.earscirev.2004.12.002.CrossRefGoogle Scholar
Marren, P. M. & Toomath, S. C. (2014). Channel pattern of proglacial rivers: Topographic forcing due to glacier retreat: Channel pattern of proglacial rivers. Earth Surface Processes and Landforms 39 (7): 943951. https://doi.org/10.1002/esp.3545.CrossRefGoogle Scholar
Mattson, L. E. & Gardner, J. S. (1991). Mass wasting on valley-side ice-cored moraines, boundary glacier, Alberta, Canada. Geografiska Annaler: Series A, Physical Geography 73 (3–4): 123128. https://doi.org/10.1080/04353676.1991.11880337.CrossRefGoogle Scholar
Meyer, L. D., Dabney, S. M. & Harmon, W. C. (1995). Sediment-trapping effectiveness of stiff-grass hedges. Transactions of the ASAE 38 (3): 809815. https://doi.org/10.13031/2013.27895.CrossRefGoogle Scholar
Meyrat, R. (2018). L’utilisation de la photogrammétrie Structure from Motion pour le suivi des glaciers rocheux. (mémoire de license non publié). Université de Lausanne, Faculté des géosciences et de l’environnement, Suisse.Google Scholar
Micheletti, N., Lambiel, C. & Lane, S. N. (2015). Investigating decadal‐scale geomorphic dynamics in an Alpine mountain setting. Journal of Geophysical Research: Earth Surface 120 (10): 21552175. https://doi.org/10.1002/2015JF003656.CrossRefGoogle Scholar
Micheletti, N. & Lane, S. N. (2016). Water yield and sediment export in small, partially glaciated Alpine watersheds in a warming climate: Climate change control in small Alpine watersheds. Water Resources Research 52 (6): 49244943. https://doi.org/10.1002/2016WR018774.CrossRefGoogle Scholar
Milan, D. J. (2012). Geomorphic impact and system recovery following an extreme flood in an upland stream: Thinhope Burn, Northern England, UK. Geomorphology 138 (1): 319328. https://doi.org/10.1016/j.geomorph.2011.09.017.CrossRefGoogle Scholar
Miller, H. R. & Lane, S. N. (2019). Biogeomorphic feedbacks and the ecosystem engineering of recently deglaciated terrain. Progress in Physical Geography: Earth and Environment 43 (1): 2445. https://doi.org/10.1177/0309133318816536.CrossRefGoogle Scholar
Orwin, J. F. & Smart, C. C. (2004). Short-term spatial and temporal patterns of suspended sediment transfer in proglacial channels, small river glacier, Canada. Hydrological Processes 18 (9): 15211542. https://doi.org/10.1002/hyp.1402.CrossRefGoogle Scholar
Otto, J.-C. (2019). Proglacial lakes in high mountain environments. In Heckmann, Tobias & Morche, David (eds.), Geomorphology of Proglacial Systems. Cham: Springer International Publishing, 231247. https://doi.org/10.1007/978-3-319-94184-4_14.CrossRefGoogle Scholar
Petts, G. E. & Bickerton, M. A. (1994). Influence of water abstraction on the macroinvertebrate community gradient within a glacial stream system: La Borgne d’Arolla, Valais, Switzerland. Freshwater Biology 32 (2): 375386. https://doi.org/10.1111/j.1365-2427.1994.tb01133.x.CrossRefGoogle Scholar
Perolo, P., Bakker, M., Gabbud, C., Moradi, G., Rennie, C. & Lane, S. N. (2019). Subglacial sediment production and snout marginal ice uplift during the late ablation season of a temperate valley glacier. Earth Surface Processes and Landforms, 44, 1117-1136.CrossRefGoogle Scholar
Piton, G., Simon, C., Recking, A., Tacnet, J. M., Liébault, F., Kuss, D., Quefféléan, Y. & Marco, O. (2017). Why do we build check dams in alpine streams? an historical perspective from the French experience: A review of the subtle knowledge of 19th century torrent-control-engineers. Earth Surface Processes and Landforms 42 (1): 91108. https://doi.org/10.1002/esp.3967.CrossRefGoogle Scholar
Porter, P. R., Smart, M. J. & Irvine-Fynn, T. D. L. (2019). Glacial sediment stores and their reworking. In Heckmann, Tobias & Morche, David, (eds.), Geomorphology of Proglacial Systems. Cham: Springer International Publishing, 157176. https://link.springer.com/chapter/10.1007/978-3-319-94184-4_10.CrossRefGoogle Scholar
Rainato, R., Picco, L., Cavalli, M., Mao, L., Neverman, A. J. & Tarolli, P. (2018). Coupling climate conditions, sediment sources and sediment transport in an Alpine basin: Climate, sediment sources and sediment transport in an Alpine basin. Land Degradation & Development 29 (4): 11541166. https://doi.org/10.1002/ldr.2813.CrossRefGoogle Scholar
Ravanel, L. & Deline, P. (2011). Climate influence on rockfalls in high-Alpine steep Rockwalls: The north side of the Aiguilles de Chamonix (Mont Blanc massif) since the end of the “Little Ice Age”. The Holocene 21 (2): 357365. https://doi.org/10.1177/0959683610374887.CrossRefGoogle Scholar
Rebetez, M., Lugon, R. & Baeriswyl, P.-A. (1997). Climatic change and debris flows in high mountain regions: The case study of the Ritigraben Torrent (Swiss Alps). In Diaz, Henry F., Beniston, Martin, & Bradley, Raymond S., (eds.), Climatic Change at High Elevation Sites. Dordrecht: Springer, 139157. https://doi.org/10.1007/978-94-015-8905-5_8.CrossRefGoogle Scholar
Renshaw, C. E., Abengoza, K., Magilligan, F. J., Dade, W. B. & Landis, J. D. (2014). Impact of flow regulation on near-channel floodplain sedimentation. Geomorphology, 205, 120127.CrossRefGoogle Scholar
Rey, F. (2004). Effectiveness of vegetation barriers for Marly sediment trapping. Earth Surface Processes and Landforms 29 (9): 11611169. https://doi.org/10.1002/esp.1108.CrossRefGoogle Scholar
