Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-ttngx Total loading time: 0 Render date: 2024-05-15T17:00:00.529Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  30 April 2020

Bruce L. Rhoads
Affiliation:
University of Illinois
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
River Dynamics
Geomorphology to Support Management
, pp. 432 - 506
Publisher: Cambridge University Press
Print publication year: 2020

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Aalto, R., Dunne, T., Guyot, J. L., 2006. Geomorphic controls on Andean denudation rates. Journal of Geology, 114(1), 8599.Google Scholar
Abad, J. D., Garcia, M. H., 2009. Experiments in a high-amplitude Kinoshita meandering channel: 2. Implications of bend orientation on bed morphodynamics. Water Resources Research, 45. 10.1029/2008wr007017.Google Scholar
Abbado, D., Slingerland, R., Smith, N. D., 2005. Origin of anastomosis in the upper Columbia River, British Columbia, Canada. In: Blum, M. D., Marriott, S. B., LeClair, S. F. (eds.), Fluvial Sedimentology VII, International Association of Sedimentologists Special Publication No. 35, pp. 315.Google Scholar
Abban, B., Papanicolaou, A. N., Cowles, M. K., et al., 2016. An enhanced Bayesian fingerprinting framework for studying sediment source dynamics in intensively managed landscapes. Water Resources Research, 52(6), 46464673. 10.1002/2015wr018030.Google Scholar
Abbe, T., Brooks, A., 2011. Geomorphic, engineering, and ecological considerations when using wood in river restoration. In: Simon, A., Bennett, S. J., Castro, J. M. (eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools. American Geophysical Union, Washington, DC, pp. 419451.Google Scholar
Abernethy, B., Rutherfurd, I. D., 2000. Stabilising stream banks with riparian vegetation. Natural Resource Management, 3(2), 29.Google Scholar
Abrahams, A. D., 1977. Factor of relief in evolution of channel networks in mature drainage basins. American Journal of Science, 277(5), 626646.Google Scholar
Abrahams, A. D., 1984. Channel networks: a geomorphological perspective. Water Resources Research, 20(2), 161188.Google Scholar
Abrahams, A. D., Li, G., Atkinson, J. F., 1995. Step-pool streams – adjustment to maximum flow resistance. Water Resources Research, 31(10), 25932602.Google Scholar
Abrams, D. M., Lobkovsky, A. E., Petroff, A. P., et al., 2009. Growth laws for channel networks incised by groundwater flow. Nature Geoscience, 2(3), 193196.CrossRefGoogle Scholar
Adami, L., Bertoldi, W., Zolezzi, G., 2016. Multidecadal dynamics of alternate bars in the Alpine Rhine River. Water Resources Research, 52(11), 89218938. 10.1002/2015wr018228.Google Scholar
Adamowski, K., 1981. Plotting formula for flood frequency. Water Resources Bulletin, 17(2), 197202.Google 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
Agouridis, C. T., Edwards, D. R., Workman, S. R., et al., 2005. Streambank erosion associated with grazing practices in the humid region. Transactions of the Association of Agricultural Engineers, 48(1), 181190.Google Scholar
Ahadiyan, J., Adeli, A., Bahmanpouri, F., Gualtieri, C., 2018. Numerical simulation of flow and scour in a laboratory junction. Geosciences, 8(5), 162.10.3390/geosciences8050162.Google Scholar
Aharonson, O., Zuber, M. T., Rothman, D. H., Schorghofer, N., Whipple, K. X., 2002. Drainage basins and channel incision on Mars. Proceedings of the National Academy of Sciences of the United States of America, 99(4), 17801783.Google Scholar
Ahilan, S., O’Sullivan, J. J., Bruen, M., Brauders, N., Healy, D., 2013. Bankfull discharge and recurrence intervals in Irish rivers. Proceedings of the Institution of Civil Engineers-Water Management, 166(7), 381393.Google Scholar
Ahmed, J., Constantine, J. A., Dunne, T., 2019. The role of sediment supply in the adjustment of channel sinuosity across the Amazon Basin. Geology, 47(9), 807810. 10.1130/G46319.1.Google Scholar
Ahnert, F., 1970. Functional relationships between denudation, relief, and uplift in large mid-latitude drainage basins. American Journal of Science, 268(3), 243263.Google Scholar
Aich, V., Zimmermann, A., Elsenbeer, H., 2014. Quantification and interpretation of suspended-sediment discharge hysteresis patterns: how much data do we need? Catena, 122, 120129.CrossRefGoogle Scholar
Aksoy, H., Kavvas, M. L., 2005. A review of hillslope and watershed scale erosion and sediment transport models. Catena, 64(2–3), 247271. 10.1016/j.catena.2005.08.008.Google Scholar
Allen, G. H., Pavelsky, T. M., 2018. Global extent of rivers and streams. Science, 361(6402), 585587.CrossRefGoogle ScholarPubMed
Allmendinger, N. E., Pizzuto, J. E., Moglen, G. E., Lewicki, M., 2007. A sediment budget for an urbanizing watershed, 1951–1996, Montgomery County, Maryland, U.S.A. Journal of the American Water Resources Association, 43(6), 14831498.Google Scholar
Allred, T. M., Schmidt, J. C., 1999. Channel narrowing by vertical accretion along the Green River near Green River, Utah. Geological Society of America Bulletin, 111(12), 17571772.Google Scholar
Alonso, C. V., 1980. Selecting a formula to estimate sediment transport in non-vegetated channels. Chapter 5. In: Knisel, W. G. (ed.), CREAMS – A Field Scale Model for Chemicals, Runoff, and Erosion from Agriculture Management Systems, Conservation Research Report No. 26. U.S. Department of Agriculture, Washington, DC.Google Scholar
Alonso, C. V., Bennett, S. J., Stein, O. R., 2002. Predicting head cut erosion and migration in concentrated flows typical of upland areas. Water Resources Research, 38(12), 39-1–39-15. 10.1029/2001wr001173.CrossRefGoogle Scholar
American Rivers, 2019. Dam Removal Database, https://figshare.com/articles/_/5234068 (accessed May 14, 2019).Google Scholar
An, H.-P., Chen, S.-C., Chan, H.-C., Hsu, Y., 2013. Dimension and frequency of bar formation in a braided river. International Journal of Sediment Research, 28(3), 358367.Google Scholar
Anderson, D. G., 1970. Effects of urban development on floods in Northern Virginia. U.S. Geological Survey Water-Supply Paper 2001-C, U.S. Government Printing Office, Washington, DC.Google Scholar
Anderson, R. S., Anderson, S. P., 2010. Geomorphology: The Mechanics and Chemistry of Landscape. Cambridge University Press, New York.Google Scholar
Andrews, E. D., 1980. Effective and bankfull discharges of streams in the Yampa River basin, Colorado and Wyoming. Journal of Hydrology, 46(3–4), 311330.Google Scholar
Andrews, E. D., 1983. Entrainment of gravel from naturally sorted riverbed material. Geological Society of America Bulletin, 94(10), 12251231.Google Scholar
Andrews, E. D., 1984. Bed-material entrainment and hydraulic geometry of gravel-bed rivers in Colorado. Geological Society of America Bulletin, 95(3), 371378.2.0.CO;2>CrossRefGoogle Scholar
Andrews, E. D., Erman, D. C., 1986. Persistence in the size distribution of surficial bed material during an extreme snowmelt flood. Water Resources Research, 22(2), 191197.Google Scholar
Anisimov, O., Vandenberghe, J., Lobanov, V., Kondratiev, A., 2008. Predicting changes in alluvial channel patterns in North-European Russia under conditions of global warming. Geomorphology, 98(3–4), 262274.Google Scholar
Annable, W. K., Watson, C. C., Thompson, P. J., 2012. Quasi-equilibrium conditions of urban gravel-bed stream channels in southern Ontario, Canada. River Research and Applications, 28(3), 302325.Google Scholar
Ansari, K., Morvan, H. P., Hargreaves, D. M., 2011. Numerical investigation into secondary currents and wall shear in trapezoidal channels. Journal of Hydraulic Engineering, 137(4), 432440.Google Scholar
Anthony, D. J., Harvey, M. D., 1991. Stage-dependent cross section adjustments in a meandering reach of Fall River, Colorado. Geomorphology, 4(3–4), 187203.CrossRefGoogle Scholar
Arcement, G. J., Schneider, V. R., 1989. Guide for selecting Manning’s roughness coefficients for natural channels and flood plains. U.S. Geological Survey Water-Supply Paper 2339, U.S. Government Printing Office, Washington, DC.Google Scholar
Arnell, N. W., Gosling, S. N., 2016. The impacts of climate change on river flood risk at the global scale. Climatic Change, 134(3), 387401.CrossRefGoogle Scholar
Arnold, C. L., Boison, P. J., Patton, P. C., 1982. Sawmill Brook: an example of rapid geomorphic change related to urbanization. Journal of Geology, 90(2), 155166. 10.1086/628660.Google Scholar
Arrospide, F., Mao, L., Escauriaza, C., 2018. Morphological evolution of the Maipo River in central Chile: influence of instream gravel mining. Geomorphology, 306, 182197.Google 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(4), 22082229. 10.1002/jgrf.20150.CrossRefGoogle Scholar
Ashmore, P. E., 1982. Laboratory modeling of gravel braided stream morphology. Earth Surface Processes and Landforms, 7(3), 201225.Google Scholar
Ashmore, P. E., 1991a. How do gravel-bed rivers braid? Canadian Journal of Earth Sciences, 28(3), 326341.Google Scholar
Ashmore, P. E., 1991b. Channel morphology and bedload pulses in braided, gravel-bed streams. Geografiska Annaler Series A Physical Geography, 73, 3752.Google Scholar
Ashmore, P. E., 1993. Anabranch confluence kinetics and sedimentation processes in gravel-braided streams. In: Best, J. L., Bristow, C. S. (eds.), Braided Rivers. Special Publication No. 75, Geological Society of London, London, pp. 129146.Google Scholar
Ashmore, P. E., 2009. Intensity and characteristic length of braided channel patterns. Canadian Journal of Civil Engineering, 36(10), 16561666.CrossRefGoogle Scholar
Ashmore, P. E., 2013. Morphology and dynamics of braided rivers. In: Shroder, J. (ed.), Treatise on Geomorphology, vol. 9, Fluvial Geomorphology, Wohl, E. (vol. ed.). Academic Press, San Diego, CA, pp. 289312.Google Scholar
Ashmore, P. E., 2015. Towards a sociogeomorphology of rivers. Geomorphology, 251, 149156.CrossRefGoogle Scholar
Ashmore, P. E., Church, M., 1998. Sediment transport and river morphology: a paradigm for study. In: Klingeman, P. C., Beschta, R. L., Komar, P. D., Bradley, J. B. (eds.), Gravel-Bed Rivers in the Environment. Water Resources Publications, Highland Ranch, CO, pp. 115148.Google Scholar
Ashmore, P. E., Day, T. J., 1988. Effective discharge for suspended sediment transport in streams of the Saskatchewan River basin. Water Resources Research, 24(6), 864870.Google Scholar
Ashmore, P. E., Gardner, J. T., 2008. Unconfined confluences in braided rivers. In: Rice, S. P., Roy, A. G., Rhoads, B. L. (eds.), River Confluences, Tributaries and the Fluvial Network. Wiley, Chichester, UK, pp. 119147.Google Scholar
Ashmore, P. E., Parker, G., 1983. Confluence scour in coarse braided streams. Water Resources Research, 19(2), 392402.CrossRefGoogle Scholar
Ashmore, P. E., Sauks, E., 2006. Prediction of discharge from water surface width in a braided river with implications for at-a-station hydraulic geometry. Water Resources Research, 42(3). 10.1029/2005wr003993.CrossRefGoogle Scholar
Ashmore, P. E., Ferguson, R. I., Prestegaard, K. L., Ashworth, P. J., Paola, C., 1992. Secondary flow in anabranch confluences of a braided, gravel-bed stream. Earth Surface Processes and Landforms, 17(3), 299311. 10.1002/esp.3290170308.Google Scholar
Ashmore, P. E., Bertoldi, W., Gardner, J. T., 2011. Active width of gravel-bed braided rivers. Earth Surface Processes and Landforms, 36(11), 15101521. 10.1002/esp.2182.CrossRefGoogle Scholar
Ashworth, P. J., 1996. Mid-channel bar growth and its relationship to local flow strength and direction. Earth Surface Processes and Landforms, 21(2), 103123. 10.1002/(sici)1096–9837(199602)21:2<103::aid-esp569>3.0.co;2-o.3.0.CO;2-O>CrossRefGoogle Scholar
Ashworth, P. J., Ferguson, R. I., 1989. Size-selective entrainment of bed load in gravel bed streams. Water Resources Research, 25(4), 627634. 10.1029/WR025i004p00627.CrossRefGoogle Scholar
Ashworth, P. J., Lewin, J., 2012. How do big rivers come to be different? Earth-Science Reviews, 114(1–2), 84107. 10.1016/j.earscirev.2012.05.003.CrossRefGoogle Scholar
Ashworth, P. J., Ferguson, R. I., Ashmore, P. E., et al., 1992. Measurements in a braided river chute and lobe: 2. Sorting of bed load during entrainment, transport, and deposition. Water Resources Research, 28(7), 18871896. 10.1029/92wr00702.CrossRefGoogle Scholar
Ashworth, P. J., Best, J. L., Roden, J. E., Bristow, C. S., Klaassen, G. J., 2000. Morphological evolution and dynamics of a large, sand braid-bar, Jamuna River, Bangladesh. Sedimentology, 47(3), 533555. 10.1046/j.1365–3091.2000.00305.x.Google Scholar
Ashworth, P. J., Smith, G. H. S., Best, J. L., et al., 2011. Evolution and sedimentology of a channel fill in the sandy braided South Saskatchewan River and its comparison to the deposits of an adjacent compound bar. Sedimentology, 58(7), 18601883.Google Scholar
Aslan, A., Autin, W. J., 1999. Evolution of the Holocene Mississippi River floodplain, Ferriday, Louisiana: insights on the origin of fine-grained floodplains. Journal of Sedimentary Research, 69(4), 800815. 10.2110/jsr.69.800.Google Scholar
Asselman, N. E. M., 2000. Fitting and interpretation of sediment rating curves. Journal of Hydrology, 234(3–4), 228248. 10.1016/s0022-1694(00)00253–5.CrossRefGoogle Scholar
Asselman, N. E. M., Middelkoop, H., 1995. Floodplain sedimentation – quantities, patterns, and processes. Earth Surface Processes and Landforms, 20(6), 481499. 10.1002/esp.3290200602.Google Scholar
Attal, M., Lave, J., 2009. Pebble abrasion during fluvial transport: experimental results and implications for the evolution of the sediment load along rivers. Journal of Geophysical Research – Earth Surface, 114. 10.1029/2009jf001328.Google Scholar
Ayles, C. P., Church, M., 2015. Downstream channel gradation in the regulated Peace River. In: Church, M. (ed.), Regulation of Peace River: A Case Study for River Management. Wiley, Chichester, UK, pp. 3966.Google Scholar
Baartman, J. E. M., Jetten, V. G., Ritsema, C. J., de Vente, J., 2012. Exploring effects of rainfall intensity and duration on soil erosion at the catchment scale using openLISEM: Prado catchment, SE Spain. Hydrological Processes, 26(7), 10341049. 10.1002/hyp.8196.Google Scholar
Bagherimiyab, F., Lemmin, U., 2013. Shear velocity estimates in rough-bed open-channel flow. Earth Surface Processes and Landforms, 38(14), 17141724. 10.1002/esp.3421.CrossRefGoogle Scholar
Bagnold, R. A., 1966. An approach to the sediment transport problem from general physics. U.S. Geological Survey Professional Paper 422-I, U.S. Government Printing Office, Washington, DC.Google Scholar
Bagnold, R. A., 1973. Nature of saltation and of “bed-load” transport in water. Proceedings of the Royal Society of London, Series A – Mathematical Physical and Engineering Sciences, 332(1591), 473504. 10.1098/rspa.1973.0038.Google Scholar
Bagnold, R. A., 1980. An empirical correlation of bedload transport rates in flumes and natural rivers. Proceedings of the Royal Society of London, Series A – Mathematical Physical and Engineering Sciences, 372(1751), 453473. 10.1098/rspa.1980.0122.Google Scholar
Bai, S., Li, J., 2013. Sediment wash-off from an impervious urban land surface. Journal of Hydrologic Engineering, 18(5), 488498. 10.1061/(asce)he.1943–5584.0000654.Google Scholar
Bailey, B. B., 1967. James Hutton – The Founder of Modern Geology. Elsevier, New York.Google Scholar
Bak, P., 1996. How Nature Works: The Science of Self-Organized Criticality. Copernicus, New York.Google Scholar
Baker, V. R., 1977. Stream channel response to floods, with examples from central Texas. Geological Society of America Bulletin, 88(8), 10571071. 10.1130/0016–7606(1977)88<1057:srtfwe>2.0.co;2.Google Scholar
Baker, V. R., Costa, J., 1987. Flood power. In: Mayer, L., Nash, D. B. (eds.), Catastrophic Flooding. Allen and Unwin, London, pp. 122.Google Scholar
Baker, V. R., Kochel, R. C., 1988. Flood sedimentation in bedrock fluvial systems. In: Baker, V. R., Kochel, R. C., Patton, P. C. (eds.), Flood Geomorphology. Wiley and Sons, New York, pp. 123137.Google Scholar
Bakke, P. D., Basdekas, P. O., Dawdy, D. R., Klingeman, P. C., 1999. Calibrated Parker-Klingeman model for gravel transport. Journal of Hydraulic Engineering, 125(6), 657660. 10.1061/(asce)0733–9429(1999)125:6(657).CrossRefGoogle Scholar
Bakke, P. D., Sklar, L. S., Dawdy, D. R., Wang, W. C., 2017. The design of a site-calibrated Parker-Klingeman gravel transport model. Water, 9(6), 441; 10.3390/w9060441.Google Scholar
Baranya, S., Olsen, N. R. B., Jozsa, J., 2015. Flow analysis of a river confluence with field measurements and RANS model with next grid approach. River Research and Applications, 31(1), 2841. 10.1002/rra.2718.Google Scholar
Barman, B., Kumar, B., Sarma, A. K., 2018. Turbulent flow structures and geomorphic characteristics of a mining affected alluvial channel. Earth Surface Processes and Landforms, 43(9), 18111824. 10.1002/esp.4355.CrossRefGoogle Scholar
Barnard, R. S., Melhorn, W. N., 1982. Morphologic and morphometric response to channelization: the case of history of the Big Pine Creek Ditch, Benton County, Iowa. In: Craig, R. G., Craft, J. L. (eds.), Applied Geomorphology. George Allen and Unwin, London, pp. 224239.Google Scholar
Barnes, H. H., 1967. Roughness characteristics of natural channels. U.S. Geological Survey Water-Supply Paper 1849, U.S. Government Printing Office, Washington, DC.Google Scholar