Schmidli, J. & Frei, C. (2005). Trends of heavy precipitation and wet and dry spells in Switzerland during the 20th century. International Journal of Climatology 25 (6): 753771. https://doi.org/10.1002/joc.1179.CrossRefGoogle Scholar
Schmidli, J., Schmutz, C., Frei, C., Wanner, H. & Schär, C. (2002). Mesoscale precipitation variability in the region of the European Alps during the 20th century: Alpine precipitation variability. International Journal of Climatology 22 (9): 10491074. https://doi.org/10.1002/joc.769.CrossRefGoogle Scholar
Schomacker, A. & Kjaer, K. H. (2008). Quantification of dead-ice melting in ice-cored moraines at the high-arctic glacier Holmstr Mbreen, Svalbard. In Boreas 37:211225.CrossRefGoogle Scholar
Schrott, L., Götz, J., Geilhausen, M. & Morche, D. (2006). Spatial and temporal variability of sediment transfer and storage in an Alpine basin (Reintal Valley, Bavarian Alps, Gemany). Geographica Helvetica 61 (3): 191200. https://doi.org/10.5194/gh-61-191-2006.CrossRefGoogle Scholar
Shakesby, R. A. (2004). Protalus ramparts. In Encyclopedia of Geomorphology 1: 813814.Google Scholar
Shugar, D. H., Burr, A., Haritashya, U. K., Kargel, J. S., Watson, C. S., Kennedy, M. C., Bevington, A. R., Betts, R. A., Harrison, S. & Strattman, K. (2020). Rapid worldwide growth of glacial lakes since 1990. Nature Climate Change 10 (10): 939945. https://doi.org/10.1038/s41558-020-0855-4.CrossRefGoogle Scholar
Sloan, J., Miller, J. R. & Lancaster, N. (2001). Response and recovery of the Eel River, California, and its tributaries to floods in 1955, 1964, and 1997. Geomorphology 36 (3–4): 129154. https://doi.org/10.1016/S0169-555X(00)00037-4.CrossRefGoogle Scholar
Smith, V. B. & Mohrig, D. (2017). Geomorphic signature of a dammed Sandy River: The lower trinity river downstream of Livingston dam in Texas, USA. Geomorphology, 297, 122136CrossRefGoogle Scholar
Sorg, A., Mosello, B., Shalpykova, G., Allan, A., Clarvis, M. H. & Stoffel, M. (2014). Coping with changing water resources: The case of the Syr Darya river basin in Central Asia. Environmental Science & Policy 43 (November): 6877. https://doi.org/10.1016/j.envsci.2013.11.003.CrossRefGoogle Scholar
Stahl, K., Moore, R. D., Shea, J. M., Hutchinson, D. & Cannon, A. J. (2008). Coupled modelling of glacier and streamflow response to future climate scenarios: modelling of glacier and streamflow. Water Resources Research 44 (2): 1-13. https://doi.org/10.1029/2007WR005956.CrossRefGoogle Scholar
Stoffel, M. & Huggel, C. (2012). Effects of climate change on mass movements in mountain environments. Progress in Physical Geography: Earth and Environment 36 (3): 421439. https://doi.org/10.1177/0309133312441010.CrossRefGoogle Scholar
Stoffel, M., Mendlik, T., Schneuwly-Bollschweiler, M. & Gobiet, A. (2014). Possible impacts of climate change on debris-flow activity in the Swiss Alps. Climatic Change 122 (1–2): 141155. https://doi.org/10.1007/s10584-013-0993-z.CrossRefGoogle Scholar
Stott, T. & Mount, N. (2007). Alpine proglacial suspended sediment dynamics in warm and cool ablation seasons: implications for global warming. Journal of Hydrology 332 (3–4): 259270. https://doi.org/10.1016/j.jhydrol.2006.07.001.CrossRefGoogle Scholar
Sturm, M. (1986). Formation of a strandline during the 1984 Jokulhlaup of strandline lake. Artic 39: 267269.Google Scholar
Stutenbecker, L., Delunel, R., Schlunegger, F., Silva, T.A., Šegvić, B., Girardclos, S. Bakker, M. & et al. (2018). Reduced sediment supply in a fast eroding landscape? A multi-proxy sediment budget of the Upper Rhône Basin, Central Alps. Sedimentary Geology 375 (November): 105–19. https://doi.org/10.1016/j.sedgeo.2017.12.013.CrossRefGoogle Scholar
Surian, N., Righini, M., Lucía, A., Nardi, L., Amponsah, W., Benvenuti, M., Borga, M. & al. (2016). Channel response to extreme floods: insights on controlling factors from six mountain rivers in northern Apennines, Italy. Geomorphology 272 (November): 7891. https://doi.org/10.1016/j.geomorph.2016.02.002.CrossRefGoogle Scholar
Swift, D. A., Nienow, P. W. & Hoey, T. B. (2005). Basal sediment evacuation by subglacial meltwater: Suspended sediment transport from Haut Glacier d’Arolla, Switzerland. Earth Surface Processes and Landforms 30 (7): 867883. https://doi.org/10.1002/esp.1197.CrossRefGoogle Scholar
Swisstopo, Office fédérale de topographie, ed. (2004). DHM25: The Digital Height Model of Switzerland.Google Scholar
Syverson, K. M. (1998). Sediment record of short-lived ice-contact lakes, Burroughs glacier, Alaska. In Boreas 27:4454.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. The Journal of Geology 115 (1): 119. https://doi.org/10.1086/509246.CrossRefGoogle Scholar
Vörösmarty, C. J, Meybeck, M., Fekete, B., Sharma, K., Green, P. & Syvitski, J. P. M. (2003). Anthropogenic sediment retention: Major global impact from registered river impoundments. Global and Planetary Change 39 (1–2): 169190. https://doi.org/10.1016/S0921-8181(03)00023-7.CrossRefGoogle Scholar
Warburton, J. (1990). An Alpine proglacial fluvial sediment budget. Geografiska Annaler: Series A, Physical Geography 72 (3–4): 261272. https://doi.org/10.1080/04353676.1990.11880322.CrossRefGoogle Scholar
Weber, M., Braun, L., Mauser, W. & Prasch, M. (2010). Contribution of rain, snow and icemelt in the upper Danube discharge today and in the future. Supplementi di Geografia Fisica e Dinamica Quaternaria 33: 221230.Google Scholar
Westoby, M. J., Glasser, N. F., Brasington, J., Hambrey, M. J., Quincey, D. J. & Reynolds, J. M. (2014). Modelling outburst floods from moraine-dammed glacial lakes. Earth-Science Reviews 134 (July): 137159. https://doi.org/10.1016/j.earscirev.2014.03.009.CrossRefGoogle Scholar
Whalley, W. B. (2003). Rock glaciers and protalus landforms: Analogous forms and ice sources on Earth and Mars. Journal of Geophysical Research 108 (E4): 8032. https://doi.org/10.1029/2002JE001864.CrossRefGoogle Scholar
Williams, G. P. & Wolman, M. G. (1984). Downstream effects of dams on alluvial rivers. US Geological Survey Professional Paper 1286, 83 pp.CrossRefGoogle Scholar
Winsemann, J., Brandes, C. & Polom, U. (2011). Response of a proglacial delta to rapid high-amplitude lake-level change: An integration of outcrop data and high-resolution shear wave seismics. Basin Research 23 (1): 2252. https://doi.org/10.1111/j.1365-2117.2010.00465.x.CrossRefGoogle Scholar
Worni, R., Huggel, C., Clague, J. J., Schaub, Y. & Stoffel, M. (2014). Coupling glacial lake impact, dam breach, and flood processes: A modeling perspective. Geomorphology 224 (November): 161176. https://doi.org/10.1016/j.geomorph.2014.06.031.CrossRefGoogle Scholar
Zappa, M. & Kan, C. (2007). Extreme heat and runoff extremes in the Swiss Alps. Natural Hazards and Earth System Sciences 7 (3): 375389. https://doi.org/10.5194/nhess-7-375-2007.CrossRefGoogle Scholar

References

Aalto, R., Lauer, J. W. & Dietrich, W. E., 2008. Spatial and temporal dynamics of sediment accumulation and exchange along Strickland River floodplains (Papua New Guinea) over decadal-to-centennial timescales. Journal of Geophysical Research: Earth Surface, 113(F1), F01S04. doi:10.1029/2006JF000627.CrossRefGoogle Scholar
Abernethy, B. & Rutherfurd, I. D., 2001. The distribution and strength of riparian tree roots in relation to riverbank reinforcement. Hydrological Processes, 15(1), 6379.CrossRefGoogle Scholar
Adams, P. N., Slingerland, R. L. & Smith, N. D., 2004. Variations in natural levee morphology in anastomosed channel flood plain complexes. Geomorphology, 61(1–2), 127142.CrossRefGoogle Scholar
Allison, M. A., Kuehl, S. A., Martin, T. C. & Hassan, A., 1998. Importance of flood-plain sedimentation for river sediment budgets and terrigenous input to the oceans: Insights from the Brahmaputra-Jamuna River. Geology, 26(2), 175178.2.3.CO;2>CrossRefGoogle Scholar
Anderson, J. B., Wallace, D. J., Simms, A. R., Rodriguez, A. B. & Milliken, K. T., 2014. Variable response of coastal environments of the northwestern Gulf of Mexico to sea-level rise and climate change: Implications for future change. Marine Geology, 352, 348366.CrossRefGoogle Scholar
Asahi, K., Shimizu, Y., Nelson, J. & Parker, G., 2013. Numerical simulation of river meandering with self-evolving banks. Journal of Geophysical Research: Earth Surface, 118, 22082229. doi:10.1002/jgrf.20150.CrossRefGoogle Scholar
Asselman, N. E. & Middelkoop, H., 1995. Floodplain sedimentation: Quantities, patterns and processes. Earth Surface Processes and Landforms, 20(6), 481499.CrossRefGoogle Scholar
Baitis, E., 2008. Grain sizes of recent siliciclastic deposits in Wax Lake Delta, Louisiana. Thesis (B.S.), The University of Texas at Austin, 26 p.Google Scholar
Bevington, A. E., Twilley, R. R., Sasser, C. E. & HolmJr, G. O., 2017. Contribution of river floods, hurricanes, and cold fronts to elevation change in a deltaic floodplain, northern Gulf of Mexico, USA. Estuarine, Coastal and Shelf Science, 191, 188200.CrossRefGoogle Scholar
Bevington, A. E. & Twilley, R. R., 2018. Island edge morphodynamics along a chronosequence in a prograding deltaic floodplain wetland. Journal of Coastal Research, 34(4), 806817.CrossRefGoogle Scholar
Bhattacharya, J. P., 2006. Deltas, in Facies Models Revisited, SEPM Society for Sedimentary Geology. doi:10.2110/pec.06.84.0237.CrossRefGoogle Scholar
Bobrovitskaya, N. M., Zubkova, C. & Meade, R. H., 1996. Discharges and yields of suspended sediment in the Ob’ and Yenisey Rivers of Siberia. In Erosion and Sediment Yield: Global and Regional Perspectives, eds. Walling, D.E. & Webb, B.W., 115123. Wallingford, UK: International Association of Hydrological Sciences Press.Google Scholar
Bracken, L. J., Wainwright, J., Ali, G. A., Tetzlaff, D., Smith, M. W., Reaney, S. M., & Roy, A. G., 2013. Concepts of hydrological connectivity: Research approaches, pathways and future agendas. Earth-Science Reviews, 119. https://doi.org/10.1016/j.earscirev.2013.02.001.CrossRefGoogle Scholar
Bridge, J. S., 2009. Rivers and Floodplains: Forms, Processes, and Sedimentary Record. John Wiley & Sons, Chichester.Google Scholar
Brierley, G. J., Ferguson, R. J. & Woolfe, K. J., 1997. What is a fluvial levee? Sedimentary Geology, 114, 19. doi:10.1016/S0037-0738(97)00114-0.CrossRefGoogle Scholar
Cahoon, D. R., White, D. A. & Lynch, J. C., 2011. Sediment infilling and wetland formation dynamics in an active crevasse splay of the Mississippi River delta. Geomorphology, 131(3–4), 5768.CrossRefGoogle Scholar
Carle, M. V., Sasser, C. E. & Roberts, H. H., 2015. Accretion and vegetation community change in the Wax Lake Delta following the historic 2011 Mississippi River flood. Journal of Coastal Research, 31(3), 569587.CrossRefGoogle Scholar