Barry, J. J., Buffington, J. M., Goodwin, P., King, J. G., Emmett, W. W., 2008. Performance of bed-load transport equations relative to geomorphic significance: predicting effective discharge and its transport rate. Journal of Hydraulic Engineering, 134(5), 601615. 10.1061/(asce)0733–9429(2008)134:5(601).Google Scholar
Barth, N. A., Villarini, G., Nayak, M. A., White, K., 2017. Mixed populations and annual flood frequency estimates in the western United States: The role of atmospheric rivers. Water Resources Research, 53(1), 257269. 10.1002/2016wr019064.Google Scholar
Basso, S., Frascati, A., Marani, M., Schirmer, M., Botter, G., 2015. Climatic and landscape controls on effective discharge. Geophysical Research Letters, 42(20), 84418447. 10.1002/2015gl066014.Google Scholar
Bates, B. C., 1990. A statistical log piecewise linear model of at-station hydraulic geometry. Water Resources Research, 26(1), 109118. 10.1029/WR026i001p00109.Google Scholar
Bathurst, J. C., 1985. Flow resistance estimation in mountain rivers. Journal of Hydraulic Engineering, 111(4), 625643.Google Scholar
Bathurst, J. C., 1987. Critical conditions for bed material movement in steep, boulder-bed streams. In: Beschta, R. L., Blinn, T., Grant, G. E., Ice, G. G., Swanson, F. J. (eds.), Erosion and Sedimentation in the Pacific Rim. IAHS Publication No. 165, IAHS Press, Wallingford, UK, pp. 309318.Google Scholar
Bathurst, J. C., 2013. Critical conditions for particle motion in coarse bed materials of nonuniform size distribution. Geomorphology, 197, 170184.CrossRefGoogle Scholar
Bathurst, J. C., Thorne, C. R., Hey, R. D., 1977. Direct measurements of secondary currents in river bends. Nature, 269(5628), 504506.Google Scholar
Bathurst, J. C., Thorne, C. R., Hey, R. D., 1979. Secondary flow and shear stress at river bends. Journal of the Hydraulics Division – ASCE, 105(10), 12771295.Google Scholar
Bathurst, J. C., Benson, I. A., Valentine, E. M., Nalluri, C., 2002. Overbank sediment deposition patterns for straight and meandering flume channels. Earth Surface Processes and Landforms, 27(6), 659665.Google Scholar
Bauer, B. O., Sherman, D. J., Wolcott, J. F., 1992. Sources of uncertainty in shear stress and roughness length estimates derived from velocity profiles. Professional Geographer, 44(4), 453464.Google Scholar
Bayat, E., Rodriguez, J. F., Saco, P. M., et al., 2017. A tale of two riffles: using multidimensional, multifractional, time-varying sediment transport to assess self-maintenance in pool-riffle sequences. Water Resources Research, 53(3), 20952113. 10.1002/2016wr019464.Google Scholar
Baynes, E. R. C., Lague, D., Attal, M., et al., 2018a. River self-organisation inhibits discharge control on waterfall migration. Scientific Reports, 8. 10.1038/s41598-018–20767-6.Google Scholar
Baynes, E. R. C., van de Lageweg, W. I., McLelland, S. J., et al., 2018b. Beyond equilibrium: Re-evaluating physical modelling of fluvial systems to represent climate changes. Earth-Science Reviews, 181, 8297.Google Scholar
Beach, T., 1994. The fate of eroded soil – sediment sink and sediment budgets of agrarian landscapes in southern Minnesota, 1851–1988. Annals of the Association of American Geographers, 84(1), 528.Google Scholar
Beaumont, C., Kooi, H., Willet, S., 2000. Coupled tectonic-surface process models with applications to rifted margins and collision orogens. In: Summerfield, M. A. (ed.), Geomorphology and Global Tectonics. Wiley and Sons, Chichester, UK, pp. 2955.Google Scholar
Beauvais, A. A., Montgomery, D. R., 1997. Are channel networks statistically self-similar? Geology, 25(12), 10631066.Google Scholar
Beechie, T., Imaki, H., 2014. Predicting natural channel patterns based on landscape and geomorphic controls in the Columbia River basin, USA. Water Resources Research, 50(1), 3957. 10.1002/2013wr013629.Google Scholar
Beechie, T. J., Liermann, M., Pollock, M. M., Baker, S., Davies, J., 2006. Channel pattern and river-floodplain dynamics in forested mountain river systems. Geomorphology, 78(1–2), 124141.Google Scholar
Beechie, T. J., Sear, D. A., Olden, J. D., et al., 2010. Process-based principles for restoring river ecosystems. Bioscience, 60(3), 209222.Google Scholar
Beechie, T., Richardson, J. S., Gurnell, A. M., Negishi, J., 2013a. Watershed processes, human impacts, and process-based restoration. In: Roni, P., Beechie, T. (eds.), Stream and Watershed Restoration: A Guide to Restoring Riverine Processes and Habitats. Wiley-Blackwell, Chichester, UK, pp. 1149.Google Scholar
Beechie, T., Pess, G.R., Morley, S., et al., 2013b. Watershed assessments and identification of restoration needs. In: Beechie, T., Roni, P. (eds.), Stream and Watershed Restoration: A Guide to Restoring Riverine Processes and Habitats. Wiley-Blackwell, Chichester, UK, pp. 50113.Google Scholar
Beer, T., Borgas, M., 1993. Horton Laws and the fractal nature of streams. Water Resources Research, 29(5), 14751487.Google Scholar
Beeson, C. E., Doyle, P. F., 1995. Comparison of bank erosion at vegetated and non-vegetated channel bends. Water Resources Bulletin, 31(6), 983990.Google Scholar
Begin, Z. B., Meyer, D. F., Schumm, S. A., 1981. Development of longitudinal profiles of alluvial channels in response to base-level lowering. Earth Surface Processes and Landforms, 6(1), 4968.Google Scholar
Belmont, P., Willenbring, J. K., Schottler, S. P., et al., 2014. Toward generalizable sediment fingerprinting with tracers that are conservative and nonconservative over sediment routing timescales. Journal of Soils and Sediments, 14(8), 14791492.CrossRefGoogle Scholar
Benaichouche, A., Stab, O., Tessier, B., Cojan, I., 2016. Evaluation of a landscape evolution model to simulate stream piracies: Insights from multivariable numerical tests using the example of the Meuse basin, France. Geomorphology, 253, 168180.Google Scholar
Benda, L., Poff, N. L., Miller, D., et al., 2004a. The network dynamics hypothesis: How channel networks structure riverine habitats. Bioscience, 54(5), 413427.Google Scholar
Benda, L., Andras, K., Miller, D., Bigelow, P., 2004b. Confluence effects in rivers: Interactions of basin scale, network geometry, and disturbance regimes. Water Resources Research, 40(5). 10.1029/2003wr002583.Google Scholar
Bendixen, M., Best, J. L., Hackney, C., Lonsmann Iverson, L., 2019. Time is running out for sand. Nature, 571, 2931.Google Scholar
Benedetti, M. M., 2003. Controls on overbank deposition in the upper Mississippi River. Geomorphology, 56(3–4), 271290.Google Scholar
Bennett, S. J., 1999. Effect of slope on the growth and migration of headcuts in rills. Geomorphology, 30(3), 273290.Google Scholar
Bennett, S. J., Alonso, C. V., 2005. Kinematics of flow within headcut scour holes on hillslopes. Water Resources Research, 41(9). 10.1029/2004wr003752.Google Scholar
Bennett, S. J., Liu, R. J., 2016. Basin self-similarity, Hack’s law, and the evolution of experimental rill networks. Geology, 44(1), 3538.CrossRefGoogle Scholar
Bennett, S. J., Hou, Y. T., Atkinson, J. F., 2014. Turbulence suppression by suspended sediment within a geophysical flow. Environmental Fluid Mechanics, 14(4), 771794.Google Scholar
Bennett, S. J., Gordon, L. M., Neroni, V., Wells, R. R., 2015. Emergence, persistence, and organization of rill networks on a soil-mantled experimental landscape. Natural Hazards, 79, S7S24.CrossRefGoogle Scholar
Benson, M. A., Thomas, D. M., 1966. A definition of dominant discharge. International Association of Hydrological Sciences Bulletin, 11, 7680.CrossRefGoogle Scholar
Bent, G. C., Waite, A. M., 2013. Equations for estimating bankfull channel geometry and discharge for streams in Massachusetts. U.S. Geological Survey Scientific Investigations Report 2013–5155.Google Scholar
Bentley, P. J., Gulbrandsen, M., Kyvik, S., 2015. The relationship between basic and applied research in universities. Higher Education, 70(4), 689709.Google Scholar
Bergeron, N. E., Abrahams, A. D., 1992. Estimating shear velocity and roughness length from velocity profiles. Water Resources Research, 28(8), 21552158.Google Scholar
Berhanu, M., Petroff, A., Devauchelle, O., Kudrolli, A., Rothman, D. H., 2012. Shape and dynamics of seepage erosion in a horizontal granular bed. Physical Review E, 86(4), 9. 10.1103/PhysRevE.86.041304.Google Scholar
Berlin, M. M., Anderson, R. S., 2007. Modeling of knickpoint retreat on the Roan Plateau, western Colorado. Journal of Geophysical Research – Earth Surface, 112(F3). 10.1029/2006jf000553.Google Scholar
Bernhardt, E. S., Palmer, M. A., Allan, J. D., et al., 2005. Synthesizing US river restoration efforts. Science, 308(5722), 636637.Google Scholar
Bernhardt, E. S., Sudduth, E. B., Palmer, M. A., et al., 2007. Restoring rivers one reach at a time: results from a survey of US river restoration practitioners. Restoration Ecology, 15(3), 482493.Google Scholar
Bertoldi, W., 2012. Life of a bifurcation in a gravel-bed braided river. Earth Surface Processes and Landforms, 37(12), 13271336.Google Scholar
Bertoldi, W., Tubino, M., 2005. Bed and bank evolution of bifurcating channels. Water Resources Research, 41(7). 10.1029/2004wr003333.Google Scholar
Bertoldi, W., Tubino, M., 2007. River bifurcations: experimental observations on equilibrium configurations. Water Resources Research, 43(10). 10.1029/2007wr005907.Google Scholar
Bertoldi, W., Amplatz, T., Miori, T., Zanoni, L., Tubino, M., 2006. Bed load fluctuations and channel processes in a braided network laboratory model. In: Ferreira, R., Alves, C., Leal, G., Cardoso, A. (eds.), River Flow 2006, Vols. 1 & 2. Taylor & Francis, London, pp. 937945.Google Scholar
Bertoldi, W., Zanoni, L., Tubino, M., 2009a. Planform dynamics of braided streams. Earth Surface Processes and Landforms, 34(4), 547557.Google Scholar
Bertoldi, W., Ashmore, P., Tubino, M., 2009b. A method for estimating the mean bed load flux in braided rivers. Geomorphology, 103(3), 330340. 10.1016/j.geomorph.2008.06.014.Google Scholar
Bertoldi, W., Zanoni, L., Miori, S., Repetto, R., Tubino, M., 2009c. Interaction between migrating bars and bifurcations in gravel bed rivers. Water Resources Research, 45. 10.1029/2008wr007086.Google 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.Google Scholar
Bertoldi, W., Gurnell, A. M., Drake, N. A., 2011a. The topographic signature of vegetation development along a braided river: results of a combined analysis of airborne lidar, color air photographs, and ground measurements. Water Resources Research, 47. 10.1029/2010wr010319.Google Scholar
Bertoldi, W., Drake, N. A., Gurnell, A. M., 2011b. Interactions between river flows and colonizing vegetation on a braided river: exploring spatial and temporal dynamics in riparian vegetation cover using satellite data. Earth Surface Processes and Landforms, 36(11), 14741486.Google Scholar
Bertoldi, W., Welber, M., Gurnell, A. M., et al., 2015. Physical modelling of the combined effect of vegetation and wood on river morphology. Geomorphology, 246, 178187.Google Scholar
Beschta, R. L., 1978. Long-term patterns of sediment production following road construction and logging in the Oregon Coast Range. Water Resources Research, 14(6), 10111016.Google Scholar
Beschta, R. L., Pyles, M. R., Skaugset, A. E., Surfleet, C. G., 2000. Peakflow responses to forest practices in the western cascades of Oregon, USA. Journal of Hydrology, 233(1–4), 102120.CrossRefGoogle Scholar
Best, J. L., 1987. Flow dynamics at river channel confluences: Implications for sediment transport and bed morphology. In: Ethridge, F. G., Flores, R. M., Harvey, M. D. (eds.), Recent Developments in Fluvial Sedimentology, Special Publication 39. Society of Economic Paleontologists and Mineralogists, Tulsa, OK, pp. 2735.Google Scholar
Best, J. L., 1988. Sediment transport and bed morphology at river channel confluences. Sedimentology, 35, 481498.Google Scholar
Best, J. L., 1992. On the entrainment of sediment and initiation of bed defects – insights from recent developments within turbulent boundary-layer research. Sedimentology, 39(5), 797811.Google Scholar
Best, J. L., 2005. The fluid dynamics of river dunes: a review and some future research directions. Journal of Geophysical Research – Earth Surface, 110(F4). 10.1029/2004jf000218.Google Scholar
Best, J. L., 2019. Anthropogenic stresses on the world’s big rivers. Nature Geoscience, 12(1), 721.Google Scholar
Best, J. L., Ashworth, P. J., 1997. Scour in large braided rivers, and the recognition of sequence stratigraphic boundaries. Nature, 387(6630), 275277.Google Scholar
Best, J. L., Reid, I., 1984. Separation zone at open-channel junctions. Journal of Hydraulic Engineering, 110(11), 15881594.Google Scholar
Best, J. L., Rhoads, B. L., 2008. Sediment transport, bed morphology, and the sedimentology of river channel confluences. In: Rice, S. P., Roy, A. G., Rhoads, B. L. (eds.), River Confluences, Tributaries and the Fluvial Network. Wiley, Chichester, UK, pp. 4572.Google Scholar
Best, J. L., Roy, A. G., 1991. Mixing layer distortion at the confluence of channels of different depth. Nature, 350(6317), 411413.Google Scholar
Bettis, E. A., Benn, D. W., Hajic, E. R., 2008. Landscape evolution, alluvial architecture, environmental history, and the archaeological record of the Upper Mississippi River Valley. Geomorphology, 101(1–2), 362377.Google Scholar
Bettis, R., White, W. R., 1987. Extremal hypotheses applied to river regime. In: Thorne, C. R., Bathurst, J. C., Hey, R. D. (eds.), Sediment Transport in Gravel-Bed Rivers. Wiley, Chichester, UK, pp. 767789.Google Scholar
Bevan, V., MacVicar, B., Chapuis, M., et al., 2018. Enlargement and evolution of a semi-alluvial creek in response to urbanization. Earth Surface Processes and Landforms, 43(11), 22952312.Google Scholar
Beven, K., 1981. The effect of ordering on the geomorphic effectiveness of hydrologic events. In: Davies, T. R. H., Pearce, A. J. (eds.), Erosion and Sediment Transport in the Pacific Rim Steeplands. IAHS Publication No. 132, IAHS Press, Wallingford, UK, pp. 510526.Google Scholar
Bezak, N., Brilly, M., Sraj, M., 2014. Comparison between the peaks-over-threshold method and the annual maximum method for flood frequency analysis. Hydrological Sciences Journal, 59(5), 959977.Google Scholar
Bhowmik, N. G., Demissie, M., 1982. Bed material sorting in pools and riffles. Journal of the Hydraulics Division – ASCE, 108(10), 12271231.Google Scholar
Biedenharn, D. S., Copeland, R. R., Thorne, C. R., et al., 2000. Effective discharge calculation: A practical guide. U.S. Army Corps of Engineers Research and Development Center, Washington, DC.Google Scholar
Bieger, K., Rathjens, H., Allen, P. M., Arnold, J. G., 2015. Development and evaluation of bankfull hydraulic geometry relationships for the physiographic regions of the United States. Journal of the American Water Resources Association, 51(3), 842858.Google Scholar
Billi, P., Preciso, E., Salemi, E., 2014. Field investigation on step-pool morphology and processes in steep mountain streams. Agriculture and Forestry, 60(3), 728.Google Scholar
Bingner, R. L., Theurer, F. D., Yuan, Y., 2015. AnnAGNPS Technical Processes. National Sedimentation Laboratory, Agricultural Research Service, U.S. Department of Agriculture, Oxford, Mississippi.Google Scholar
Bird, G., Brewer, P. A., Macklin, M. G., et al., 2008. River system recovery following the Novat-Rosu tailings dam failure, Maramures County, Romania. Applied Geochemistry, 23(12), 34983518.Google Scholar
Biron, P., Roy, A. G., Best, J. L., Boyer, C. J., 1993. Bed morphology and sedimentology at the confluence of unequal depth channels. Geomorphology, 8, 115129.Google Scholar
Biron, P., Best, J. L., Roy, A. G., 1996a. Effects of bed discordance on flow dynamics at open channel confluences. Journal of Hydraulic Engineering, 122(12), 676682.Google Scholar
Biron, P., Boy, A. G., Best, J. L., 1996b. Turbulent flow structure at concordant and discordant open-channel confluences. Experiments in Fluids, 21(6), 437446.Google Scholar
Biron, P. M., Richer, A., Kirkbride, A. D., Roy, A. G., Han, S., 2002. Spatial patterns of water surface topography at a river confluence. Earth Surface Processes and Landforms, 27(9), 913928.Google Scholar
Biron, P. M., Robson, C., Lapointe, M. F., Gaskin, S. J., 2004a. Comparing different methods of bed shear stress estimates in simple and complex flow fields. Earth Surface Processes and Landforms, 29(11), 14031415.CrossRefGoogle Scholar
Biron, P. M., Ramamurthy, A. S., Han, S., 2004b. Three-dimensional numerical modeling of mixing at river confluences. Journal of Hydraulic Engineering, 130(3), 243253.Google Scholar
Biron, P. M., Buffin-Belanger, T., Larocque, M., et al., 2014. Freedom space for rivers: A sustainable management approach to enhance river resilience. Environmental Management, 54(5), 10561073.Google Scholar
Biron, P. M., Buffin-Belanger, T., Masse, S., 2018. The need for river management and stream restoration practices to integrate hydrogeomorphology. Canadian Geographer, 62(2), 288295.Google Scholar
Bishop, P., 2007. Long-term landscape evolution: Linking tectonics and surface processes. Earth Surface Processes and Landforms, 32(3), 329365.Google Scholar
Bishop, P., Hoey, T. B., Jansen, J. D., Artza, I. L., 2005. Knickpoint recession rate and catchment area: The case of uplifted rivers in Eastern Scotland. Earth Surface Processes and Landforms, 30(6), 767778.Google Scholar
Blainey, J. B., Webb, R. H., Moss, M. E., Baker, V. R., 2002. Bias and information content of paleoflood data in flood-frequency analysis. In: House, P. K., Webb, R. H., Baker, V. R., Levish, D. R. (eds.), Ancient Floods, Modern Hazards: Principles and Applications of Paleoflood Hydrology. American Geophysical Union, Washington, DC, pp. 161174.Google Scholar
Blanckaert, K., 2009. Saturation of curvature-induced secondary flow, energy losses, and turbulence in sharp open-channel bends: Laboratory experiments, analysis, and modeling. Journal of Geophysical Research – Earth Surface, 114. 10.1029/2008jf001137.Google Scholar
Blanckaert, K., 2010. Topographic steering, flow recirculation, velocity redistribution, and bed topography in sharp meander bends. Water Resources Research, 46. 10.1029/2009wr008303.Google Scholar
Blanckaert, K., 2011. Hydrodynamic processes in sharp meander bends and their morphological implications. Journal of Geophysical Research – Earth Surface, 116. 10.1029/2010jf001806.Google Scholar
Blanckaert, K., 2015. Flow separation at convex banks in open channels. Journal of Fluid Mechanics, 779, 432467.Google Scholar
Blanckaert, K., de Vriend, H. J., 2004. Secondary flow in sharp open-channel bends. Journal of Fluid Mechanics, 498, 353380.Google Scholar
Blanckaert, K., de Vriend, H. J., 2010. Meander dynamics: a nonlinear model without curvature restrictions for flow in open-channel bends. Journal of Geophysical Research – Earth Surface, 115. 10.1029/2009jf001301.Google Scholar
Blanckaert, K., Graf, W. H., 2004. Momentum transport in sharp open-channel bends. Journal of Hydraulic Engineering, 130(3), 186198.Google Scholar
Blanckaert, K., Duarte, A., Chen, Q., Schleiss, A. J., 2012. Flow processes near smooth and rough (concave) outer banks in curved open channels. Journal of Geophysical Research – Earth Surface, 117. 10.1029/2012jf002414.Google Scholar
Blanckaert, K., Duarte, A., Schleiss, A. J., 2010. Influence of shallowness, bank inclination and bank roughness on the variability of flow patterns and boundary shear stress due to secondary currents in straight open-channels. Advances in Water Resources, 33(9), 10621074. 10.1016/j.advwatres.2010.06.012.Google Scholar
Blanckaert, K., Kleinhans, M. G., McLelland, S. J., et al., 2013. Flow separation at the inner (convex) and outer (concave) banks of constant-width and widening open-channel bends. Earth Surface Processes and Landforms, 38(7), 696716.CrossRefGoogle Scholar
Bledsoe, B. P., Watson, C. C., 2001a. Logistic analysis of channel pattern thresholds: Meandering, braiding, and incising. Geomorphology, 38(3–4), 281300.Google Scholar
Bledsoe, B. P., Watson, C. C., 2001b. Effects of urbanization on channel instability. Journal of the American Water Resources Association, 37(2), 255270.Google Scholar
Bledsoe, B. P., Watson, C. C., Biedenharn, D. S., 2002. Quantification of incised channel evolution and equilibrium. Journal of the American Water Resources Association, 38(3), 861870.Google Scholar
Blettler, M. C. M., Amsler, M. L., Ezcurra de Drago, I., et al., 2015. The impact of significant input of fine sediment on benthic fauna at tributary junctions: A case study of the Bermejo-Paraguay River confluence, Argentina. Ecohydrology, 8(2), 340352.Google Scholar
Blom, A., Viparelli, E., Chavarrias, V. C., 2016. The graded alluvial river: Profile concavity and downstream fining. Geophysical Research Letters, 43(12), 62856293. 10.1002/2016gl068898.Google Scholar
Blom, A., Arkesteijn, L., Chavarrias, V., Viparelli, E., 2017a. The equilibrium alluvial river under variable flow and its channel-forming discharge. Journal of Geophysical Research – Earth Surface, 122(10), 19241948.Google Scholar
Blom, A., Chavarrias, V., Ferguson, R. I., Viparelli, E., 2017b. Advance, retreat, and halt of abrupt gravel-sand transitions in alluvial rivers. Geophysical Research Letters, 44(19), 97519760. 10.1002/2017gl074231.Google Scholar
Blondeaux, P., Seminara, G., 1985. A unified bar bend theory of river meanders. Journal of Fluid Mechanics, 157(AUG), 449470.Google Scholar
Bloschl, G., Gaal, L., Hall, J., et al., 2015. Increasing river floods: fiction or reality? Water, 2(4), 329344.Google Scholar
Blott, S. J., Pye, K., 2001. GRADISTAT: A grain size distribution and statistics package for the analysis of unconsolidated sediments. Earth Surface Processes and Landforms, 26(11), 12371248.Google Scholar
Blott, S. J., Pye, K., 2012. Particle size scales and classification of sediment types based on particle size distributions: Review and recommended procedures. Sedimentology, 59(7), 20712096.Google Scholar
Bluck, B. J., 1987. Bed forms and clast size changes in gravel-bed rivers. In: Richards, K. S. (ed.), River Channels: Environment and Process. Blackwell, Oxford, UK, pp. 159178.Google Scholar
Blue, B., 2018. What’s wrong with healthy rivers? Promise and practice in the search for a guiding ideal for freshwater management. Progress in Physical Geography, 42(4), 462477.Google Scholar
Blue, B., Brierley, G., 2016. “But what do you measure?” Prospects for a constructive critical physical geography. Area, 48(2), 190197.CrossRefGoogle Scholar
Blum, M. D., Guccione, M. J., Wysocki, D. A., Robnett, P. C., Rutledge, E. M., 2000. Late Pleistocene evolution of the lower Mississippi River valley, southern Missouri to Arkansas. Geological Society of America Bulletin, 112(2), 221235.Google Scholar
Bolla Pittaluga, M., Repetto, R., Tubino, M., 2003. Channel bifurcation in braided rivers: Equilibrium configurations and stability. Water Resources Research, 39(3). 10.1029/2001wr001112.Google Scholar
Bolla Pittaluga, M., Coco, G., Kleinhans, M. G., 2015. A unified framework for stability of channel bifurcations in gravel and sand fluvial systems. Geophysical Research Letters, 42(18), 75217536.Google Scholar
Bomhof, J., Rennie, C. D., Jenkinson, R. W., 2015. Use of local soil and vegetation classifications to improve regional downstream hydraulic geometry relations. Journal of Hydraulic Engineering, 141(5). 10.1061/(asce)hy.1943–7900.0000978.Google Scholar
Booker, D. J., Sear, D. A., Payne, A. J., 2001. Modelling three-dimensional flow structures and patterns of boundary shear stress in a natural pool-riffle sequence. Earth Surface Processes and Landforms, 26(5), 553576.Google Scholar
Booth, D. B., Fischenich, C. J., 2015. A channel evolution model to guide sustainable urban stream restoration. Area, 47(4), 408421.Google Scholar
Booth, D. B., Roy, A. H., Smith, B., Capps, K. A., 2016. Global perspectives on the urban stream syndrome. Freshwater Science, 35(1), 412420.Google Scholar
Borrelli, P., Robinson, D. A., Fleischer, L. R., et al., 2017. An assessment of the global impact of 21st century land use change on soil erosion. Nature Communications, 8. 10.1038/s41467-017–02142-7.Google Scholar
Bouchez, J., Lajeunesse, E., Gaillardet, J., et al., 2010. Turbulent mixing in the Amazon River: The isotopic memory of confluences. Earth and Planetary Science Letters, 290(1–2), 3743.Google Scholar
Boyer, C., Roy, A. G., Best, J. L., 2006. Dynamics of a river channel confluence with discordant beds: Flow turbulence, bed load sediment transport, and bed morphology. Journal of Geophysical Research – Earth Surface, 111(F4). 10.1029/2005JF000458.Google Scholar
Bracken, L. J., Wainwright, J., 2006. Geomorphological equilibrium: myth and metaphor? Transactions of the Institute of British Geographers, 31(2), 167178.Google Scholar
Bradbrook, K. F., Lane, S. N., Richards, K. S., 2000. Numerical simulation of three-dimensional, time-averaged flow structure at river confluences. Water Resources Research, 36 (9), 27312746.Google Scholar
Bradbrook, K. F., Lane, S. N., Richards, K. S., Biron, P. M., Roy, A. G., 2001. Role of bed discordance at asymmetrical river confluences. Journal of Hydraulic Engineering, 127(5), 351368.Google Scholar
Brandt, S. A., 2000. Classification of geomorphological effects downstream of dams. Catena, 40(4), 375401.Google Scholar
Brasington, J., Rumsby, B. T., McVey, R. A., 2000. Monitoring and modelling morphological change in a braided gravel-bed river using high resolution GPS-based survey. Earth Surface Processes and Landforms, 25(9), 973990.Google Scholar
Brasington, J., Langham, J., Rumsby, B., 2003. Methodological sensitivity of morphometric estimates of coarse fluvial sediment transport. Geomorphology, 53(3–4), 299316.Google Scholar
Braudrick, C. A., Dietrich, W. E., Leverich, G. T., Sklar, L. S., 2009. Experimental evidence for the conditions necessary to sustain meandering in coarse-bedded rivers. Proceedings of the National Academy of Sciences, of the United States of America 106(40), 1693616941.Google Scholar
Braun, J., Sambridge, M., 1997. Modelling landscape evolution on geological time scales: A new method based on irregular spatial discretization. Basin Research, 9(1), 2752.Google Scholar
Bray, D. I., 1982. Regime equations for gravel-bed rivers. In: Hey, R. D., Bathurst, J. C., Thorne, C. R. (eds.), Gravel-Bed Rivers. Wiley, Chichester, UK, pp. 517542.Google Scholar
Brayshaw, A. C., 1984. Characteristics and origin of cluster bedforms in coarse-grained alluvial channels. In: Koster, E. H., Steele, R. J. (eds.), Sedimentology of Gravels and Conglomerates. Canadian Society of Petroleum Geologists Memoir 10, pp. 7785.Google Scholar
Brebner, A., Wilson, K. C., 1967. Derivation of regime equations from relationships for pressurized flow by use of minimum energy degradation rate. Proceedings of the Institution of Civil Engineers, 36(JAN), 4762.Google Scholar
Brenna, A., Surian, N., Mao, L., 2019. Virtual velocity approach for estimating bed material transport in gravel-bed rivers: Key factors and significance. Water Resources Research, 55(2), 16511674. 10.1029/2018wr023556.Google Scholar
Brennan, S. R., Schindler, D. E., Cline, T. J., et al., 2019. Shifting habitat mosaics and fish production across river basins. Science, 364(6442), 783786.Google Scholar
Bressan, F., Papanicolaou, A. N., Abban, B., 2014. A model for knickpoint migration in first- and second-order streams. Geophysical Research Letters, 41(14), 49874996. 10.1002/2014gl060823.CrossRefGoogle Scholar
Brewer, P. A., Lewin, J., 1998. Planform cyclicity in an unstable reach: Complex fluvial response to environmental change. Earth Surface Processes and Landforms, 23(11), 9891008.Google Scholar
Brewer, P. A., Leeks, G. J. L., Lewin, J., 1992. Direct measurement of in-channel abrasion processes. In: Bogen, J., Walling, D. E., Day, T. J. (eds.), Erosion and Sediment Transport Monitoring Programmes in River Basins. IAHS Publication No. 210, IAHS Press, Wallingford, UK, pp. 2129.Google Scholar
Bridge, J. S., 1985. Paleochannel patterns inferred from alluvial deposits – a critical evaluation. Journal of Sedimentary Petrology, 55(4), 579589.Google Scholar
Bridge, J. S., 1993. The interaction between channel geometry, water flow, sediment transport and deposition in braided rivers. In: Best, J. L., Bristow, C. S. (eds.), Braided Rivers. Special Publication No. 75. Geological Society of London, London, pp. 1371.Google Scholar
Bridge, J. S., 2003. Rivers and Floodplains: Form, Process and Sedimentary Record. Blackwell Publishing, Oxford, UK.Google Scholar
Bridge, J. S., Bennett, S. J., 1992. A model for the entrainment and transport of sediment grains of mixed sizes, shapes, and densities. Water Resources Research, 28(2), 337363.Google Scholar
Bridge, J. S., Gabel, S. L., 1992. Flow and sediment dynamics in a low sinuosity braided river – Calamus River, Nebraska Sandhills. Sedimentology, 39(1), 125142.Google Scholar
Bridge, J. S., Lunt, I. A., 2006. Depositional models of braided rivers. In: Smith, G. H. S., Best, J. L., Bristow, C. S., Petts, G. E. (eds.), Braided Rivers: Process, Deposits, Ecology and Management. Special Publications of the International Association of Sedimentologists, 36, Blackwell, Malden, MA, pp. 1150.Google Scholar
Bridge, J. S., Smith, N. D., Trent, F., Gabel, S. L., Bernstein, P., 1986. Sedimentology and morphology of a low-sinuosity river – Calamus River, Nebraska Sand Hills. Sedimentology, 33(6), 851870.Google Scholar
Bridge, J. S., Alexander, J., Collier, R. E. L., Gawthorpe, R. L., Jarvis, J., 1995. Ground-penetrating radar and coring used to study the large-scale structure of point-bar deposits in 3 dimensions. Sedimentology, 42(6), 839852.Google Scholar
Brierley, G. J., 1989. River planform facies models – the sedimentology of braided, wandering and meandering reaches of the Squamish River, British Columbia. Sedimentary Geology, 61(1–2), 1735.Google Scholar
Brierley, G. J., 1991a. Bar sedimentology of the Squamish River, British Columbia – definition and application of morphostratigraphic units. Journal of Sedimentary Petrology, 61(2), 211225.Google Scholar
Brierley, G. J., 1991b. Floodplain sedimentology of the Squamish River, British Columbia – relevance of element analysis. Sedimentology, 38(4), 735750.Google Scholar
Brierley, G. J., 2010. Landscape memory: The imprint of the past on contemporary landscape forms and processes. Area, 42(1), 7685.Google Scholar
Brierley, G. J., Hickin, E. J., 1985. The downstream gradation of particle size in the Squamish River, British Columbia. Earth Surface Processes and Landforms, 10(6), 597606.Google Scholar
Brierley, G. J., Hickin, E. J., 1991. Channel planform as a non-controlling factor in fluvial sedimentology – the case of the Squamish River floodplain, British Columbia. Sedimentary Geology, 75(1–2), 6783.Google Scholar
Brierley, G. J., Hickin, E. J., 1992. Floodplain development based on selective preservation of sediments, Squamish River, British Columbia. Geomorphology, 4(6), 381391.Google Scholar
Brierley, G., Fryirs, K., 1998. A fluvial sediment budget for upper Wolumla Creek, south coast, New South Wales, Australia. Australian Geographer, 29(2), U1-111.Google Scholar
Brierley, G., Fryirs, K., 2005. Geomorphology and River Management. Blackwell, Malden, MA.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.Google Scholar
Brierley, G., Fryirs, K., 2012. Geomorphic Analysis of Rivers: An Approach to Reading the Landscape. Wiley, Chichester, UK.Google Scholar
Brierley, G. J., Fryirs, K. A., 2016. The use of evolutionary trajectories to guide “moving targets” in the management of river futures. River Research and Applications, 32(5), 823835.Google Scholar
Brierley, G. J., Ferguson, R. J., Woolfe, K. J., 1997. What is a fluvial levee? Sedimentary Geology, 114(1–4), 19.Google Scholar
Bristow, C. S., Skelly, R. L., Ethridge, F. G., 1999. Crevasse splays from the rapidly aggrading, sand-bed, braided Niobrara River, Nebraska: Effect of base-level rise. Sedimentology, 46(6), 10291047.Google Scholar
Brizga, S. O., Finlayson, B. L., 1990. Channel avulsion and river metamorphosis – the case of the Thomson River, Victoria, Australia. Earth Surface Processes and Landforms, 15(5), 391404.Google Scholar
Brookes, A., 1985. River channelization: Traditional engineering methods, physical consequences and alternative practices. Progress in Physical Geography, 9(1), 4473.Google Scholar
Brookes, A., 1987a. The distribution and management of channelized streams in Denmark. Regulated Rivers, 1, 316.Google Scholar
Brookes, A., 1987b. River channel adjustments downstream from channelization works in England and Wales. Earth Surface Processes and Landforms, 12(4), 337351.Google Scholar
Brookes, A., 1987c. Restoring the sinuosity of artificially straightened stream channels. Environmental Geology and Water Sciences, 10(1), 3341.Google Scholar
Brookes, A., 1988. Channelized Rivers: Perspectives for Environmental Management. Wiley, Chichester, UK.Google Scholar
Brookes, A., 1995. Challenges and objectives for geomorphology in UK river management. Earth Surface Processes and Landforms, 20(7), 593610.Google Scholar
Brooks, A. P., Brierley, G. J., 1997. Geomorphic responses of lower Bega River to catchment disturbance, 1851–1926. Geomorphology, 18(3–4), 291304.Google Scholar
Brooks, A. P., Brierley, G. J., Millar, R. G., 2003. The long-term control of vegetation and woody debris on channel and flood-plain evolution: Insights from a paired catchment study in southeastern Australia. Geomorphology, 51(1–3), 729.Google Scholar
Browand, F. K., 1986. The structure of the turbulent mixing layer. Physica D, 18(1–3), 135148.Google Scholar
Brown, A. G., Carey, C., Erkens, G., et al., 2009. From sedimentary records to sediment budgets: Multiple approaches to catchment sediment flux. Geomorphology, 108(1–2), 3547.Google Scholar
Brown, A. G., Tooth, S., Chiverrell, R. C., et al., 2013. The Anthropocene: Is there a geomorphological case? Earth Surface Processes and Landforms, 38(4), 431434.Google Scholar
Brown, A. G., Petit, F., James, L. A., 2016. Archeology and human artefacts. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology, 2nd Edition. Wiley, Chichester, UK, pp. 4055.Google Scholar
Brown, A. G., Tooth, S., Bullard, J. E., et al., 2017. The geomorphology of the Anthropocene: Emergence, status and implications. Earth Surface Processes and Landforms, 42(1), 7190.Google Scholar
Brune, G. M., 1953. Trap efficiency of reservoirs. Transactions, American Geophysical Union, 34(3), 407418.Google Scholar
Brunsden, D., 2001. A critical assessment of the sensitivity concept in geomorphology. Catena, 42(2–4), 99123.Google Scholar
Brunsden, D., Thornes, J. B., 1979. Landscape sensitivity and change. Transactions of the Institute of British Geographers, 4(4), 463484.Google Scholar
Brush, L. M. Jr., 1961. Drainage basins, channels, and flow characteristics of selected streams in central Pennsylvania U.S. Geological Survey Professional Paper 282-F. U.S. Government Printing Office, Washington, DC.Google Scholar