Cazanacli, D. & Smith, N. D., 1998. A study of morphology and texture of natural levees – Cumberland Marshes, Saskatchewan, Canada. Geomorphology, 25(1–2), 4355.CrossRefGoogle Scholar
Chow, V. T., 1959. Open-Channel Hydraulics. New York, McGraw-Hill, 680 p.Google Scholar
Christiansen, T., Wiberg, P. L. & Milligan, T. G., 2000. Flow and sediment transport on a tidal salt marsh surface. Estuarine, Coastal and Shelf Science, 50(3), 315331.CrossRefGoogle Scholar
Cowell, P. J. & Thom, B. G., 1994. Morphodynamics of coastal evolution. In Coastal Evolution: Late Quaternary Shoreline Morphodynamics, (eds.) Carter, R. W. G., Woodroffe, C. D., 3386, Cambridge: Cambridge University Press.Google Scholar
Czuba, J. A., David, S. R., Edmonds, D. A. & Ward, A. S., 2019. Dynamics of surface-water connectivity in a low-gradient meandering river floodplain. Water Resources Research, 55, 18491870, doi:10.1029/2018WR023527.CrossRefGoogle Scholar
Darby, S. E., Alabyan, A. M. & Van de Wiel, M. J. 2008b. Numerical simulation of bank erosion and channel migration in meandering rivers. Water Resources Research, 38, 1163, doi:10.1029/2001WR000602, 2002.Google Scholar
David, S. R., Edmonds, D. A. & Letsinger, S. L., 2017. Controls on the occurrence and prevalence of floodplain channels in meandering rivers: Controls on floodplain channels in Meandering Rivers. Earth Surface Processes and Landforms, 42(3), 460472. https://doi.org/10.1002/esp.4002.CrossRefGoogle Scholar
Day, G., Dietrich, W. E., Rowland, J. C. & Marshall, A., 2008. The depositional web on the floodplain of the Fly River, Papua New Guinea. Journal of Geophysical Research: Earth Surface, 113, 119. doi:10.1029/2006JF000622.CrossRefGoogle Scholar
Dean, D. J., Topping, D. J., Schmidt, J. C., Griffiths, R. E. & Sabol, T. A., 2016. Sediment supply versus local hydraulic controls on sediment transport and storage in a river with large sediment loads. Journal of Geophysical Research: Earth Surface, 121, 82110. doi:10.1002/2015JF003436.CrossRefGoogle Scholar
Dietrich, W. E. & Perron, J. T., 2006. The search for a topographic signature of life. Nature, 439(7075), 411418.CrossRefGoogle ScholarPubMed
Dong, T. Y., McElroy, J. A. N. B., Il’icheva, E., Pavlov, M., Ma, H., Moodie, A. J. & Moreido, V. M., 2020. Predicting water and sediment partitioning in a delta channel network under varying discharge conditions. Water Resources Research, 56(11), p.e2020WR027199. https://doi.org/10.1029/2020WR027199.CrossRefGoogle Scholar
Draut, A. E., Kineke, G. C., Velasco, D. W., Allison, M. A. & Prime, R. J., 2005. Influence of the Atchafalaya River on recent evolution of the chenier-plain inner continental shelf, northern Gulf of Mexico. Continental Shelf Research, 25(1), 91112.CrossRefGoogle Scholar
Edmonds, D. A. L., Caldwell, R., Brondizio, E. S. & Mo Siani, S., 2020. Coastal flooding will disproportionately impact people on river deltas. Nature Communications, 11, 1–8, https://doi.org/10.1038/s41467-020-18531-4.CrossRefGoogle ScholarPubMed
Eke, E., Parker, G. & Shimizu, Y. 2014. Numerical modeling of erosional and depositional bank processes in migrating river bends with self-formed width: morphodynamics of bar push and bank pull. Journal of Geophysical Research: Earth Surface, 119, 1455–1483, doi:10.1002/2013JF003020.Google Scholar
Elliot, T., 1986. Deltas. In Sedimentary Environments: Processes, Facies, and Stratigraphy, (ed.) Reading, H. G., 113154. Blackwell Scientific Publication, Oxford.Google Scholar
Fagherazzi, S., Kirwan, M. L., Mudd, S. M., Guntenspergen, G. R., Temmerman, S., D’Alpaos, A., et al. 2012. Numerical models of salt marsh evolution: Ecological, geomorphic, and climatic factors. Reviews of Geophysics, 50, RG1002. https://doi.org/10.1029/2011RG000359.CrossRefGoogle Scholar
Ferguson, R. J. & Brierley, G. J., 1999. Levee morphology and sedimentology along the lower Tuross River, south‐eastern Australia. Sedimentology, 46(4), 627648.CrossRefGoogle Scholar
Fernandes, A. M., Tornqvist, T. E., Straub, K. M. & Mohrig, D., 2016. Connecting the backwater hydraulics of coastal rivers to fluvio-deltaic sedimentology and stratigraphy. Geology, 44 (12), 979982. doi:10.1130/G37965.1.CrossRefGoogle Scholar
Fisk, H. N. 1952. Geological Investigations of the Atchafalaya Basin and Problem of Mississippi River Diversion. Vicksburg, MS: US Army Corps of Engineers.Google Scholar
Fraticelli, C. M., 2006. Climate forcing in a wave-dominated delta: The effects of drought–flood cycles on delta progradation. Journal of Sedimentary Research, 76(9), 10671076.CrossRefGoogle Scholar
Geleynse, N., Hiatt, M., Sangireddy, H. & Passalacqua, P., 2015. Identifying environmental controls on the shoreline of a natural river delta, Journal of Geophysical Research Earth Surface, 120, 877893, doi:10.1002/2014JF003408.CrossRefGoogle Scholar
Gibling, M. R. & Davies, N. S., 2012. Palaeozoic landscapes shaped by plant evolution. Nature Geoscience, 5(2), 99105.CrossRefGoogle Scholar
Giosan, L., Syvitski, J., Constantinescu, S. & Day, J., 2014. Climate change: Protect the world’s deltas. Nature 516, 3133.CrossRefGoogle ScholarPubMed
Goodbred, S. L. & Kuehl, S. A., 1998. Floodplain processes in the Bengal Basin and the storage of Ganges-Brahmaputra river sediment: An accretion study using 137Cs and 210Pb geochronology. Sedimentary Geology, 121, 239258.CrossRefGoogle Scholar