Bryan, R. B., Jones, J. A. A., 1997. The significance of soil piping processes: inventory and prospect. Geomorphology, 20(3–4), 209218.Google Scholar
Buffin-Belanger, T., Roy, A. G., 1998. Effects of a pebble cluster on the turbulent structure of a depth-limited flow in a gravel-bed river. Geomorphology, 25(3–4), 249267.Google Scholar
Buffin-Belanger, T., Roy, A. G., Kirkbride, A. D., 2000. On large-scale flow structures in a gravel-bed river. Geomorphology, 32(3–4), 417435.Google Scholar
Buffin-Belanger, T., Biron, P. M., Larocque, M., et al., 2015. Freedom space for rivers: An economically viable river management concept in a changing climate. Geomorphology, 251, 137148.Google Scholar
Buffington, J. M., 1999. The legend of A. F. Shields. Journal of Hydraulic Engineering, 125(4), 376387.Google Scholar
Buffington, J. M., Montgomery, D. R., 1997. A systematic analysis of eight decades of incipient motion studies, with special reference to gravel-bedded rivers. Water Resources Research, 33(8), 19932029.Google Scholar
Buffington, J. M., Montgomery, D. R., 2013. Geomorphic classification of rivers. In: Shroder, J. W. (ed.), Treatise on Geomorphology, Vol. 9, Fluvial geomorphology, Wohl, E. (vol. ed.). Academic Press, San Diego, CA, pp. 730767.Google Scholar
Buffington, J. M., Dietrich, W. E., Kirchner, J. W., 1992. Friction angle measurements on a naturally formed gravel streambed – implications for critical boundary shear stress. Water Resources Research, 28(2), 411425.Google Scholar
Buffington, J. M., Lisle, T. E., Woodsmith, R. D., Hilton, S., 2002. Controls on the size and occurrence of pools in coarse-grained forest rivers. River Research and Applications, 18(6), 507531.Google Scholar
Bull, L. J., 1997. Relative velocities of discharge and sediment waves for the River Severn, UK. Hydrological Sciences Journal, 42(5), 649660.Google Scholar
Bull, L. J., Kirkby, M. J., 1997. Gully processes and modelling. Progress in Physical Geography, 21(3), 354374.CrossRefGoogle Scholar
Bull, L. J., Lawler, D. M., Leeks, G. J. L., Marks, S., 1995. Downstream changes in suspended sediment fluxes in the River Severn, UK. In: Osterkamp, W. R. (ed.), Effects of Scale on Interpretation and Management of Sediment and Water Quality. IAHS Publication 226, IAHS Press, Wallingford, UK, pp. 2737.Google Scholar
Bull, W. B., Scott, K. M., 1974. Impact of mining gravel from urban streambeds in the southwestern United States. Geology, 2, 171174.Google Scholar
Bunte, K., Abt, S. R., 2001. Sampling Surface and Subsurface Particle-Size Distributions in Wadable Gravel- and Cobble-Bed Streams for Analyses in Sediment Transport, Hydraulics, and Streambed Monitoring. General Technical Report RMRS-GTR-74, Rocky Mountain Research Station, Forest Service, U.S. Department of Agriculture, Fort Collins, CO.Google Scholar
Bunte, K., Abt, S. R., Swingle, K. W., Cenderelli, D. A., 2014. Effective discharge in Rocky Mountain headwater streams. Journal of Hydrology, 519, 21362147.Google Scholar
Buraas, E. M., Renshaw, C. E., Magilligan, F. J., Dade, W. B., 2014. Impact of reach geometry on stream channel sensitivity to extreme floods. Earth Surface Processes and Landforms, 39(13), 17781789.Google Scholar
Burckhardt, J. C., Todd, B. L., 1998. Riparian forest effect on lateral stream channel migration in the glacial till plains. Journal of the American Water Resources Association, 34(1), 179184.Google Scholar
Burge, L. M., 2005. Wandering Miramichi rivers, New Brunswick, Canada. Geomorphology, 69(1–4), 253274. 10.1016/j.geomorph.2005.01.010.Google Scholar
Burge, L. M., 2006. Stability, morphology and surface grain size patterns of channel bifurcation in gravel-cobble bedded anabranching rivers. Earth Surface Processes and Landforms, 31(10), 12111226.Google Scholar
Burge, L. M., Lapointe, M. F., 2005. Understanding the temporal dynamics of the wandering Renous River, New Brunswick, Canada. Earth Surface Processes and Landforms, 30(10), 12271250.Google Scholar
Burkham, D. E., 1972. Channel changes of the Gila River in the Safford Valley, Arizona 1846–1970. U.S. Geological Survey Professional Paper 655-G, U.S. Government Printing Office Washington, DC.Google Scholar
Buscombe, D., Conley, D. C., 2012. Effective shear stress of graded sediments. Water Resources Research, 48. 10.1029/2010wr010341.Google Scholar
Bussi, G., Dadson, S. J., Bowes, M. J., Whitehead, P. G., 2017. Seasonal and interannual changes in sediment transport identified through sediment rating curves. Journal of Hydrologic Engineering, 22(2). 10.1061/(asce)he.1943–5584.0001466.Google Scholar
Caamano, D., Goodwin, P., Buffington, J. M., Liou, J. C. P., Daley-Laursen, S., 2009. Unifying criterion for the velocity reversal hypothesis in gravel-bed rivers. Journal of Hydraulic Engineering, 135(1), 6670.Google Scholar
Caamano, D., Goodwin, P., Buffington, J. M., 2012. Flow structure through pool-riffle sequences and a conceptual model for their sustainability in gravel-bed rivers. River Research and Applications, 28(3), 377389.Google Scholar
Cabezas, A., Angulo-Martinez, M., Gonzalez-Sanchis, M., Jimenez, J. J., Comin, F. A., 2010. Spatial variability in floodplain sedimentation: The use of generalized linear mixed-effects models. Hydrology and Earth System Sciences, 14(8), 16551668.Google 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.Google Scholar
Call, B. C., Belmont, P., Schmidt, J. C., Wilcock, P. R., 2017. Changes in floodplain inundation under nonstationary hydrology for an adjustable, alluvial river channel. Water Resources Research, 53(5), 38113834.Google Scholar
Calle, M., Alho, P., Benito, G., 2017. Channel dynamics and geomorphic resilience in an ephemeral Mediterranean river affected by gravel mining. Geomorphology, 285, 333346.Google Scholar
Calver, A., Anderson, M. G., 2004. Conceptual framework for the persistence of flood-initiated geomorphological features. Transactions of the Institute of British Geographers, 29(1), 129137.Google Scholar
Camacho, R., Yen, B. C., 1991. Nonlinear resistance relationships for alluvial channels. In: Yen, B. C. (ed.), Channel Flow Resistance: Centennial of Manning’s Formula. Water Resources Publications, Highlands, Colorado, pp. 186194.Google Scholar
Campbell, A. J., Sidle, R. C., 1985. Bedload transport in a pool-riffle sequence of a coastal Alaska stream. Water Resources Bulletin, 21(4), 579590.Google Scholar
Campodonico, V. A., Garcia, M. G., Pasquini, A. I., 2015. The dissolved chemical and isotopic signature downflow the confluence of two large rivers: The case of the Parana and Paraguay rivers. Journal of Hydrology, 528, 161176.Google Scholar
Camporeale, C., Perona, P., Porporato, A., Ridolfi, L., 2007. Hierarchy of models for meandering rivers and related morphodynamic processes. Reviews of Geophysics, 45(1). 10.1029/2005rg000185.Google Scholar
Camporeale, C., Perucca, E., Ridolfi, L., 2008. Significance of cutoff in meandering river dynamics. Journal of Geophysical Research – Earth Surface, 113(F1). 10.1029/2006jf000694.Google Scholar
Canovaro, F., Solari, L., 2007. Dissipative analogies between a schematic macro-roughness arrangement and step-pool morphology. Earth Surface Processes and Landforms, 32(11), 16281640.Google Scholar
Cant, D. J., Walker, R. G., 1978. Fluvial processes and facies sequences in sandy braided south Saskatchewan River, Canada. Sedimentology, 25(5), 625648.Google Scholar
Cantelli, A., Muto, T., 2014. Multiple knickpoints in an alluvial river generated by a single instantaneous drop in base level: experimental investigation. Earth Surface Dynamics, 2(1), 271278.Google Scholar
Cantelli, A., Wong, M., Parker, G., Paola, C., 2007. Numerical model linking bed and bank evolution of incisional channel created by dam removal. Water Resources Research, 43(7). 10.1029/2006wr005621.Google Scholar
Cao, S. Y., Knight, D. W., 1997. Entropy-based design approach of threshold alluvial channels. Journal of Hydraulic Research, 35(4), 505524.Google Scholar
Cao, S. Y., Knight, D. W., 1998. Design for hydraulic geometry of alluvial channels. Journal of Hydraulic Engineering, 124(5), 484492.Google Scholar
Cao, Z. X., Egashira, S. J., Carling, P. A., 2003. Role of suspended-sediment particle size in modifying velocity profiles in open channel flows. Water Resources Research, 39(2). 10.1029/2001wr000934.Google Scholar
Capra, A., Di Stefano, C., Ferro, V., Scicolone, B., 2009. Similarity between morphological characteristics of rills and ephemeral gullies in Sicily, Italy. Hydrological Processes, 23(23), 33343341.Google Scholar
Carling, P. A., 1983. Threshold of coarse sediment transport in broad and narrow natural streams. Earth Surface Processes and Landforms, 8(1), 118.Google Scholar
Carling, P. A., 1984. Deposition of fine and coarse sand in an open-work gravel bed. Canadian Journal of Fisheries and Aquatic Sciences, 41(2), 263280.Google Scholar
Carling, P., 1988. The concept of dominant discharge applied to two gravel-bed streams in relation to channel stability thresholds. Earth Surface Processes and Landforms, 13(4), 355367.Google Scholar
Carling, P. A., 1991. An appraisal of the velocity-reversal hypothesis for stable pool riffle sequences in the River Severn, England. Earth Surface Processes and Landforms, 16(1), 1931.Google Scholar
Carling, P. A., 1999. Subaqueous gravel dunes. Journal of Sedimentary Research, 69(3), 534545.Google Scholar
Carling, P. A., Orr, H. G., 2000. Morphology of riffle-pool sequences in the River Severn, England. Earth Surface Processes and Landforms, 25(4), 369384.Google Scholar
Carling, P. A., Reader, N. A., 1982. Structure, composition and bulk properties of upland stream gravels. Earth Surface Processes and Landforms, 7(4), 349365.Google Scholar
Carling, P. A., Wood, N., 1994. Simulation of flow over pool-riffle topography – a consideration of the velocity reversal hypothesis. Earth Surface Processes and Landforms, 19(4), 319332.Google Scholar
Carling, P. A., Kelsey, A., Glaister, M. S. 1992. Effect of bed roughness, particle shape and orientation on initial motion criteria. In: Billi, P., Hey, R. D., Thorne, C. R., Tacconi, P. (eds.), Dynamics of Gravel-Bed Rivers. Wiley, Chichester, UK, pp. 2338.Google Scholar
Carling, P. A., Cao, Z. X., Holland, M. J., Ervine, D. A., Babaeyan-Koopaei, K., 2002. Turbulent flow across a natural compound channel. Water Resources Research, 38(12). 10.1029/2001wr000902.Google Scholar
Carling, P., Jansen, J., Meshkova, L., 2014. Multichannel rivers: Their definition and classification. Earth Surface Processes and Landforms, 39(1), 2637.Google Scholar
Carling, P. A., Gupta, N., Atkinson, P. M., He, H. Q., 2016. Criticality in the planform behavior of the Ganges River meanders. Geology, 44(10), 859862.Google Scholar
Carling, P. A., Perillo, M., Best, J., Garcia, M. H., 2017. The bubble bursts for cavitation in natural rivers: Laboratory experiments reveal minor role in bedrock erosion. Earth Surface Processes and Landforms, 42(9), 13081316.Google Scholar
Carlston, C. W., 1969. Downstream variations in hydraulic geometry of streams – special emphasis on mean velocity. American Journal of Science, 267(4), 499509.Google Scholar
Carretier, S., Poisson, B., Vassallo, R., Pepin, E., Farias, M., 2009. Tectonic interpretation of transient stage erosion rates at different spatial scales in an uplifting block. Journal of Geophysical Research – Earth Surface, 114, 19. 10.1029/2008jf001080.Google Scholar
Carson, M. A., 1984a. The meandering-braided river threshold – a reappraisal. Journal of Hydrology, 73(3–4), 315334. 10.1016/0022–1694(84)90006–4.Google Scholar
Carson, M. A., 1984b. Observations on the meandering-braided transition, the Canterbury Plains, New Zealand. Part One. New Zealand Geographer, 40, 1217.Google Scholar
Carson, M. A., 1986. Characteristics of high-energy “meandering” rivers – the Canterbury Plains, New Zealand. Geological Society of America Bulletin, 97(7), 886895.Google Scholar
Carson, M. A., Griffiths, G. A., 1989. Gravel transport in the braided Waimakariri River – mechanisms, measurements, and predictions. Journal of Hydrology, 109(3–4), 201220.Google Scholar
Carson, M. A., Lapointe, M. F., 1983. The inherent asymmetry of river meander planform. Journal of Geology, 91(1), 4155.Google Scholar
Cartigny, M. J. B., Ventra, D., Postma, G., van den Berg, J. H., 2014. Morphodynamics and sedimentary structures of bedforms under supercritical-flow conditions: new insights from flume experiments. Sedimentology, 61(3), 712748. 10.1111/sed.12076.Google Scholar
Castelltort, S., Simpson, G., 2006. River spacing and drainage network growth in widening mountain ranges. Basin Research, 18(3), 267276.Google Scholar
Castillo, C., Gomez, J. A., 2016. A century of gully erosion research: Urgency, complexity and study approaches. Earth-Science Reviews, 160, 300319.Google Scholar
Castillo, M., Bishop, P., Jansen, J. D., 2013. Knickpoint retreat and transient bedrock channel morphology triggered by base-level fall in small bedrock river catchments: The case of the Isle of Jura, Scotland. Geomorphology, 180, 19.Google Scholar
Castro, J. M., Jackson, P. L., 2001. Bankfull discharge recurrence intervals and regional hydraulic geometry relationships: Patterns in the Pacific Northwest, USA. Journal of the American Water Resources Association, 37(5), 12491262.Google 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.Google Scholar
Celik, A. O., Diplas, P., Dancey, C. L., 2013. Instantaneous turbulent forces and impulse on a rough bed: Implications for initiation of bed material movement. Water Resources Research, 49(4), 22132227. 10.1002/wrcr.20210.Google Scholar
Chang, H. H., 1979. Minimum stream power and river channel patterns. Journal of Hydrology, 41, 303327.Google Scholar
Chang, H. H., 1980. Geometry of gravel streams. Journal of the Hydraulics Division – ASCE, 106(9), 14431456.Google Scholar
Chang, H. H., 1988a. Fluvial Processes in River Engineering. Wiley, New York.Google Scholar
Chang, H. H., 1988b. On the cause of river meandering. In: White, W. R. (ed.), International Conference on River Regime. Wiley, New York, pp. 8393.Google Scholar
Chang, H. Y., Simons, D. B., Woolhiser, D. A., 1971. Flume experiments on alternate bar formation. Journal of the Waterways, Harbors and Coastal Engineering Division, ASCE, 97, 155165.Google Scholar
Chartrand, S. M., Whiting, P. J., 2000. Alluvial architecture in headwater streams with special emphasis on step-pool topography. Earth Surface Processes and Landforms, 25(6), 583600.Google Scholar
Chartrand, S. M., Jellinek, M., Whiting, P. J., Stamm, J., 2011. Geometric scaling of step-pools in mountain streams: observations and implications. Geomorphology, 129(1–2), 141151. 10.1016/j.geomorph.2011.01.020.Google Scholar
Chartrand, S. M., Jellinek, A. M., Hassan, M. A., Ferrer-Boix, C., 2018. Morphodynamics of a width-variable gravel bed stream: new insights on pool-riffle formation from physical experiments. Journal of Geophysical Research – Earth Surface, 123(11), 27352766. 10.1029/2017jf004533.Google Scholar
Chen, A., Darbon, J., Morel, J.-M., 2014. Landscape evolution models: A review of their fundamental equations. Geomorphology, 219, 6886. 10.1016/j.geomorph.2014.04.037.Google Scholar
Chen, D., Jirka, G. H., 1995. Experimental study of plane turbulent wakes in a shallow water layer. Fluid Dynamics Research, 16, 1141.Google Scholar
Chen, X., Zhu, D. Z., Steffler, P. M., 2017. Secondary currents induced mixing at channel confluences. Canadian Journal of Civil Engineering, 44(12), 10711083. 10.1139/cjce-2016–0228.Google Scholar
Cheng, C., Song, Z. Y., Wang, Y. G., Zhang, J. S., 2013. Parameterized expressions for an improved Rouse equation. International Journal of Sediment Research, 28(4), 523534.Google Scholar
Cheng, L., Yaeger, M., Viglione, A., et al., 2012. Exploring the physical controls of regional patterns of flow duration curves – Part 1: insights from statistical analyses. Hydrology and Earth System Sciences, 16(11), 44354446.Google Scholar
Cheng, N. S., 1997. Simplified settling velocity formula for sediment particle. Journal of Hydraulic Engineering, 123(2), 149152.Google Scholar
Chew, L. C., Ashmore, P. E., 2001. Channel adjustment and a test of rational regime theory in a proglacial braided stream. Geomorphology, 37(1–2), 4363.Google Scholar
Chin, A., 1989. Step pools in stream channels. Progress in Physical Geography, 13(3), 390407.Google Scholar
Chin, A., 1998. On the stability of step-pool mountain streams. Journal of Geology, 106(1), 5969.Google Scholar
Chin, A., 1999a. On the origin of step-pool sequences in mountain streams. Geophysical Research Letters, 26(2), 231234.Google Scholar
Chin, A., 1999b. The morphologic structure of step-pools in mountain streams. Geomorphology, 27(3–4), 191204.Google Scholar
Chin, A., 2002. The periodic nature of step-pool mountain streams. American Journal of Science, 302(2), 144167.Google Scholar
Chin, A., 2003. The geomorphic significance of step-pools in mountain streams. Geomorphology, 55(1–4), 125137.Google Scholar
Chin, A., 2006. Urban transformation of river landscapes in a global context. Geomorphology, 79(3–4), 460487.Google Scholar
Chin, A., Phillips, J. D., 2007. The self-organization of step-pools in mountain streams. Geomorphology, 83(3–4), 346358.Google Scholar
Chin, A., Wohl, E., 2005. Toward a theory for step pools in stream channels. Progress in Physical Geography, 29(3), 275296.Google Scholar