Goodwell, A. E. & Kumar, P., 2017a. Temporal information partitioning: Characterizing synergy, uniqueness, and redundancy in interacting environmental variables. Water Resources Research, 53(7), 59205942. doi:10.1002/2016WR020216.CrossRefGoogle Scholar
Goodwell, A. E. & Kumar, P., 2017b. Temporal information partitioning networks (TIPNets): A process network approach to infer ecohydrologic shifts. Water Resources Research, 53(7), 58995919. doi:10.1002/2016WR020218.CrossRefGoogle Scholar
Hariharan, J., Piliouras, A., Schwenk, J., & Passalacqua, P., 2022. Width-based discharge partitioning in distributary networks: How right we are, Geophysical Research Letters, 49(14), e2022GL097897. doi:10.1029/2022GL097897.CrossRefGoogle Scholar
Hassenruck-Gudipati, H. J., 2021. Understanding Fluvial Topography: Morphodynamic Processes That Build River Levees and Cut Terraces. The University of Texas at Austin, Dissertation, 115 p.Google Scholar
Hassenruck-Gudipati, H.J., Passalacqua, P. and Mohrig, D., 2022. Natural levees increase in prevalence in the backwater zone: Coastal Trinity River, Texas, USA. Geology, 50(9), 1068–1072.CrossRefGoogle Scholar
Hiatt, M. & Passalacqua, P., 2015. Hydrological connectivity in river deltas: The first-order importance of channel-island exchange, Water Resources Research, 51, 22642282, doi:10.1002/2014WR016149.CrossRefGoogle Scholar
Hiatt, M. & Passalacqua, P., 2017. What controls the transition from confined to unconfined flow? Analysis of hydraulics in a coastal river delta, Journal of Hydraulic Engineering, 143, p. 6, doi:10.1061/(ASCE)HY.1943-7900.0001309.CrossRefGoogle Scholar
Hickin, E. J., 1979. Concave-bank benches on the Squamish River, British Columbia, Canada. Canadian Journal of Earth Sciences, 16, 200203, doi:10.1139/e79-018.CrossRefGoogle Scholar
Hirabayashi, Y., Mahendran, R., Koirala, S., Konoshima, L., Yamazaki, D., Watanabe, S., Kim, H. & Kanae, S., 2013. Global flood risk under climate change. Nature Climate Change, 3, 816821. doi:10.1038/NCLIMATE1911.CrossRefGoogle Scholar
Hooke, J. M. 1979. An analysis of the processes of river bank erosion. Journal of Hydrology 42, 3962, ISSN 0022-1694, https://doi.org/10.1016/0022-1694(79)90005-2.CrossRefGoogle Scholar
Howard, A., 1992. Modeling channel migration and floodplain sedimentation in meandering streams. In Lowland Floodplain Rivers: Geomorphological Perspectives, (eds.) Carlingand, P. A. & Petts, G. E., pp. 141, Hoboken NJ: John Wiley & Sons Ltd.Google Scholar
Howard, A. D. & Knutson, T. R., 1984. Sufficient conditions for river meandering: A simulation approach. Water Resources Research, 20, 16591667. doi:10.1029/WR020i011p01659.CrossRefGoogle Scholar
Ielpi, A. & Lapôtre, M. G., 2020. A tenfold slowdown in river meander migration driven by plant life. Nature Geoscience, 13(1), 8286.CrossRefGoogle Scholar
Ikeda, S., Parker, G. & Sawai, K., 1981. Bend theory of river meanders: Part 1. Linear development. Journal of Fluid Mechanics, 112, 363377. doi:10.1017/S0022112081000451.CrossRefGoogle Scholar
Janes, V. J. J., Nicholas, A. P., Collins, A. L. et al. 2017. Analysis of fundamental physical factors influencing channel bank erosion: results for contrasting catchments in England and Wales. Environmental and Earth Science, 76, 307, https://doi.org/10.1007/s12665-017-6593-x.CrossRefGoogle Scholar
Kesel, R. H., Dunne, K. C., McDonald, R. C., Allison, K. R. & Spicer, B. E., 1974. Lateral erosion and overbank deposition on the Mississippi River in Louisiana caused by 1973 flooding. Geology, 2(9), 461464. doi:10.1130/0091-7613(1974)2h461:LEAODOi2.0.CO;2.2.0.CO;2>CrossRefGoogle Scholar
Kim, W., Mohrig, D., Twilley, R., Paola, C. & Parker, G., 2009. Is it feasible to build new land in the Mississippi River delta?: EOS. Transactions, American Geophysical Union, 90(42), 373384.CrossRefGoogle Scholar
Kirwan, M. L., Guntenspergen, G. R., D’Alpaos, A., Morris, J. T., Mudd, S. M. & Temmerman, S., 2010. Limits on the adaptability of coastal marshes to rising sea level. Geophysical Research Letters, 37, L23401. doi:10.1029/2010GL045489.CrossRefGoogle Scholar
Kirwan, M. L. & Megonigal, J. P., 2013. Tidal wetland stability in the face of human impacts and sea-level rise. Nature, 504(7478), 5360.CrossRefGoogle ScholarPubMed
Kirwan, M. L. & Murray, A. B. 2007. A coupled geomorphic and ecological model of tidal marsh evolution. Proceedings of the National Academy of Sciences, 104(15), 61186122.CrossRefGoogle ScholarPubMed
Kleinhans, M. G., de Vries, B., Braat, L. & van Oorschot, M., 2018. Living landscapes: Muddy and vegetated floodplain effects on fluvial pattern in an incised river. Earth Surface Processes and Landforms, 43(14), 29482963. https://doi.org/10.1002/esp.4437.CrossRefGoogle Scholar
Lane, E. W., 1957. A Study of the Shape of Channels Formed by Natural Streams Flowing in Erodible Material. Missouri River Division Sediment Series Report 9: Omaha, Nebraska, U.S. Army Corps of Engineers, pp. 1–106.Google Scholar
Larsen, L. G., 2019. Multiscale flow‐vegetation‐sediment feedbacks in low‐gradient landscapes. Geomorphology.CrossRefGoogle Scholar
Lauer, J.W. & Parker, G. 2008a. Net local removal of floodplain sediment by river meander migration. Geomorphology, 96, 123149, ISSN 0169-555X, https://doi.org/10.1016/j.geomorph.2007.08.003.CrossRefGoogle Scholar