Chin, A., Daniels, M. D., Urban, M. A., et al., 2008. Perceptions of wood in rivers and challenges for stream restoration in the United States. Environmental Management, 41(6), 893903.Google Scholar
Chin, A., Gidley, R., Tyner, L., Gregory, K., 2017. Adjustment of dryland stream channels over four decades of urbanization. Anthropocene, 20, 2436.Google Scholar
Chiu, C. L., Lin, G. F., 1983. Computation of 3-D flow and shear in open channels. Journal of Hydraulic Engineering, 109(11), 14241440.Google Scholar
Chone, G., Biron, P. M., 2016. Assessing the relationship between river mobility and aquatic habitat. River Research and Applications, 32(4), 528539.Google Scholar
Chorley, R. J., Beckinsdale, R. P., 1964. The History of the Study of Landforms or The Development of Geomorphology. Volume One: Geomorphology before Davis. Methuen, London.Google Scholar
Chorley, R. J., Beckinsdale, R. P., Dunn, A. J., 1973. The History of the Study of Landforms or the Development of Geomorphology, Volume 2. The Life and Work of William Morris Davis. Methuen, London.Google Scholar
Chorley, R. J., Schumm, S. A., Sugden, D. E., 1984. Geomorphology. Methuen, London.Google Scholar
Chow, V. T., 1959. Open-Channel Hydraulics. McGraw-Hill, New York.Google Scholar
Chu, V. H., Babarutsi, S., 1988. Confinement and bed-friction effects in shallow turbulent mixing layers. Journal of Hydraulic Engineering, 114, 12571274.Google Scholar
Chuang, S.-C., Chen, H., Lin, G.-W., Lin, C.-W., Chang, C.-P., 2009. Increase in basin sediment yield from landslides in storms following major seismic disturbance. Engineering Geology, 103(1–2), 5965.Google Scholar
Church, M., 1972. Baffin Island Sandurs: A Study of Arctic Fluvial Processes. (Bulletin of the Geological Survey of Canada 216). Ottawa, Canada.Google Scholar
Church, M., 1983. Pattern of instability in a wandering gravel bed channel. In: Collinson, J. D., Lewin, J. (eds.), Modern and Ancient Fluvial Systems, International Association of Sedimentologists Special Publication No. 6. Blackwell, Oxford, UK, pp. 169180.Google Scholar
Church, M., 1995. Geomorphic response to river flow regulation – case studies and time-scales. Regulated Rivers-Research & Management, 11(1), 322. 10.1002/rrr.3450110103.Google Scholar
Church, M., 2006. Bed material transport and the morphology of alluvial river channels. Annual Review of Earth and Planetary Sciences, 34, 325354.Google Scholar
Church, M., 2011. Observations and experiments. In: Gregory, K. J., Goudie, A. S. (eds.), The Sage Handbook of Geomorphology. Sage Publications, London, pp. 121141.Google Scholar
Church, M., Ferguson, R. I., 2015. Morphodynamics: rivers beyond steady state. Water Resources Research, 51(4), 1883–1897. 10.1002/2014wr016862.Google Scholar
Church, M., Haschenburger, J. K., 2017. What is the “active layer”? Water Resources Research, 53(1), 510. 10.1002/2016wr019675.Google Scholar
Church, M., Hassan, M. A., 1992. Size and distance of travel of unconstrained clasts on a streambed. Water Resources Research, 28(1), 299303.Google Scholar
Church, M., Kellerhals, R., 1978. Statistics of grain-size variation along a gravel river. Canadian Journal of Earth Sciences, 15(7), 11511160.Google Scholar
Church, M., Rice, S. P., 2009. Form and growth of bars in a wandering gravel-bed river. Earth Surface Processes and Landforms, 34(10), 14221432.Google Scholar
Church, M., Slaymaker, O., 1989. Disequilibrum of Holocene sediment yield in glaciated British Columbia. Nature, 337(6206), 452454.Google Scholar
Church, M., Zimmermann, A., 2007. Form and stability of step-pool channels: research progress. Water Resources Research, 43(3). 10.1029/2006wr005037.Google Scholar
Church, M. A., McLean, D. G., Wolcott, J. F., 1987. River bed gravels: sampling and analysis. In: Thorne, C. R., Bathurst, J. C., Hey, R. D. (eds.), Sediment Transport in Gravel-Bed Rivers. Wiley, Chichester, UK, pp. 4348.Google Scholar
Church, M., Ham, D., Hassan, M., Slaymaker, O., 1999. Fluvial clastic sediment yield in Canada: Scaled analysis. Canadian Journal of Earth Sciences, 36(8), 12671280.Google Scholar
Citterio, A., Piegay, H., 2009. Overbank sedimentation rates in former channel lakes: Characterization and control factors. Sedimentology, 56(2), 461482.Google Scholar
Claude, N., Rodrigues, S., Bustillo, V., et al., 2012. Estimating bedload transport in a large sand-gravel bed river from direct sampling, dune tracking and empirical formulas. Geomorphology, 179, 4057.Google Scholar
Clayton, J. A., Pitlick, J., 2007. Spatial and temporal variations in bed load transport intensity in a gravel bed river bend. Water Resources Research, 43(2). 10.1029/2006wr005253.Google Scholar
Clifford, N. J., 1993. Differential bed sedimentology and the maintenance of riffle-pool sequences. Catena, 20(5), 447468.Google Scholar
Clifford, N. J., Richards, K. S., 1992. The reversal hypothesis and the maintenance of riffle-pool sequences: A review and field appraisal. In: Carling, P. A., Petts, G. E. (eds.), Lowland Floodplain Rivers: Geomorphological Perspectives. Wiley, Chichester, UK, pp. 4370.Google Scholar
Clifford, N. J., Hardisty, J., French, J. R., Hart, S., 1993. Downstream variation in bed material characteristics: A turbulence-controlled form-process feedback mechanism In: Best, J. L., Bristow, C. S. (eds.), Braided Rivers. Geological Society of London Special Publication 75, Geological Society, London, pp. 89104.Google Scholar
Clubb, F. J., Mudd, S. M., Milodowski, D. T., Hurst, M. D., Slater, L. J., 2014. Objective extraction of channel heads from high-resolution topographic data. Water Resources Research, 50(5), 42834304. 10.1002/2013wr015167.Google Scholar
Cockerill, K., Anderson, W. P., Jr., 2014. Creating false images: stream restoration in an urban setting. Journal of the American Water Resources Association, 50, 468482.Google Scholar
Cohen, S., Kettner, A. J., Syvitski, J. P. M., 2014. Global suspended sediment and water discharge dynamics between 1960 and 2010: continental trends and intra-basin sensitivity. Global and Planetary Change, 115, 4458.Google Scholar
Cohen, Y., Devauchelle, O., Seybold, H. F., et al., 2015. Path selection in the growth of rivers. Proceedings of the National Academy of Sciences of the United States of America, 112(46), 1413214137.Google Scholar
Colberg, J. S., Anders, A. M., 2014. Numerical modeling of spatially-variable precipitation and passive margin escarpment evolution. Geomorphology, 207, 203212.Google Scholar
Collins, A. L., Walling, D. E., 2004. Documenting catchment suspended sediment sources: Problems, approaches and prospects. Progress in Physical Geography, 28(2), 159196.Google Scholar
Collins, A. L., Walling, D. E., Leeks, G. J. L., 1997. Source type ascription for fluvial suspended sediment based on a quantitative composite fingerprinting technique. Catena, 29(1), 127.Google Scholar
Collins, A. L., Walling, D. E., Leeks, G. J. L., 1998. Use of composite fingerprints to determine the provenance of the contemporary suspended sediment load transported by rivers. Earth Surface Processes and Landforms, 23(1), 3152.Google Scholar
Collins, A. L., Walling, D. E., Webb, L., King, P., 2010. Apportioning catchment scale sediment sources using a modified composite fingerprinting technique incorporating property weightings and prior information. Geoderma, 155(3–4), 249261.Google Scholar
Collins, B. D., Dunne, T., 1989. Gravel transport, gravel harvesting, and channel-bed degradation in rivers draining the southern Olympic Mountains, Washington, USA. Environmental Geology and Water Sciences, 13(3), 213224.Google Scholar
Collins, B. D., Montgomery, D. R., Fetherston, K. L., Abbe, T. B., 2012. The floodplain large-wood cycle hypothesis: a mechanism for the physical and biotic structuring of temperate forested alluvial valleys in the North Pacific coastal ecoregion. Geomorphology, 139, 460470.Google Scholar
Collinson, J. D., 1996. Alluvial sedimentation. In: Reading, H. G. (ed.), Sedimentary Environments: Processes, Facies, and Stratigraphy. Blackwell Science, Oxford, UK, pp. 3782.Google Scholar
Colombini, M., 1993. Turbulence-driven secondary flows and formation of sand ridges. Journal of Fluid Mechanics, 254, 701719.Google Scholar
Colombini, M., Parker, G., 1995. Longitudinal streaks. Journal of Fluid Mechanics, 304, 161183.Google Scholar
Colombini, M., Seminara, G., Tubino, M., 1987. Finite-amplitude alternate bars. Journal of Fluid Mechanics, 181, 213232.Google Scholar
Colosimo, M. F., Wilcock, P. R., 2007. Alluvial sedimentation and erosion in an urbanizing watershed, Gwynns Falls, Maryland. Journal of the American Water Resources Association, 43(2), 499521.Google Scholar
Comiti, F., Andreoli, A., Lenzi, M. A., 2005. Morphological effects of local scouring in step-pool streams. Earth Surface Processes and Landforms, 30(12), 15671581.Google Scholar
Comiti, F., Mao, L., Wilcox, A., Wohl, E. E., Lenzi, M. A., 2007. Field-derived relationships for flow velocity and resistance in high-gradient streams. Journal of Hydrology, 340(1–2), 4862.Google Scholar
Constantine, C. R., Mount, M. F., Florsheim, J. L., 2003. The effects of longitudinal differences in gravel mobility on the downstream fining pattern in the Cosumnes River, California. Journal of Geology, 111(2), 233241.Google Scholar
Constantine, C. R., Dunne, T., Hanson, G. J., 2009. Examining the physical meaning of the bank erosion coefficient used in meander migration modeling. Geomorphology, 106(3–4), 242252.Google Scholar
Constantine, J. A., Dunne, T., 2008. Meander cutoff and the controls on the production of oxbow lakes. Geology, 36(1), 2326.Google Scholar
Constantine, J. A., McLean, S. R., Dunne, T., 2010a. A mechanism of chute cutoff along large meandering rivers with uniform floodplain topography. Geological Society of America Bulletin, 122(5–6), 855869.Google Scholar
Constantine, J. A., Dunne, T., Piegay, H., Kondolf, G. M., 2010b. Controls on the alluviation of oxbow lakes by bed-material load along the Sacramento River, California. Sedimentology, 57(2), 389407.Google Scholar
Constantine, J. A., Dunne, T., Ahmed, J., Legleiter, C., Lazarus, E. D., 2014. Sediment supply as a driver of river meandering and floodplain evolution in the Amazon Basin. Nature Geoscience, 7(12), 899903.Google Scholar
Constantinescu, G., Koken, M., Zeng, J., 2011a. The structure of turbulent flow in an open channel bend of strong curvature with deformed bed: Insight provided by detached eddy simulation. Water Resources Research, 47. 10.1029/2010wr010114.Google Scholar
Constantinescu, G., Miyawaki, S., Rhoads, B., Sukhodolov, A., Kirkil, G., 2011b. Structure of turbulent flow at a river confluence with momentum and velocity ratios close to 1: Insight provided by an eddy-resolving numerical simulation. Water Resources Research, 47. 10.1029/2010wr010018.Google Scholar
Constantinescu, G., Miyawaki, S., Rhoads, B., Sukhodolov, A., 2012. Numerical analysis of the effect of momentum ratio on the dynamics and sediment-entrainment capacity of coherent flow structures at a stream confluence. Journal of Geophysical Research – Earth Surface, 117. 10.1029/2012jf002452.Google Scholar
Constantinescu, G., Miyawaki, S., Rhoads, B., Sukhodolov, A., 2014. Numerical evaluation of the effects of planform geometry and inflow conditions on flow, turbulence structure, and bed shear velocity at a stream confluence with a concordant bed. Journal of Geophysical Research – Earth Surface, 119(10), 20792097. 10.1002/2014jf003244.Google Scholar
Constantinescu, G., Miyawaki, S., Rhoads, B., Sukhodolov, A., 2016. Influence of planform geometry and momentum ratio on thermal mixing at a stream confluence with a concordant bed. Environmental Fluid Mechanics, 16(4), 845873.Google Scholar
Contos, J., Tripcevich, N., 2014. Correct placement of the most distant source of the Amazon River in the Mantaro River drainage. Area, 46(1), 2739.Google Scholar
Cook, K. L., Turowski, J. M., Hovius, N., 2013. A demonstration of the importance of bedload transport for fluvial bedrock erosion and knickpoint propagation. Earth Surface Processes and Landforms, 38(7), 683695.Google Scholar
Coon, W. F., 1998. Estimates of roughness coefficients for selected natural stream channels with vegetated banks in New York. U.S. Geological Survey Open-file Report 93–161, U.S. Geological Survey, Denver, CO.Google Scholar
Cooper, A. H., Brown, T. J., Price, S. J., Ford, J. R., Waters, C. N., 2018. Humans are the most significant global geomorphological driving force of the 21st century. Anthropocene Review, 5(3), 222229.Google Scholar
Coopersmith, E., Yaeger, M. A., Ye, S., Cheng, L., Sivapalan, M., 2012. Exploring the physical controls of regional patterns of flow duration curves – Part 3: a catchment classification system based on regime curve indicators. Hydrology and Earth System Sciences, 16(11), 44674482.Google Scholar
Costa, J. E., 1974. Response and recovery of a Piedmont watershed from tropical storm Agnes, June 1972. Water Resources Research, 10(1), 106112.Google Scholar
Costa, J. E., 1975. Effects of agriculture on erosion and sedimentation in the Piedmont Province, Maryland. Geological Society of America Bulletin, 86(9), 12811286.Google Scholar
Costa, J. E., O’Connor, J. E., 1995. Geomorphically effective floods. In: Costa, J. E., Miller, A. J., Potter, K. P., Wilcock, P. R. (eds.), Natural and Anthropogenic Influences in Fluvial Geomorphology (The Wolman Volume). American Geophysical Union, Washington, DC, pp. 4556.Google Scholar
Costigan, K. H., Gerken, J. E., 2016. Channel morphology and flow structure of an abandoned channel under varying stages. Water Resources Research, 52(7), 54585472. 10.1002/2015wr017601.Google Scholar
Costigan, K. H., Daniels, M. D., Perkin, J. S., Gido, K. B., 2014. Longitudinal variability in hydraulic geometry and substrate characteristics of a Great Plains sand-bed river. Geomorphology, 210, 4858.Google Scholar
Coulthard, T. J., 2005. Effects of vegetation on braided stream pattern and dynamics. Water Resources Research, 41(4). 10.1029/2004wr003201.Google Scholar
Coulthard, T. J., van de Wiel, M. J., 2013a. Numerical modeling in fluvial geomorphology. In: Shroder, J. W. (ed.), Treatise on Geomorphology, Vol. 9, Fluvial Geomorphology, Wohl, E. (vol. ed.). Academic Press, San Diego, CA, pp. 694710.Google Scholar
Coulthard, T. J., Van de Wiel, M. J., 2013b. Climate, tectonics or morphology: what signals can we see in drainage basin sediment yields? Earth Surface Dynamics, 1(1), 1327.Google 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(3), 269288.Google Scholar
Coulthard, T. J., Hancock, G. R., Lowry, J. B. C., 2012. Modelling soil erosion with a downscaled landscape evolution model. Earth Surface Processes and Landforms, 37(10), 10461055.Google Scholar
Covault, J. A., Craddock, W. H., Romans, B. W., Fildani, A., Gosai, M., 2013. Spatial and temporal variations in landscape evolution: Historic and longer-term sediment flux through global catchments. Journal of Geology, 121(1), 3556.Google Scholar
Cowan, W. L., 1956. Estimating hydraulic roughness coefficients. Agricultural Engineering, 37, 473475.Google Scholar
Cox, N. J., Warburton, J., Armstrong, A., Holliday, V. J., 2008. Fitting concentration and load rating curves with generalized linear models. Earth Surface Processes and Landforms, 33(1), 2539.Google Scholar
Creelle, S., Schindfessel, L., De Mulder, T., 2017. Modelling of the tributary momentum contribution to predict confluence head losses. Journal of Hydraulic Research, 55(2), 175189.Google Scholar
Crosato, A., Mosselman, E., 2009. Simple physics-based predictor for the number of river bars and the transition between meandering and braiding. Water Resources Research, 45. 10.1029/2008wr007242.Google Scholar
Crosato, A., Mosselman, E., Desta, F. B., Uijttewaal, W. S. J., 2011. Experimental and numerical evidence for intrinsic nonmigrating bars in alluvial channels. Water Resources Research, 47. 10.1029/2010wr009714.Google Scholar
Crosato, A., Desta, F. B., Cornelisse, J., Schuurman, F., Uijttewaal, W. S. J., 2012. Experimental and numerical findings on the long-term evolution of migrating alternate bars in alluvial channels. Water Resources Research, 48. 10.1029/2011wr011320.Google Scholar
Crosby, B. T., Whipple, K. X., 2006. Knickpoint initiation and distribution within fluvial networks: 236 waterfalls in the Waipaoa River, North Island, New Zealand. Geomorphology, 82(1–2), 1638.Google Scholar
Crosby, B. T., Whipple, K. X., Gasparini, N. M., Wobus, C. W., 2007. Formation of fluvial hanging valleys: Theory and simulation. Journal of Geophysical Research – Earth Surface, 112(F3). 10.1029/2006jf000566.Google Scholar
Crowder, D. W., Knapp, H. V., 2005. Effective discharge recurrence intervals of Illinois streams. Geomorphology, 64(3–4), 167184.Google Scholar
Crowder, D. W., Demissie, M., Markus, M., 2007. The accuracy of sediment loads when log-transformation produces nonlinear sediment load-discharge relationships. Journal of Hydrology, 336(3–4), 250268.Google Scholar
Cruise, J. F., Laymon, C. A., Al-Hamdan, O. Z., 2010. Impact of 20 years of land-cover change on the hydrology of streams in the southeastern United States. Journal of the American Water Resources Association, 46(6), 11591170.Google Scholar
Crutzen, P. J., 2002. Geology of mankind. Nature, 415(6867), 2323.Google Scholar
Csiki, S., Rhoads, B. L., 2010. Hydraulic and geomorphological effects of run-of-river dams. Progress in Physical Geography, 34(6), 755780.Google Scholar
Csiki, S. J. C., Rhoads, B. L., 2014. Influence of four run-of-river dams on channel morphology and sediment characteristics in Illinois, USA. Geomorphology, 206, 215229.Google Scholar
Cui, Y. T., Parker, G., 1998. The arrested gravel front: Stable gravel-sand transitions in rivers – Part 2: General numerical solution. Journal of Hydraulic Research, 36(2), 159182.Google Scholar
Cui, Y. T., Parker, G., Paola, C., 1996. Numerical simulation of aggradation and downstream fining. Journal of Hydraulic Research, 34(2), 185204.Google Scholar
Cui, Y. T., Parker, G., Braudrick, C., Dietrich, W. E., Cluer, B., 2006. Dam removal express assessment models (DREAM). Part 1: model development and validation. Journal of Hydraulic Research, 44(3), 291307.Google Scholar
Cui, Y., Booth, D. B., Monschke, J., et al., 2017. Analyses of the erosion of fine sediment deposit for a large dam-removal project: An empirical approach. International Journal of River Basin Management, 15(1), 103114.Google Scholar
Curran, J. C., 2007. Step-pool formation models and associated step spacing. Earth Surface Processes and Landforms, 32(11), 16111627. 10.1002/esp.1589.Google Scholar
Curran, J. C., Wilcock, P. R., 2005. Characteristic dimensions of the step-pool bed configuration: an experimental study. Water Resources Research, 41(2). 10.1029/2004wr003568.Google Scholar
Curran, J. H., Wohl, E. E., 2003. Large woody debris and flow resistance in step-pool channels, Cascade Range, Washington. Geomorphology, 51(1–3), 141157.Google Scholar
Curtis, K. E., Renshaw, C. E., Magilligan, F. J., Dade, W. B., 2010. Temporal and spatial scales of geomorphic adjustments to reduced competency following flow regulation in bedload-dominated systems. Geomorphology, 118(1–2), 105117.Google Scholar
Czegledi, I., Saly, P., Takacs, P., et al., 2016. The scales of variability of stream fish assemblages at tributary confluences. Aquatic Sciences, 78(4), 641654.Google 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. 10.1002/2014wr016139.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(3), 18491870. 10.1029/2018wr023527.Google Scholar
da Silva, A. M. F., 2006. On why and how do rivers meander. Journal of Hydraulic Research, 44(5), 579590.Google Scholar
da Silva, A. M. F., 2009. On the stable geometry of self-formed alluvial channels: theory and practical application. Canadian Journal of Civil Engineering, 36(10), 16671679.Google Scholar
da Silva, A. M. F., Ahmari, H., 2009. Size and effect on the mean flow of large-scale horizontal coherent structures in open-channel flows: An experimental study. Canadian Journal of Civil Engineering, 36(10), 16431655.Google Scholar
Dade, W. B., Friend, P. F., 1998. Grain-size, sediment-transport regime, and channel slope in alluvial rivers. Journal of Geology, 106(6), 661675.Google Scholar
Dadson, S. J., Hovius, N., Chen, H., et al., 2004. Earthquake-triggered increase in sediment delivery from an active mountain belt. Geology, 32(8), 733736.Google Scholar
Dai, A., 2016. Historical and future changes in streamflow and continental runoff: A review. In: Tang, Q., Oki, T. (eds.), Terrestrial Water Cycle and Climate Change: Natural and Human-induced Impacts. American Geophysical Union, Washington, DC, pp. 1737.Google Scholar
Dalrymple, T., 1960. Flood Frequency Analyses. Manual of Hydrology, Part 3. Flood-Flow Techniques, U.S. Geological Survey Water-supply Paper 1543-A. U.S. Government Printing Office, Washington, DC.Google Scholar
D’Ambrosio, J. L., Ward, A. D., Witter, J. D., 2015. Evaluating geomorphic change in constructed two-stage ditches. Journal of the American Water Resources Association, 51(4), 910922.Google Scholar
Daniels, M. D., Rhoads, B. L., 2004. Effect of large woody debris configuration on three-dimensional flow structure in two low-energy meander bends at varying stages. Water Resources Research, 40(11). 10.1029/2004wr003181.Google Scholar
Daniels, R. B., 1960. Entrenchment of the Willow drainage ditch, Harrison County, Iowa. American Journal of Science, 258(3), 161176.Google Scholar
Darby, S. E., Thorne, C. R., 1992. Impact of channelization on the Mimmshall Brook, Hertfordshire, UK. Regulated Rivers Research and Management, 7(2), 193204.Google Scholar
Darby, S. E., Thorne, C. R., 1996. Development and testing of riverbank-stability analysis. Journal of Hydraulic Engineering, ASCE, 122(8), 443454. 10.1061/(asce)0733–9429(1996)122:8(443).Google Scholar
Darby, S. E., Gessler, D., Thorne, C. R., 2000. Computer program for stability analysis of steep, cohesive riverbanks. Earth Surface Processes and Landforms, 25(2), 175190.Google Scholar
Darby, S. E., Alabyan, A. M., Van de Wiel, M. J., 2002. Numerical simulation of bank erosion and channel migration in meandering rivers. Water Resources Research, 38(9), 2–1-2–12.Google Scholar
Darby, S. E., Rinaldi, M., Dapporto, S., 2007. Coupled simulations of fluvial erosion and mass wasting for cohesive river banks. Journal of Geophysical Research – Earth Surface, 112(F3). 10.1029/2006jf000722.Google Scholar
David, G. C. L., Wohl, E., Yochum, S. E., Bledsoe, B. P., 2010. Controls on at-a-station hydraulic geometry in steep headwater streams, Colorado, USA. Earth Surface Processes and Landforms, 35(15), 18201837. 10.1002/esp.2023.Google Scholar
David, S. R., Edmonds, D. A., Letsinger, S. L., 2017. Controls on the occurrence and prevalence of floodplain channels in meandering rivers. Earth Surface Processes and Landforms, 42(3), 460472. 10.1002/esp.4002.Google Scholar
Davidson, G. R., Carnley, M., Lange, T., Galicki, S. J., Douglas, A., 2004. Changes in sediment accumulation rate in an oxbow lake following late 19th century clearing of land for agricultural use: A Pb-210, Cs-137, and C-14 study in Mississippi, USA. Radiocarbon, 46(2), 755764.Google Scholar
Davidson, S. K., Hey, R. D., 2011. Regime equations for natural meandering cobble- and gravel-bed rivers. Journal of Hydraulic Engineering, 137(9), 894910.Google Scholar
Davidson, S. L., Eaton, B. C., 2018. Beyond regime: A stochastic model of floods, bank erosion, and channel migration. Water Resources Research, 54(9), 62826298. 10.1029/2017wr022059.Google Scholar
Davies, T. R. H., Sutherland, A. J., 1983. Extremal hypotheses for river behavior. Water Resources Research, 19(1), 141148.Google Scholar
Davies, T. R. H., Tinker, C. C., 1984. Fundamental characteristics of stream meanders. Geological Society of America Bulletin, 95(5), 505512.Google Scholar
Davies-Vollum, K. S., Smith, N. D., 2008. Factors affecting the accumulation of organic-rich deposits in a modern avulsive floodplain: Examples from the Cumberland Marshes, Saskatchewan, Canada. Journal of Sedimentary Research, 78(9–10), 683692.Google Scholar
Davis, R. T., Tank, J. L., Mahl, U. H., Winikoff, S. G., Roley, S. S., 2015. The influence of two-stage ditches with constructed floodplains on water column nutrients and sediments in agricultural streams. Journal of the American Water Resources Association, 51(4), 941955.Google Scholar
Davis, W. M., 1889. The rivers and valleys of Pennsylvania. National Geographic Magazine, 1, 1126.Google Scholar
Davis, W. M., 1899. The geographical cycle. Geographical Journal, 14, 481504.Google Scholar
Davis, W. M., 1902. Base level, grade and peneplain. Journal of Geology, 10, 77111.Google Scholar
Davis, W. M., 1903. The development of river meanders. The Geological Magazine, New Series, Decade IV, Vol. X, 145148.Google Scholar
Davoren, A., Mosley, M. P., 1986. Observations of bedload movement, bar development and sediment supply in the braided Ohua River. Earth Surface Processes and Landforms, 11(6), 643652.Google Scholar
Davy, B. W., Davies, T. R. H., 1979. Entropy concepts in fluvial geomorphology – re-evaluation. Water Resources Research, 15(1), 103106.Google Scholar
Day, S. S., Gran, K. B., Belmont, P., Wawrzyniec, T., 2013. Measuring bluff erosion part 2: Pairing aerial photographs and terrestrial laser scanning to create a watershed scale sediment budget. Earth Surface Processes and Landforms, 38(10), 10681082.Google Scholar
de Almeida, G. A. M., Rodriguez, J. F., 2011. Understanding pool-riffle dynamics through continuous morphological simulations. Water Resources Research, 47. 10.1029/2010wr009170.Google Scholar
de Almeida, G. A. M., Rodriguez, J. F., 2012. Spontaneous formation and degradation of pool-riffle morphology and sediment sorting using a simple fractional transport model. Geophysical Research Letters, 39. 10.1029/2012gl051059.Google Scholar
de Azeredo Freitas, H. R., Freitas, C. d. C., Rosim, S., de Freitas Oliveira, J. R., 2016. Drainage networks and watersheds delineation derived from TIN-based digital elevation models. Computers & Geosciences, 92, 2137.Google Scholar
De Cacqueray, N., Hargreaves, D. M., Morvan, H. P., 2009. A computational study of shear stress in smooth rectangular channels. Journal of Hydraulic Research, 47(1), 5057.Google Scholar
De Cesare, G., Schleiss, A., Hermann, F., 2001. Impact of turbidity currents on reservoir sedimentation. Journal of Hydraulic Engineering, 127(1), 616.Google Scholar
De Serres, B., Roy, A. G., 1990. Flow direction and branching geometry at junctions in dendritic river networks. Professional Geographer, 42(2), 194201.Google Scholar
De Serres, B., Roy, A. G., Biron, P. M., Best, J. L., 1999. Three-dimensional structure of flow at a confluence of river channels with discordant beds. Geomorphology, 26(4), 313335.Google Scholar
de Vente, J., Poesen, J., Arabkhedri, M., Verstraeten, G., 2007. The sediment delivery problem revisited. Progress in Physical Geography, 31(2), 155178.Google Scholar
de Vente, J., Poesen, J., Verstraeten, G., et al., 2013. Predicting soil erosion and sediment yield at regional scales: Where do we stand? Earth-Science Reviews, 127, 1629.Google Scholar
Dean, D. J., Schmidt, J. C., 2013. The geomorphic effectiveness of a large flood on the Rio Grande in the Big Bend region: Insights on geomorphic controls and post-flood geomorphic response. Geomorphology, 201, 183198.Google Scholar
Dearing, J. A., Jones, R. T., 2003. Coupling temporal and spatial dimensions of global sediment flux through lake and marine sediment records. Global and Planetary Change, 39(1–2), 147168.Google Scholar
Dedkov, A., 2004. The relationship between sediment yield and drainage basin area. In: Golosov, V., Belyaev, V., Walling, D. E. (eds.), Sediment Transfer through the Fluvial System. IAHS Publication No. 288, IAHS Press, Wallingford, UK, pp. 197204.Google Scholar
Dedkov, A. P., Gusarov, A. V., 2006. Suspended sediment yield from continents into the World Ocean: spatial and temporal changeability. In: Rowan, J. S., Duck, R. W., Werritty, A. (eds.), Sediment Dynamics and the Hydromorphology of Fluvial Systems. IAHS Publication No. 306, IAHS Press, Wallingford, UK, pp. 311.Google Scholar
Dedkov, A., Moszherin, V. I., 1992. Erosion and sediment yields in mountain regions of the world. In: Walling, D. E., Davies, T. R., Hasholt, B. (eds.), Erosion, Debris Flows and Environment in Mountain Regions. IAHS Publication No. 209, IAHS Press, Wallingford, UK, pp. 2936.Google Scholar
Defina, A., 2003. Numerical experiments on bar growth. Water Resources Research, 39(4). 10.1029/2002wr001455.Google Scholar
Delmas, M., Pak, L. T., Cerdan, O., et al., 2012. Erosion and sediment budget across scale: A case study in a catchment of the European loess belt. Journal of Hydrology, 420, 255263.Google Scholar
Desloges, J. R., Church, M., 1987. Channel and floodplain facies in a wandering gravel-bed river. In: Ethridge, F. G., Flores, R. M., Harvey, M. D. (eds.), Recent Developments in Fluvial Sedimentology, SEPM Special Publication No. 39. SEPM, Tulsa, OK, pp. 99109.Google Scholar
Desloges, J. R., Church, M. A., 1989. Canadian landform examples – 13: Wandering gravel-bed rivers. Canadian Geographer, 33(4), 360364.Google Scholar
Devauchelle, O., Petroff, A. P., Seybold, H. F., Rothman, D. H., 2012. Ramification of stream networks. Proceedings of the National Academy of Sciences of the United States of America, 109(51), 2083220836.Google Scholar
DeVries, P., 2003. Bedload layer thickness and disturbance depth in gravel bed streams. Journal of Hydraulic Engineering, 128, 983991.Google Scholar
Dey, A. K., Tsujimoto, T., Kitamura, T., 2007. Experimental investigations on different modes of headcut migration. Journal of Hydraulic Research, 45(3), 333346.Google Scholar
Dey, S., 2014. Fluvial Hydrodynamics: Hydrodynamic and Sediment Transport Phenomena. Springer, Berlin.Google Scholar
Dey, S., Sarkar, S., Solari, L., 2011. Near-bed turbulence characteristics at the entrainment threshold of sediment beds. Journal of Hydraulic Engineering, 137(9), 945958.Google Scholar
DeZiel, B. A., Krider, L., Hansen, B., et al., 2019. Habitat improvements and fish community response associated with an agricultural two-stage ditch in Mower County, Minnesota. Journal of the American Water Resources Association, 55(1), 154188.Google Scholar
D’Haen, K., Verstraeten, G., Degryse, P., 2012. Fingerprinting historical fluvial sediment fluxes. Progress in Physical Geography, 36(2), 154186.Google Scholar
Dieras, P. L., Constantine, J. A., Hales, T. C., Piegay, H., Riquier, J., 2013. The role of oxbow lakes in the off-channel storage of bed material along the Ain River, France. Geomorphology, 188, 110119.Google Scholar
Dietrich, J. T., 2016. Riverscape mapping with helicopter-based Structure-from-Motion photogrammetry. Geomorphology, 252, 144157.Google Scholar
Dietrich, W. E., 1982. Settling velocity of natural particles. Water Resources Research, 18(6), 16151626.Google Scholar
Dietrich, W. E., 1987. Mechanics of flow and sediment transport in river bends. In: Richards, K. S. (ed.), River Channels: Environment and Process. Basil Blackwell, London, pp. 179227.Google Scholar
Dietrich, W. E., Dunne, T., 1978. Sediment budget for a small catchment in mountainous terrain. Zeitschrift fur Geomorphologie, Supplementband 29, 191206.Google Scholar
Dietrich, W. E., Dunne, T., 1993. Channel heads. In: Kirkby, M. J., Beven, K. (eds.), Channel Network Hydrology. Wiley, New York, pp. 175219.Google Scholar
Dietrich, W. E., Smith, J. D., 1983. Influence of the point bar on flow through curved channels. Water Resources Research, 19(5), 11731192.Google Scholar
Dietrich, W. E., Smith, J. D., 1984. Bed load transport in a river meander. Water Resources Research, 20(10), 13551380.Google Scholar
Dietrich, W. E., Whiting, P. J., 1989. Boundary shear stress and sediment transport in river meanders of sand and gravel. In: Ikeda, S., Parker, G. (eds.), River Meandering. American Geophysical Union, Washington, DC., pp. 150.Google Scholar
Dietrich, W. E., Smith, J. D., Dunne, T., 1979. Flow and sediment transport in a sand bedded meander. Journal of Geology, 87(3), 305315.Google Scholar
Dietrich, W. E., Kirchner, J. W., Ikeda, H., Iseya, F., 1989. Sediment supply and the development of the coarse surface-layer in gravel-bedded rivers. Nature, 340(6230), 215217.Google Scholar
Dietrich, W. E., Wilson, C. J., Montgomery, D. R., McKean, J., Bauer, R., 1992. Erosion thresholds and land surface morphology. Geology, 20(8), 675679.Google Scholar
Dietrich, W. E., Wilson, C. J., Montgomery, D. R., McKean, J., 1993. Analysis of erosion thresholds, channel networks, and landscape morphology using a digital terrain model. Journal of Geology, 101(2), 259278.Google Scholar
Dietrich, W. E., Bellugi, G. E., Sklar, L. S., et al., 2003. Geomorphic transport laws for predicting landscape form and dynamics. In: Wilcock, P. R., Iverson, R. M. (eds.), Prediction in Geomorphology. American Geophysical Union, Washington, DC, pp. 103132.Google Scholar
Diffenbaugh, N. S., Singh, D., Mankin, J. S., et al., 2017. Quantifying the influence of global warming on unprecedented extreme climate events. Proceedings of the National Academy of Sciences of the United States of America, 114(19), 48814886.Google Scholar
Dinehart, R. L., 1992. Evolution of coarse gravel bed forms – field measurements at flood stage. Water Resources Research, 28(10), 26672689.Google Scholar
Ding, Y., Langendoen, E. J., 2018. Simulation and control of sediment transport due to dam removal. Journal of Applied Water Engineering and Research, 6(2), 95108.Google Scholar
Dingman, S. L., 1984. Fluvial Hydrology. W. H. Freeman and Co., New York.Google Scholar
Dingman, S. L., 2007. Analytical derivation of at-a-station hydraulic-geometry relations. Journal of Hydrology, 334(1–2), 1727.Google Scholar
Dingman, S. L., 2009. Fluvial Hydraulics. Oxford University Press, New York.Google Scholar
Diplas, P., 1987. Bedload transport in gravel-bed streams. Journal of Hydraulic Engineering, 113(3), 277292.Google Scholar
Diplas, P., 1990. Characteristics of self-formed straight channels. Journal of Hydraulic Engineering, 116(5), 707728.Google Scholar
Diplas, P., Vigilar, G., 1992. Hydraulic geometry of threshold channels. Journal of Hydraulic Engineering, 118(4), 597614.Google Scholar
Diplas, P., Dancey, C. L., Celik, A. O., et al., 2008. The role of impulse on the initiation of particle movement under turbulent flow conditions. Science, 322(5902), 717720.Google Scholar
Dixon, S. J., Sambrook Smith, G. H., Best, J. L., et al., 2018. The planform mobility of river channel confluences: Insights from analysis of remotely sensed imagery. Earth-Science Reviews, 176, 118.Google Scholar
Dodds, P. S., Rothman, D. H., 2000. Scaling, universality, and geomorphology. Annual Review of Earth and Planetary Sciences, 28, 571610.Google Scholar