Lauer, J. W. & Parker, G. 2008b, Modeling framework for sediment deposition, storage, and evacuation in the floodplain of a meandering river: Theory. Water Resources Research, 44, W04425, doi:10.1029/2006WR005528.Google Scholar
Lauzon, R. & Murray, A. B., 2018. Comparing the cohesive effects of mud and vegetation on delta evolution. Geophysical Research Letters, 45(19), 10437.CrossRefGoogle Scholar
Marani, M., D’Alpaos, A., Lanzoni, S., Carniello, L. & Rinaldo, A., 2010. The importance of being coupled: Stable states and catastrophic shifts in tidal biomorphodynamics. Journal of Geophysical Research: Earth Surface, 115, F04004, doi:10.1029/2009JF001600.CrossRefGoogle Scholar
Mason, J. & Mohrig, D., 2019. Differential bank migration and the maintenance of channel width in meandering river bends. Geology. doi:10.1130/G46651.1.CrossRefGoogle Scholar
Mason, J. & Mohrig, D., 2019. Scroll bars are inner bank levees along meandering river bends. Earth Surface Processes and Landforms. doi:10.1002/esp.4690.CrossRefGoogle Scholar
Mason, J. & Mohrig, D., 2018. Using time-lapse lidar to quantify river bend evolution on the meandering coastal Trinity River, Texas, USA. Journal of Geophysical Research – Earth Surface, 123(5), 11331144, https://doi.org/10.1029/2017JF004492.CrossRefGoogle Scholar
Maun, M. A., 1998. Adaptations of plants to burial in coastal sand dunes. Canadian Journal of Botany, 76(5), 713738.CrossRefGoogle Scholar
Melton, M. A., 1957. An Analysis of the Relations Among Elements of Climate, Surface Properties, and Geomorphology. Technical Report 11 Office of Naval Research Department of Geology, Columbia University.CrossRefGoogle Scholar
Mertes, L. A., 1997. Documentation and significance of the perirheic zone on inundated floodplains. Water Resources Research, 33(7), 17491762.CrossRefGoogle Scholar
Michael, H. A. & Voss, C. I., 2008. Evaluation of the sustainability of deep groundwater as an arsenic-safe resource in the Bengal Basin. Proceedings of the National academy of Sciences, 85318536, doi:10.1073/pnas.0710477105.CrossRefGoogle Scholar
Millard, C., Hajek, E. & Edmonds, D. A., 2017. Evaluating controls on crevasse-splay size: Implications for floodplain-basin filling. Journal of Sedimentary Research, 87(7), 722739.CrossRefGoogle Scholar
Mullenbach, B. L., Nittrouer, C. A., Puig, P. & Orange, D. L., 2004. Sediment deposition in a modern submarine canyon: Eel Canyon, northern California. Marine Geology, 211(1–2), 101119.CrossRefGoogle Scholar
Nardin, W. & Edmonds, D. A., 2014. Optimum vegetation height and density for inorganic sedimentation in deltaic marshes. Nature Geoscience, 7(10), 722726.CrossRefGoogle Scholar
Nardin, W., Edmonds, D. A. & Fagherazzi, S., 2016. Influence of vegetation on spatial patterns of sediment deposition in deltaic islands during flood. Advances in Water Resources, 93, 236248.CrossRefGoogle Scholar
Nienhuis, J. H., Törnqvist, T. E. & Esposito, C. R., 2018. Crevasse Splays Versus Avulsions: A Recipe for Land Building with Levee Breaches. Geophysical Research Letters, 45, 40584067, doi:10.1029/2018GL077933.CrossRefGoogle Scholar
Nittrouer, J. A., Mohrig, D., Allison, M. A. & Peyret, A.-P., 2011. The Lowermost Mississippi River: A Mixed Bedrock-Alluvial Channel: Sedimentology, 58, 19141934, doi:10.1111/j.1365-3091.2011.01245.x.CrossRefGoogle Scholar
Nittrouer, J. A., Mohrig, D. & Allison, M. A., 2011. Punctuated sand transport in the lowermost Mississippi River. Journal of Geophysical Research-Earth Surface, 116, p.F04025, doi:10.1029/2011JF002026.CrossRefGoogle Scholar
Nittrouer, J. A., Shaw, J., Lamb, M. P. & Mohrig, D., 2012. Spatial and temporal trends for water-flow velocity and bed material sediment transport in the lower Mississippi River. Geological Society of America Bulletin, 124, 400414, doi:10.1130/B30497.1.CrossRefGoogle Scholar
North, C. P. & Davidson, S. K., 2012. Unconfined alluvial flow processes: Recognition and interpretation of their deposits, and the significance for palaeogeographic reconstruction. Earth-Science Reviews, 111(1–2), 199223.CrossRefGoogle Scholar
Odezulu, C. I., Swanson, T. & Anderson, J. B., 2021. Holocene progradation and retrogradation of the Central Texas Coast regulated by alongshore and cross‐shore sediment flux variability. The Depositional Record, 7(1), 7792.CrossRefGoogle Scholar
Olliver, E. A., Edmonds, D. A. & Shaw, J. B. 2020. Influence of floods, tides, and vegetation on sediment retention in Wax Lake Delta, Louisiana, USA. Journal of Geophysical Research: Earth Surface, 125, e2019JF005316. https://doi.org/10.1029/2019JF005316.Google Scholar
Page, K. & Nanson, G., 1982. Concave-bank benches and associated floodplain formation. Earth Surface Processes and Landforms, 7, 529543, doi:10.1002/esp.3290070603.CrossRefGoogle Scholar
Paola, C. & Mohrig, D., 1996. Paleohydraulics revisited: Paleoslope estimation in coarse-grained braided rivers. Basin Research, 8, 243254, doi:10.1046/j.1365-2117.1996.00253.x.CrossRefGoogle Scholar
Passalacqua, P., Lanzoni, S., Paola, C. & Rinaldo, A., 2013. Geomorphic signatures of deltaic processes and vegetation: The Ganges-Brahmaputra-Jamuna case study. Journal of Geophysical Research Earth Surface, 118(3), 18381849, doi:10.1002/jgrf.20128.CrossRefGoogle Scholar
Passalacqua, P., 2017. The Delta Connectome: A network-based framework for studying connectivity in river deltas. Geomorphology, 277, 5062, doi:10.1016/j.geomorph.2016.04.001.CrossRefGoogle Scholar