Dodds, P. S., Rothman, D. H., 2001a. Geometry of river networks. I. Scaling, fluctuations, and deviations. Physical Review E, 63(1), 016115.Google Scholar
Dodds, P. S., Rothman, D. H., 2001b. Geometry of river networks. II. Distributions of component size and number. Physical Review E, 63(1), 016116.Google Scholar
Dodds, P. S., Rothman, D. H., 2001c. Geometry of river networks. III. Characterization of component connectivity. Physical Review E, 63(1), 016117.Google Scholar
Dodov, B., Foufoula-Georgiou, E., 2004. Generalized hydraulic geometry: Derivation based on a multiscaling formalism. Water Resources Research, 40(6). 10.1029/2003wr002082.Google Scholar
Doeschl, A. B., Ashmore, P. E., Davison, M., 2006. Methods for assessing exploratory computational models of braided rivers. In: Smith, G. H. S., Best, J. L., Bristow, C. S., Petts, G. E. (eds.), Braided Rivers: Process, Deposits, Ecology and Management. Special Publications of the International Association of Sedimentologists, 36, Blackwell, Malden, MA, pp. 177197.Google Scholar
Doeschl-Wilson, A. B., Ashmore, P. E., 2005. Assessing a numerical cellular braided-stream model with a physical model. Earth Surface Processes and Landforms, 30(5), 519540.Google Scholar
Domokos, G., Jerolmack, D. J., Sipos, A. A., Toeroek, A., 2014. How river rocks round: Resolving the shape-size paradox. PLoS ONE, 9(2). 10.1371/journal.pone.0088657.Google Scholar
Dooge, J. C. I., 1991. The Manning formula in context. In: Yen, B. C. (ed.), Channel Flow Resistance: Centennial of Manning’s Formula. Water Resources Publications, Highland Ranch, Colorado, pp. 136185.Google Scholar
Dordevic, D., 2013. Numerical study of 3D flow at right-angled confluences with and without upstream planform curvature. Journal of Hydroinformatics, 15(4), 10731088.Google Scholar
Dotterweich, M., 2013. The history of human-induced soil erosion: geomorphic legacies, early descriptions and research, and the development of soil conservation – a global synopsis. Geomorphology, 201, 134.Google Scholar
Douglas, I., 1967. Man, vegetation and the sediment yields of rivers. Nature, 215(5104), 925928.Google Scholar
Downs, P. W., Gregory, K. J., 1995. Approaches to river channel sensitivity. Professional Geographer, 47(2), 168175.Google Scholar
Downs, P. W., Kondolf, G. M., 2002. Post-project appraisals in adaptive management of river channel restoration. Environmental Management, 29(4), 477496.Google Scholar
Downs, P. W., Soar, P. J., Taylor, A., 2016. The anatomy of effective discharge: The dynamics of coarse sediment transport revealed using continuous bedload monitoring in a gravel-bed river during a very wet year. Earth Surface Processes and Landforms, 41(2), 147161.Google Scholar
Doyle, M. W., Shields, C. A., 2008. An alternative measure of discharge effectiveness. Earth Surface Processes and Landforms, 33(2), 308316.Google Scholar
Doyle, M. W., Shields, F. D., 2012. Compensatory mitigation for streams under the Clean Water Act: Reassessing science and redirecting policy. Journal of the American Water Resources Association, 48(3), 494509.Google Scholar
Doyle, M. W., Stanley, E. H., Harbor, J. M., 2003. Channel adjustments following two dam removals in Wisconsin. Water Resources Research, 39(1). 10.1029/2002wr001714.Google Scholar
Doyle, M. W., Shields, D., Boyd, K. F., Skidmore, P. B., Dominick, D., 2007. Channel-forming discharge selection in river restoration design. Journal of Hydraulic Engineering, 133(7), 831837.Google Scholar
Doyle, M. W., Lave, R., Robertson, M. M., Ferguson, J., 2013. River federalism. Annals of the Association of American Geographers, 103(2), 290298.Google Scholar
Doyle, M. W., Singh, J., Lave, R., Robertson, M. M., 2015. The morphology of streams restored for market and nonmarket purposes: Insights from a mixed natural-social science approach. Water Resources Research, 51(7), 56035622. 10.1002/2015wr017030.Google Scholar
Drake, T. G., Shreve, R. L., Dietrich, W. E., Whiting, P. J., Leopold, L. B., 1988. Bedload transport of fine gravel observed by motion-picture photography. Journal of Fluid Mechanics, 192, 193217.Google Scholar
Duan, J. G., Julien, P. Y., 2010. Numerical simulation of meandering evolution. Journal of Hydrology, 391(1–2), 3648.Google Scholar
Duckson, D. W., Duckson, L. J., 1995. Morphology of bedrock step pool systems. Water Resources Bulletin, 31(1), 4351.Google Scholar
Duckson, D. W., Duckson, L. J., 2001. Channel bed steps and pool shapes along Soda Creek, Three Sisters Wilderness, Oregon. Geomorphology, 38(3–4), 267279. 10.1016/s0169-555x(00)00098-2.Google Scholar
Dufour, S., Piégay, H., 2009. From the myth of a lost paradise to targeted river restoration: Forget natural references and focus on human benefits. River Research and Applications, 25(5), 568581.Google Scholar
Dunne, T., 1980. Formation and controls of channel networks. Progress in Physical Geography, 4, 211239.Google Scholar
Dunne, T., 1990. Hydrology, mechanics, and geomorphic implications of erosion by subsurface flow. In: Higgins, C. G., Coates, D. R. (eds.), Groundwater Geomorphology: The Role of Subsurface Water in Earth-Surface Processes and Landforms, Geological Society of America Special Paper 252, pp. 128.Google Scholar
Dunne, T., Aubry, B. F., 1986. Evaluation of Horton’s theory of sheetwash and rill erosion on the basis of field experiments. In: Abrahams, A. D. (ed.), Hillslope Processes. Allen and Unwin, Winchester, MA, pp. 3153.Google Scholar
Dunne, T., Leopold, L. B., 1978. Water in Environmental Planning. Freeman, San Francisco, CA.Google Scholar
Dunne, T., Mertes, L. A. K., Meade, R. H., Richey, J. E., Forsberg, B. R., 1998. Exchanges of sediment between the flood plain and channel of the Amazon River in Brazil. Geological Society of America Bulletin, 110(4), 450467.Google Scholar
Durkin, P. R., Hubbard, S. M., Holbrook, J., Boyd, R., 2018. Evolution of fluvial meander-belt deposits and implications for the completeness of the stratigraphic record. Geological Society of America Bulletin, 130(5–6), 721739.Google Scholar
Dury, G. H., 1964. Principles of underfit streams. U.S. Geological Survey Professional Paper 452-A. U.S. Government Printing Office, Washington, DC.Google Scholar
Dury, G. H., 1966. The concept of grade. In: Dury, G. H. (ed.), Essays in Geomorphology. Heinemann, London, pp. 211234.Google Scholar
Dyhouse, G. R., 1985. Stage-frequency analysis at a major river junction. Journal of Hydraulic Engineering, 111(4), 565583.Google Scholar
Dzubakova, K., Piegay, H., Riquier, J., Trizna, M., 2015. Multi-scale assessment of overflow-driven lateral connectivity in floodplain and backwater channels using LiDAR imagery. Hydrological Processes, 29(10), 23152330.Google Scholar
East, A. E., Logan, J. B., Mastin, M. C., et al., 2018. Geomorphic evolution of a gravel-bed river under sediment-starved versus sediment-rich conditions: river response to the world’s largest dam removal. Journal of Geophysical Research – Earth Surface, 123(12), 33383369. 10.1029/2018jf004703.Google Scholar
Eaton, B. C., 2006. Bank stability analysis for regime models of vegetated gravel bed rivers. Earth Surface Processes and Landforms, 31(11), 14381444.Google Scholar
Eaton, B. C., Church, M., 2004. A graded stream response relation for bed load-dominated streams. Journal of Geophysical Research – Earth Surface, 109(F3), 18. 10.1029/2003jf000062.Google Scholar
Eaton, B. C., Church, M., 2007. Predicting downstream hydraulic geometry: a test of rational regime theory. Journal of Geophysical Research – Earth Surface, 112(F3). 10.1029/2006jf000734.Google Scholar
Eaton, B. C., Millar, R. G., 2017. Predicting gravel bed river response to environmental change: The strengths and limitations of a regime-based approach. Earth Surface Processes and Landforms, 42(6), 9941008.Google Scholar
Eaton, B. C., Church, M., Millar, R. G., 2004. Rational regime model of alluvial channel morphology and response. Earth Surface Processes and Landforms, 29(4), 511529.Google Scholar
Eaton, B. C., Church, M., Davies, T. R. H., 2006. A conceptual model for meander initiation in bedload-dominated streams. Earth Surface Processes and Landforms, 31(7), 875891.Google Scholar
Eaton, B. C., Millar, R. G., Davidson, S., 2010. Channel patterns: braided, anabranching, and single-thread. Geomorphology, 120(3–4), 353364.Google Scholar
Eden, S., Tunstall, S. M., Tapsell, S. M., 2000. Translating nature: river restoration as nature culture. Environment and Planning D-Society & Space, 18(2), 257273.Google Scholar
Edwards, T. K., Glysson, G. D., 1999. Field methods for measurement of fluvial sediment. U.S. Geological Survey Techniques of Water Resources Investigations, Book 3, Chapter C2. Reston, VA, pp. 189.Google Scholar
Eekhout, J. P. C., Hoitink, A. J. F., Mosselman, E., 2013. Field experiment on alternate bar development in a straight sand-bed stream. Water Resources Research, 49(12), 83578369. 10.1002/2013wr014259.Google Scholar
Egozi, R., Ashmore, P., 2008. Defining and measuring braiding intensity. Earth Surface Processes and Landforms, 33(14), 21212138.Google Scholar
Egozi, R., Ashmore, P., 2009. Experimental analysis of braided channel pattern response to increased discharge. Journal of Geophysical Research – Earth Surface, 114. 10.1029/2008jf001099.Google Scholar
Einstein, H. A., 1942. Formulas for the transportation of bedload. Transactions of the American Society of Civil Engineers, 117, 561597.Google Scholar
Einstein, H. A., 1950. The bedload function for bedload transportation in open channel flows. Technical Bulletin No. 1026, U.S.D.A., Soil Conservation Service.Google Scholar
Einstein, H. A., Chien, N., 1953. Can the rate of wash load be predicted from the bed-load function? Transactions of the American Geophysical Union, 34, 876882.Google Scholar
Einstein, H. A., Li, H., 1958. Secondary currents in straight channels. Transactions, American Geophysical Union, 39, 10851088.Google Scholar
Einstein, H. A., Shen, H. W., 1964. Study on meandering in straight alluvial channels. Journal of Geophysical Research, 69(24), 52395247.Google Scholar
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(7), 14551483. 10.1002/2013jf003020.Google Scholar
Ellery, W. N., Ellery, K., Rogers, K. H., McCarthy, T. S., 1995. The role of Cyperus papyrus L. In channel blockage and abandonment in the northeastern Okavango Delta, Botswana. African Journal of Ecology, 33(1), 2549.Google Scholar
Ellis, E. R., Church, M., 2005. Hydraulic geometry of secondary channels of lower Fraser River, British Columbia, from acoustic Doppler profiling. Water Resources Research, 41(8). 10.1029/2004wr003777.Google Scholar
Ely, L. L., Enzel, Y., Baker, V. R., Cayan, D. R., 1993. A 5000-year record of extreme floods and climate change in the Southwestern United States. Science, 262(5132), 410412.Google Scholar
Emerson, J. W., 1971. Channelization – a case study. Science, 173(3994), 325326.Google Scholar
Emery, S. B., Perks, M. T., Bracken, L. J., 2013. Negotiating river restoration: The role of divergent reframing in environmental decision-making. Geoforum, 47, 167177.Google Scholar
Emmett, W. W., 1975. The channels and waters of the upper Salmon River area, Idaho. U.S. Geological Survey Professional Paper 870-A. U.S. Government Printing Office, Washington, DC.Google Scholar
Emmett, W. W., Wolman, M. G., 2001. Effective discharge and gravel-bed rivers. Earth Surface Processes and Landforms, 26(13), 13691380.Google Scholar
Endreny, T. A., 2007. Estimation of channel bankfull occurrence from instantaneous discharge data. Journal of Hydrologic Engineering, 12(5), 524531.Google Scholar
Engel, F. L., Rhoads, B. L., 2012. Interaction among mean flow, turbulence, bed morphology, bank failures and channel planform in an evolving compound meander loop. Geomorphology, 163, 7083.Google Scholar
Engel, F. L., Rhoads, B. L., 2016. Three-dimensional flow structure and patterns of bed shear stress in an evolving compound meander bend. Earth Surface Processes and Landforms, 41(9), 12111226.Google Scholar
Engel, F. L., Rhoads, B. L., 2017. Velocity profiles and the structure of turbulence at the outer bank of a compound meander bend. Geomorphology, 295, 191201.Google Scholar
Engelund, F., Skovgaard, O., 1973. Origin of meandering and braiding in alluvial streams. Journal of Fluid Mechanics, 57(FEB6), 289302.Google Scholar
England, J. F., Jr., Cohn, T. A., Faber, B. A., et al., 2018. Guidelines for determining flood flow frequency – Bulletin 17C. U.S. Geological Survey Techniques and Methods Book 4 Chapter B5.Google Scholar
Ergenzinger, P., Schmidt, K. H., 1990. Stochastic elements of bed load transport in a step-pool mountain river. In: Sinninger, R. O., Monbaron, M. (eds.), Hydrology in Mountainous Regions. II – Artificial Reservoirs; Water and Slopes. IAHS Publication No. 194, IAHS Press, Wallingford, UK, pp. 3946.Google Scholar
Ergenzinger, P., Schmidt, K. H., Busskamp, R., 1989. The pebble transmitter system (PETS) – 1st results of a technique for studying coarse material erosion, transport and deposition. Zeitschrift fur Geomorphologie, 33(4), 503508.Google Scholar
Erskine, W. D., 2011. Geomorphic controls on historical channel planform changes on the lower Pages River, Hunter Valley, Australia. Australian Geographer, 42(3), 289307. 10.1080/00049182.2011.595768.Google Scholar
Erskine, W., McFadden, C., Bishop, P., 1992. Alluvial cutoffs as indicators of former channel conditions. Earth Surface Processes and Landforms, 17(1), 2337.Google Scholar
Eshel, G., Levy, G. J., Mingelgrin, U., Singer, M. J., 2004. Critical evaluation of the use of laser diffraction for particle-size distribution analysis. Soil Science Society of America Journal, 68(3), 736743.Google Scholar
Evans, R., 1998. The erosional impacts of grazing animals. Progress in Physical Geography, 22(2), 251268.Google Scholar
Everitt, B., 1993. Channel responses to declining flow on the Rio Grande between Ft. Quitman and Presidio, Texas. Geomorphology, 6(3), 225242.Google Scholar
Fagan, S. D., Nanson, G. C., 2004. The morphology and formation of floodplain-surface channels, Cooper Creek, Australia. Geomorphology, 60(1–2), 107126.Google Scholar
Fahnestock, R. K., 1963. Morphology and hydrology of a glacial stream – White River, Mt. Rainier, Washington. U.S. Geological Survey Professional Paper 422-A. U.S. Government Printing Office, Washington, DC.Google Scholar
Farnsworth, K. L., Milliman, J. D., 2003. Effects of climatic and anthropogenic change on small mountainous rivers: The Salinas River example. Global and Planetary Change, 39(1–2), 5364.Google Scholar
Fathel, S., Furbish, D., Schmeeckle, M., 2016. Parsing anomalous versus normal diffusive behavior of bedload sediment particles. Earth Surface Processes and Landforms, 41(12), 17971803.Google Scholar
Faulkner, D. J., 1998. Spatially variable historical alluviation and channel incision in west-central Wisconsin. Annals of the Association of American Geographers, 88(4), 666685.Google Scholar
Faulkner, D. J., Larson, P. H., Jol, H. M., et al., 2016. Autogenic incision and terrace formation resulting from abrupt late-glacial base-level fall, lower Chippewa River, Wisconsin, USA. Geomorphology, 266, 7595.Google Scholar
Federici, B., Paola, C., 2003. Dynamics of channel bifurcations in noncohesive sediments. Water Resources Research, 39(6). 10.1029/2002wr001434.Google Scholar
Federici, B., Seminara, R., 2003. On the convective nature of bar instability. Journal of Fluid Mechanics, 487, 125145.Google Scholar
Fencl, J. S., Mather, M. E., Costigan, K. H., Daniels, M. D., 2015. How big of an effect do small dams have? Using geomorphological footprints to quantify spatial impact of low-head dams and identify patterns of across-dam variation. PLoS ONE, 10(11), 22. 10.1371/journal.pone.0141210.Google Scholar
Ferdowsi, B., Ortiz, C. P., Houssais, M., Jerolmack, D. J., 2017. River-bed armouring as a granular segregation phenomenon. Nature Communications, 8. 10.1038/s41467-017–01681-3.Google Scholar
Ferguson, R. I., 1973. Sinuosity of supraglacial streams. Geological Society of America Bulletin, 84(1), 251255.Google Scholar
Ferguson, R. I., 1981. Channel form and channel changes. In: Lewin, J. (ed.), British Rivers. Allen and Unwin, London, pp. 90125.Google Scholar
Ferguson, R. I., 1986a. River loads underestimated by rating curves. Water Resources Research, 22(1), 7476.Google Scholar
Ferguson, R. I., 1986b. Hydraulics and hydraulic geometry. Progress in Physical Geography, 10(1), 131.Google Scholar
Ferguson, R. I., 1987a. Accuracy and precision of methods for estimating river loads. Earth Surface Processes and Landforms, 12(1), 95104.Google Scholar
Ferguson, R., 1987b. Hydraulic and sedimentary controls of channel pattern. In: Richards, K. (ed.), River Channels: Environment and Process. Blackwell, New York, pp. 129158.Google Scholar
Ferguson, R. I., 1993. Understanding braiding processes in gravel-bed rivers: Progress and unsolved problems. In: Best, J. L., Bristow, C. S. (eds.), Braided Rivers. Geological Society of London Special Publication No. 75, Geological Society, London, pp. 7387.Google Scholar
Ferguson, R. I., 1994. Critical discharge for entrainment of poorly sorted gravel. Earth Surface Processes and Landforms, 19(2), 179186.Google Scholar
Ferguson, R. I., 2003. Emergence of abrupt gravel to sand transitions along rivers through sorting processes. Geology, 31(2), 159162.Google Scholar
Ferguson, R. I., 2005. Estimating critical stream power for bedload transport calculations in gravel-bed rivers. Geomorphology, 70(1–2), 3341.Google Scholar