Pearson, S. G., van Prooijen, B. C., Elias, E. P., Vitousek, S. & Wang, Z. B., 2020. Sediment connectivity: A framework for analyzing coastal sediment transport pathways. Journal of Geophysical Research: Earth Surface, 125(10), e2020JF005595.Google Scholar
Perignon, M., Adams, J., Overeem, I. & Passalacqua, P., 2020. Dominant process zones in a mixed fluvial-tidal delta are morphologically distinct. eSurf, 8, 809824, https://doi.org/10.5194/esurf-8-809-2020.Google Scholar
Pethick, J., 1993. Shoreline adjustments and coastal management: Physical and biological processes under accelerated sea-level rise. Geographical Journal, 159, 162168.CrossRefGoogle Scholar
Pierce, A. R. & King, S. L., 2008. Spatial dynamics of overbank sedimentation in floodplain systems. Geomorphology, 100, 256268. doi:10.1016/j.geomorph.2007.12.008.CrossRefGoogle Scholar
Pizzuto, J. E., 1987. Sediment diffusion during overbank flows. Sedimentology, 34, 301317. doi:10.1111/j.1365-3091.1987.tb00779.x.CrossRefGoogle Scholar
Rodriguez, A. B., Hamilton, M. D. & Anderson, J. B., 2000. Facies and evolution of the modern Brazos Delta, Texas: Wave versus flood influence. Journal of Sedimentary Research, 70(2), 283295.CrossRefGoogle Scholar
Rowland, J. C., Dietrich, W. E., Day, G. & Parker, G., 2009. Formation and maintenance of single‐thread tie channels entering floodplain lakes: Observations from three diverse river systems. Journal of Geophysical Research, 114, F02013, doi:10.1029/2008JF001073.CrossRefGoogle Scholar
Rowland, J. C., Lepper, K., Dietrich, W. E., Wilson, C. J. & Sheldon, R., 2005. Tie channel sedimentation rates, oxbow formation age and channel migration rate from optically stimulated luminescence (OSL) analysis of floodplain deposits. Earth Surface Processes and Landforms, 30, 11611179, doi:10.1002/esp.1268.CrossRefGoogle Scholar
Ruddell, B. L. & Kumar, P., 2009a. Ecohydrologic process networks: 1. Identification. Water Resources Research, 45, p. W03419, doi:10.1029/2008WR007279.Google Scholar
Ruddell, B. L. & Kumar, P., 2009b. Ecohydrologic process networks: 2. Analysis and characterization. Water Resources Research, 45, W03420, doi:10.1029/2008WR007280.Google Scholar
Scheidt, C., Fernandes, A., Paola, C. & Caers, J., 2015. Can geostatistical models represent natures variability? An analysis using flume experiments. Petroleum Geostatistics. doi:10.3997/2214-4609.201413624.Google Scholar
Schreiber, T., 2000. Measuring information transfer. Physical Review Letters, 85, 461464. doi:10.1103/PhysRevLett.85.461.CrossRefGoogle ScholarPubMed
Sendrowski, A. & Passalacqua, P., 2017. Process connectivity in a naturally prograding river delta. Water Resources Research, 53, 3, 18411863. doi:10.1002/2016WR019768.CrossRefGoogle Scholar
Sendrowski, A., Sadid, K., Meselhe, E., Wagner, R. W., Mohrig, D. & Passalacqua, P., 2018. Transfer entropy as a tool for hydrodynamic model validation. Entropy, 20, 58. doi:10.3390/e20010058.CrossRefGoogle ScholarPubMed
Shaw, J. B., Estep, J. D., Whaling, A. R., Sanks, K. M. & Edmonds, D. A., 2018. Measuring subaqueous progradation of the Wax Lake Delta with a model of flow direction divergence. Earth Surface Dynamics, 6(4), 11551168.CrossRefGoogle Scholar
Shaw, J. B., Mohrig, D. & Wagner, R. W., 2016. Flow patterns and morphology of a prograding river delta. Journal of Geophysical Research – Earth Surface, 121(2), 372391. doi:10.1002/2015JF003570.CrossRefGoogle Scholar
Shaw, J. B. & Mohrig, D., 2014a. The importance of erosion in distributary channel network growth, Wax Lake Delta, Louisiana, USA. Geology, 42(1), 3134.CrossRefGoogle Scholar
Shaw, J. B. & Mohrig, D., 2014b. Supplemental material: The importance of erosion in distributary channel network growth, Wax Lake Delta, Louisiana, USA. GSA DATA REPOSITORY 2014008.CrossRefGoogle Scholar
Shaw, J. B., Mohrig, D. & Whitman, S. K., 2013. The morphology and evolution of channels on the Wax Lake Delta, Louisiana, USA. Journal of Geophysical Research – Earth Surface, 118, 15621584. doi:10.1002/jgrf.20123.CrossRefGoogle Scholar
Simon, A. & Collison, A. J., 2002. Quantifying the mechanical and hydrologic effects of riparian vegetation on streambank stability. Earth surface processes and landforms, 27(5), 527546.CrossRefGoogle Scholar
Smith, N. D., Cross, T. A., Dufficy, J. P. & Clough, S. R., 1989. Anatomy of an avulsion. Sedimentology, 36(1), 123. https://doi.org/10.1111/j.1365-3091.1989.tb00817.x.CrossRefGoogle Scholar
Smith, D. G., Hubbard, S. M., Lecki, D. A. & Fustic, M., 2009. Counter point bar deposits: Lithofacies and reservoir significance in the meandering modern Peace River and ancient McMurray Formation, Alberta, Canada. Sedimentology, 56, 16551669, doi:10.1111/j.1365-3091.2009.01050.x.CrossRefGoogle Scholar
Smith, B. C., Moffett, K. B. & Mohrig, D., 2020. Short-term ecogeomorphic evolution of a fluvial delta from hindcasting intertidal marsh-top elevations (HIME). Remote Sensing, 12, 1517, doi:10.3390/rs12091517.CrossRefGoogle Scholar
Smith, V., Mason, J. & Mohrig, D., 2020. Reach-scale changes in channel geometry and dynamics due to the coastal backwater effect: The lower Trinity River, Texas. Earth Surface Processes and Landforms, 45(3), 565573, doi:10.1002/esp.4754.CrossRefGoogle Scholar
Smith, N. D. & Pérez-Arlucea, M., 2008. Natural levee deposition during the 2005 flood of the Saskatchewan River. Geomorphology, 101(4), 583594.CrossRefGoogle Scholar
Shields, M. R., Bianchi, T. S., Mohrig, D., Hutchings, J., Kenney, W. F., Kolker, A. S. & Curtis, J. H., 2017. Carbon storage in the Mississippi River Delta enhanced by ecosystem engineering. Nature Geoscience, 10(11), doi:10.1038/NGEO3044.CrossRefGoogle Scholar
Slingerland, R. & Smith, N. D., 1998. Necessary conditions for a meandering-river avulsion. Geology, 26(5), 435438.2.3.CO;2>CrossRefGoogle Scholar
Sun, T., Meakin, P., Jossang, T. & Schwarz, K., 1996. A simulation model for meandering rivers. Water Resources Research, 32, 29372954. doi:10.1029/96WR00998.CrossRefGoogle Scholar
Swanson, K. M., Watson, E., Aalto, R., Lauer, J.W., Bera, M.T., Marshall, A., Taylor, M.P., Apte, S.C. & Dietrich, W.E., 2008. Sediment load and floodplain deposition rates: Comparison of the Fly and Strickland rivers, Papua New Guinea. Journal of Geophysical Research, 113, F01S03, doi:10.1029/2006JF000623.CrossRefGoogle Scholar
Sylvester, Z., Durkin, P. R., Hubbard, S. M. & Mohrig, D., 2021. Autogenic translation and counter-point-bar deposition in meandering rivers. Geological Society of America Bulletin, https://doi.org/10.1130/B35829.1.CrossRefGoogle Scholar
Temmerman, S. & Kirwan, M. L., 2015. Building landwith a rising sea. Science, 349, 588589. http://dx.doi.org/10.1126/science.aac8312.CrossRefGoogle ScholarPubMed
Temmerman, S., Bouma, T. J., Van de Koppel, J., Van der Wal, D., De Vries, M. B. & Herman, P. M. J., 2007. Vegetation causes channel erosion in a tidal landscape. Geology, 35(7), 631634.CrossRefGoogle Scholar
Tessler, Z. D., Vörösmarty, C. J., Grossberg, M., Gladkova, I., Aizenman, H., Syvitski, J. P. M. & Foufoula-Georgiou, E., 2015. Profiling risk and sustainability in coastal deltas of the world. Science, 349, 638643. doi:10.1126/science.aab3574.CrossRefGoogle ScholarPubMed
Törnqvist, T. E. & Bridge, J. S., 2002. Spatial variation of overbank aggradation rate and its influence on avulsion frequency. Sedimentology, 49(5), 891905.CrossRefGoogle Scholar
Tull, N., Passalacqua, P., Hassenruck-Gudipati, H., Rahman, S., Wright, K., Hariharan, J., & Mohrig, D., 2022. Bidirectional river-floodplain connectivity during combined pluvial-fluvial events. Water Resources Research, 58(3), e2021WR030492, https://doi.org/10.1029/2021WR030492.CrossRefGoogle Scholar
van Dijk, W. M., Densmore, A. L., Sinha, R., Singh, A., & Voller, V. R., 2016. Reduced-complexity probabilistic reconstruction of alluvial aquifer stratigraphy, and application to sedimentary fans in northwestern India. Journal of Hydrology, 541, 12411257, https://doi.org/10.1016/j.jhydrol.2016.08.028.CrossRefGoogle Scholar
Wagner, R. W., Lague, D., Mohrig, D., Passalacqua, P., Shaw, J. & Moffett, K., 2017. Elevation change and stability on a prograding delta. Geophysical Research Letters, 44(4), 17861794, doi:10.1002/2016GL072070.CrossRefGoogle Scholar
Walker, N. D. & Hammack, A. B., 2000. Impacts of winter storms on circulation and sediment transport: Atchafalaya-Vermilion Bay Region, Louisiana, U.S.A. Journal of Coastal Research, 16(4), 9961010.Google Scholar
Walters, D. C. & Kirwan, M. L., 2016. Optimal hurricane overwash thickness for maximizing marsh resilience to sea level rise. Ecology and Evolution, 6(9), 29482956.CrossRefGoogle ScholarPubMed
Weight, R. W., Anderson, J. B. & Fernandez, R., 2011. Rapid mud accumulation on the central Texas shelf linked to climate change and sea-level rise. Journal of Sedimentary Research, 81(10), 743764.CrossRefGoogle Scholar
Wolman, M. G. & Leopold, L. B., 1957. River flood plains: Some observations on their formation (No. 282-C, 87–109). US Government Printing Office.CrossRefGoogle Scholar
Wright, K., Hiatt, M. & Passalacqua, P., 2018. Hydrological connectivity in vegetated river deltas: The importance of patchiness below a threshold. Geophysical Research Letters, 45, 10416–10427, https://doi.org/10.1029/2018GL079183.CrossRefGoogle Scholar
Wright, L. D. & Thom, B. G., 1977. Coastal depositional landforms: A morphodynamic approach. Progress in Physical Geography, 1(3), 412459.CrossRefGoogle Scholar
Yamasaki, T. N., de Lima, P. H., Silva, D. F., Cristiane, G. D. A., Janzen, J. G. & Nepf, H. M., 2019. From patch to channel scale: The evolution of emergent vegetation in a channel. Advances in Water Resources, 129, 131145.CrossRefGoogle Scholar
Yang, J. Q., Chung, H., & Nepf, H. M., 2016. The onset of sediment transport in vegetated channels predicted by turbulent kinetic energy. Geophysical Research Letters, 43(21), 11261.CrossRefGoogle Scholar
Yuill, B., Lavoie, D. & Reed, D. J., 2009. Understanding subsidence processes in coastal Louisiana. Journal of Coastal Research, 10054, 2336.CrossRefGoogle Scholar
Yuill, B. T., Khadka, A. K., Pereira, J., Allison, M. A. & Meselhe, E. A., 2016. Morphodynamics of the erosional phase of crevasse-splay evolution and implications for river sediment diversion function. Geomorphology, 259, 1229.CrossRefGoogle Scholar
Zong, L. & Nepf, H., 2010. Flow and deposition in and around a finite patch of vegetation. Geomorphology, 116(3–4), 363372.CrossRefGoogle Scholar

Save book to Kindle

To save this book to your Kindle, first ensure [email protected] is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×