Ferguson, R., 2007. Flow resistance equations for gravel- and boulder-bed streams. Water Resources Research, 43(5). 10.1029/2006WR005422.Google Scholar
Ferguson, R. I., 2012. River channel slope, flow resistance, and gravel entrainment thresholds. Water Resources Research, 48. 10.1029/2011wr010850.Google Scholar
Ferguson, R., Ashworth, P., 1991. Slope-induced changes in channel character along a gravel-bed stream – the Allt Dubhaig, Scotland. Earth Surface Processes and Landforms, 16(1), 6582.Google Scholar
Ferguson, R. J., Brierley, G. J., 1999. Levee morphology and sedimentology along the lower Tuross River, south-eastern Australia. Sedimentology, 46(4), 627648.Google Scholar
Ferguson, R. I., Church, M., 2004. A simple universal equation for grain settling velocity. Journal of Sedimentary Research, 74(6), 933937. 10.1306/051204740933.Google Scholar
Ferguson, R. I., Hoey, T. B., 2002. Long-term slowdown of river tracer pebbles: Generic models and implications for interpreting short-term tracer studies. Water Resources Research, 38(8). 10.1029/2001WR000637.Google Scholar
Ferguson, R. I., Wathen, S. J., 1998. Tracer-pebble movement along a concave river profile: Virtual velocity in relation to grain size and shear stress. Water Resources Research, 34(8), 20312038.Google Scholar
Ferguson, R., Werritty, A., 1983. Bar development and channel changes in the gravelly River Feshie, Scotland. In: Collinson, J. D., Lewin, J. (eds.), Modern and Ancient Fluvial Systems. Special Publication of the International Association of Sedimentologists No. 6, Blackwell, Oxford, UK, pp.181193.Google Scholar
Ferguson, R. I., Ashmore, P. E., Ashworth, P. J., Paola, C., Prestegaard, K. L., 1992. Measurements in a braided river chute and lobe. 1. Flow pattern, sediment transport, and channel change. Water Resources Research, 28(7), 18771886. 10.1029/92wr00700.Google Scholar
Ferguson, R. I., Kirkbride, A. D., Roy, A. G., 1996a. Markov analysis of velocity fluctuations in gravel-bed rivers. In: Ashworth, P. J., Bennett, S. J., Best, J., McLelland, S. (eds.), Coherent Flow Structures in Open Channels. John Wiley, New York, pp. 165183.Google Scholar
Ferguson, R., Hoey, T., Wathen, S., Werritty, A., 1996b. Field evidence for rapid downstream fining of river gravels through selective transport. Geology, 24(2), 179182.Google Scholar
Ferguson, R. I., Bloomer, D. J., Hoey, T. B., Werritty, A., 2002. Mobility of river tracer pebbles over different timescales. Water Resources Research, 38(5). 10.1029/2001wr000254.Google Scholar
Ferguson, R. I., Parsons, D. R., Lane, S. N., Hardy, R. J., 2003. Flow in meander bends with recirculation at the inner bank. Water Resources Research, 39(11). 10.1029/2003wr001965.Google Scholar
Ferguson, R. I., Cudden, J. R., Hoey, T. B., Rice, S. P., 2006. River system discontinuities due to lateral inputs: generic styles and controls. Earth Surface Processes and Landforms, 31(9), 11491166.Google Scholar
Ferguson, R. I., Bloomer, D. J., Church, M., 2011. Evolution of an advancing gravel front: Observations from Vedder Canal, British Columbia. Earth Surface Processes and Landforms, 36(9), 11721182.Google Scholar
Ferguson, R. I., Church, M., Rennie, C. D., Venditti, J. G., 2015. Reconstructing a sediment pulse: Modeling the effect of placer mining on Fraser River, Canada. Journal of Geophysical Research – Earth Surface, 120(7), 14361454. 10.1002/2015jf003491.Google Scholar
Ferguson, R. I., Sharma, B. P., Hodge, R. A., Hardy, R. J., Warburton, J., 2017. Bed load tracer mobility in a mixed bedrock/alluvial channel. Journal of Geophysical Research – Earth Surface, 122(4), 807822.Google Scholar
Fernandez Luque, R., Van Beek, R., 1976. Erosion and transport of bed sediment. Journal of Hydraulic Research, 14, 127144.Google Scholar
Ferreira, R. M. L., 2015. The von Karman constant for flows over rough mobile beds. Lessons learned from dimensional analysis and similarity. Advances in Water Resources, 81, 1932.Google Scholar
Ferrer-Boix, C., Chartrand, S. M., Hassan, M. A., Martin-Vide, J. P., Parker, G., 2016. On how spatial variations of channel width influence river profile curvature. Geophysical Research Letters, 43(12), 63136323. 10.1002/2016gl069824.Google Scholar
Ferro, V., Porto, P., 2012. Identifying a dominant discharge for natural rivers in southern Italy. Geomorphology, 139, 313321.Google Scholar
Filgueira-Rivera, M., Smith, N. D., Slingerland, R. L., 2007. Controls on natural levee development in the Columbia river, British Columbia, Canada. Sedimentology, 54(4), 905919.Google Scholar
Finnegan, N. J., Roe, G., Montgomery, D. R., Hallet, B., 2005. Controls on the channel width of rivers: Implications for modeling fluvial incision of bedrock. Geology, 33(3), 229232.Google Scholar
Fischer, H. B., Imberger, J., List, E. J., Koh, R. C. Y., Brooks, N. H., 1979. Mixing in Inland and Coastal Waters. Academic Press, San Diego, CA.Google Scholar
Fisk, H. N., 1947. Fine-grained alluvial deposits and their effects on Mississippi River activity. Mississippi River Commission, U.S. Waterways Experiment Station, Vols. 1 & 2.Google Scholar
FitzHugh, T. W., Vogel, R. M., 2011. The impact of dams on flood flows in the United States. River Research and Applications, 27(10), 11921215.Google Scholar
Flint, J. J., 1973. Experimental development of headward growth of channel networks. Geological Society of America Bulletin, 84(3), 10871093.Google Scholar
Flint, J. J., 1980. Tributary arrangements in fluvial systems. American Journal of Science, 280(1), 2645.Google Scholar
Flores-Cervantes, J. H., Istanbulluoglu, E., Bras, R. L., 2006. Development of gullies on the landscape: a model of headcut retreat resulting from plunge pool erosion. Journal of Geophysical Research – Earth Surface, 111(F1). 10.1029/2004jf000226.Google Scholar
Florsheim, J. L., Mount, J. F., 2002. Restoration of floodplain topography by sand-splay complex formation in response to intentional levee breaches, Lower Cosumnes River, California. Geomorphology, 44(1–2), 6794.Google Scholar
Florsheim, J. L., Mount, J. F., Chin, A., 2008. Bank erosion as a desirable attribute of rivers. Bioscience, 58(6), 519529.Google Scholar
Flynn, K. M., Kirby, W. H., Hummel, P. R., 2006. User’s Manual for PeakFQ Flood Frequency Analysis for Program Peak FQ Flood Frequency Analysis using Bulletin 17B Guidelines. U.S. Geological Survey, Techniques and Methods Book4, Chapter B4, Reston, VA.Google Scholar
Fola, M. E., Rennie, C. D., 2010. Downstream hydraulic geometry of clay-dominated cohesive bed rivers. Journal of Hydraulic Engineering, 136(8), 524527.Google Scholar
Foley, M. M., Magilligan, F. J., Torgersen, C. E., et al., 2017a. Landscape context and the biophysical response of rivers to dam removal in the United States. PLoS ONE, 12(7). 10.1371/journal.pone.0180107.Google Scholar
Foley, M. M., Bellmore, J. R., O’Connor, J. E., et al., 2017b. Dam removal: Listening in. Water Resources Research, 53(7), 52295246. 10.1002/2017wr020457.Google Scholar
Folk, R. L., 1980. Petrology of Sedimentary Rocks. Hemphill Publishing Co., Austin, TX.Google Scholar
Folk, R. L., Ward, W. C., 1957. Brazos River bar: A study in the significance of grain size parameters. Journal of Sedimentary Petrology, 27, 326.Google Scholar
Fonstad, M. A., Marcus, W. A., 2010. High resolution, basin extent observations and implications for understanding river form and process. Earth Surface Processes and Landforms, 35(6), 680698.Google Scholar
Forte, A. M., Whipple, K. X., 2018. Criteria and tools for determining drainage divide stability. Earth and Planetary Science Letters, 493, 102117.Google Scholar
Foufoula-Georgiou, E., Sapozhnikov, V. B., 1998. Anisotropic scaling in braided rivers: An integrated theoretical framework and results from application to an experimental river. Water Resources Research, 34(4), 863867.Google Scholar
Foufoula-Georgiou, E., Sapozhnikov, V., 2001. Scale invariances in the morphology and evolution of braided rivers. Mathematical Geology, 33(3), 273291.Google Scholar
Fournier, F., 1960. Climate et Erosion. Presses Universitaires de France, Paris.Google Scholar
Fox, C. A., Magilligan, F. J., Sneddon, C. S., 2016. “You kill the dam, you are killing a part of me”: Dam removal and the environmental politics of river restoration. Geoforum, 70, 93104.Google Scholar
Fox, G. A., Felice, R. G., 2014. Bank undercutting and tension failure by groundwater seepage: predicting failure mechanisms. Earth Surface Processes and Landforms, 39(6), 758765.Google Scholar
Fox, G. A., Wilson, G. V., Simon, A., et al., 2007. Measuring streambank erosion due to ground water seepage: Correlation to bank pore water pressure, precipitation and stream stage. Earth Surface Processes and Landforms, 32(10), 15581573.Google Scholar
Fox, J. F., Papanicolaou, A. N., 2008. An un-mixing model to study watershed erosion processes. Advances in Water Resources, 31(1), 96108.Google Scholar
Franca, M. J., Lemmin, U., 2015. Detection and reconstruction of large-scale coherent flow structures in gravel-bed rivers. Earth Surface Processes and Landforms, 40(1), 93104.Google Scholar
Francalanci, S., Solari, L., Toffolon, M., Parker, G., 2012. Do alternate bars affect sediment transport and flow resistance in gravel-bed rivers? Earth Surface Processes and Landforms, 37(8), 866875.Google Scholar
Frankel, K. L., Pazzaglia, F. J., Vaughn, J. D., 2007. Knickpoint evolution in a vertically bedded substrate, upstream-dipping terraces, and Atlantic slope bedrock channels. Geological Society of America Bulletin, 119(3–4), 476486.Google Scholar
Frascati, A., Lanzoni, S., 2010. Long-term river meandering as a part of chaotic dynamics? A contribution from mathematical modelling. Earth Surface Processes and Landforms, 35(7), 791802.Google Scholar
Fredsoe, J., 1978. Meandering and braiding of rivers. Journal of Fluid Mechanics, 84(FEB), 609624.Google Scholar
Frias, C. E., Abad, J. D., Mendoza, A., et al., 2015. Planform evolution of two anabranching structures in the Upper Peruvian Amazon River. Water Resources Research, 51(4), 27422759.Google Scholar
Friedkin, J. F., 1945. A laboratory study of the meandering of alluvial rivers. U.S. Waterways Experiment Station, Vicksburg, MS.Google Scholar
Frings, R. M., 2008. Downstream fining in large sand-bed rivers. Earth-Science Reviews, 87(1–2), 3960.Google Scholar
Frings, R. M., Ottevanger, W., Sloff, K., 2011. Downstream fining processes in sandy lowland rivers. Journal of Hydraulic Research, 49(2), 178193.Google Scholar
Frothingham, K. M., Rhoads, B. L., 2003. Three-dimensional flow structure and channel change in an asymmetrical compound meander loop, Embarras River, Illinois. Earth Surface Processes and Landforms, 28(6), 625644.Google Scholar
Frothingham, K. M., Rhoads, B. L., Herricks, E. E., 2001. Stream geomorphology and fish community structure in channelized and meandering reaches of an agricultural stream. In: Dorava, J. M., Montgomery, D. R., Palcsak, B. B., Fitzpatrick, F. A. (eds.), Geomorphic Processes and Riverine Habitat, American Geophysical Union, Washington, DC, pp. 105117.Google Scholar
Frothingham, K. M., Rhoads, B. L., Herricks, E. E., 2002. A multiscale conceptual framework for integrated ecogeomorphological research to support stream naturalization in the agricultural Midwest. Environmental Management, 29(1), 1633.Google Scholar
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.Google Scholar
Fryirs, K. A., 2017. River sensitivity: A lost foundation concept in fluvial geomorphology. Earth Surface Processes and Landforms, 42(1), 5570.Google 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.Google Scholar
Fryirs, K. A., Brierley, G. J., 2018. What’s in a name? A naming convention for geomorphic river types using the River Styles Framework. PLoS ONE, 13(9). 10.1371/journal.pone.0201909.Google Scholar
Fryirs, K. A., Brierley, G. J., Preston, N. J., Kasai, M., 2007. Buffers, barriers and blankets: The (dis)connectivity of catchment-scale sediment cascades. Catena, 70(1), 4967.Google Scholar
Fryirs, K., Spink, A., Brierley, G., 2009. Post-European settlement response gradients of river sensitivity and recovery across the upper Hunter catchment, Australia. Earth Surface Processes and Landforms, 34(7), 897918.Google Scholar
Fryirs, K., Brierley, G. J., Erskine, W. D., 2012. Use of ergodic reasoning to reconstruct the historical range of variability and evolutionary trajectory of rivers. Earth Surface Processes and Landforms, 37(7), 763773.Google Scholar
Fryirs, K., Lisenby, P., Croke, J., 2015. Morphological and historical resilience to catastrophic flooding: The case of Lockyer Creek, SE Queensland, Australia. Geomorphology, 241, 5571.Google Scholar
Fryirs, K. A., Wheaton, J. M., Brierley, G. J., 2016. An approach for measuring confinement and assessing the influence of valley setting on river forms and processes. Earth Surface Processes and Landforms, 41(5), 701710.Google Scholar
Fujita, Y., 1989. Bar and channel formation in braided streams. In: Ikeda, S., Parker, G. (eds.), River Meandering. American Geophysical Union, Washington, DC, pp. 417462.Google Scholar
Fuller, I. C., Large, A. R. G., Milan, D. J., 2003. Quantifying channel development and sediment transfer following chute cutoff in a wandering gravel-bed river. Geomorphology, 54(3–4), 307323.Google Scholar
Furbish, D. J., 1991. Spatial autoregressive structure in meander evolution. Geological Society of America Bulletin, 103(12), 15761589.Google Scholar
Furbish, D. J., 1997. Fluid Physics in Geology. Oxford University Press, New York.Google Scholar
Furbish, D. J., Haff, P. K., Roseberry, J. C., Schmeeckle, M. W., 2012. A probabilistic description of the bed load sediment flux: 1. theory. Journal of Geophysical Research – Earth Surface, 117. 10.1029/2012jf002352.Google Scholar
Furbish, D. J., Schmeeckle, M., Fathel, S., 2017. Particle motions and bedload theory: The entrainment forms of the flux and the Exner equation. In: Tsutsumi, D., Laronne, J. B. (eds.), Gravel-Bed Rivers: Processes and Disasters. Wiley, Chichester, UK, pp. 97120.Google Scholar
Furey, P. R., Troutman, B. M., 2008. A consistent framework for Horton regression statistics that leads to a modified Hack’s law. Geomorphology, 102(3–4), 603614.Google Scholar
Gagliano, S. M., Howard, P. C., 1984. The neck cutoff oxbow lake cycle along the lower Mississippi River. In: Elliott, C. M. (ed.), River Meandering. American Society of Civil Engineers, New York, pp. 147158.Google Scholar
Galay, V. J., 1983. Causes of river bed degradation. Water Resources Research, 19(5), 10571090.Google Scholar
Gale, S. J., Hoare, P. G., 1992. Bulk sampling of coarse clastic sediments for particle size analysis. Earth Surface Processes and Landforms, 17(7), 729733.Google Scholar
Galster, J. C., Pazzaglia, F. J., Germanoski, D., 2008. Measuring the impact of urbanization on channel widths using historic aerial photographs and modern surveys. Journal of the American Water Resources Association, 44(4), 948960.Google Scholar
Ganti, V., Meerschaert, M. M., Foufoula-Georgiou, E., Viparelli, E., Parker, G., 2010. Normal and anomalous diffusion of gravel tracer particles in rivers. Journal of Geophysical Research – Earth Surface, 115. 10.1029/2008jf001222.Google Scholar
Gao, P., 2008. Understanding watershed suspended sediment transport. Progress in Physical Geography, 32(3), 243263.Google Scholar
Gao, P., 2013. Rill and gully development processes. In: Schroder, J. W. (ed.), Treatise on Geomorphology, Vol. 7, Hillslope Processes, Stoffel, M., Marston, R. (vol. eds.). Wiley, New York, pp. 122131.Google Scholar
Gao, P., Josefson, M., 2012a. Temporal variations of suspended sediment transport in Oneida Creek watershed, central New York. Journal of Hydrology, 426, 1727.Google Scholar
Gao, P., Josefson, M., 2012b. Event-based suspended sediment dynamics in a central New York watershed. Geomorphology, 139, 425437.Google Scholar
Gao, P., Nearing, M. A., Commons, M., 2013. Suspended sediment transport at the instantaneous and event time scales in semiarid watersheds of southeastern Arizona, USA. Water Resources Research, 49(10), 68576870. 10.1002/wrcr.20549.Google Scholar
Garcia, M. H., 2008. Sediment transport and morphodynamics. In: Garcia, M. H. (ed.), Sedimentation Engineering: Processes, Measurements, Modeling, and Practice. American Society of Civil Engineers, Reston, VA, pp. 21163.Google Scholar
Garcia, M., Nino, Y., 1993. Dynamics of sediment bars in straight and meandering channels – experiments on the resonance phenomenon. Journal of Hydraulic Research, 31(6), 739761.Google Scholar
Garcia, M., Parker, G., 1991. Entrainment of bed sediment into suspension. Journal of Hydraulic Engineering, 117(4), 414435.Google Scholar
Garcia-Flores, M., Maza-Alvarez, J. A., 1997. Inicio de movimiento y acorazamiento, Capitulo 8 del Manual de Ingeniería de Ríos, Series del Instituto de Ingeniería 592, UNAM, Mexico (in Spanish).Google Scholar
Gardner, T. W., 1983. Experimental study of knickpoint and longitudinal profile evolution in cohesive, homogenous material. Geological Society of America Bulletin, 94(5), 664672.Google Scholar
Gardner, T., Ashmore, P., Leduc, P., 2018. Morpho-sedimentary characteristics of proximal gravel braided river deposits in a Froude-scaled physical model. Sedimentology, 65(3), 877896.Google Scholar
Garrels, R. M., MacKenzie, F. T., 1971. Evolution of Sedimentary Rocks. Norton, New York.Google Scholar
Gasparini, N. M., Tucker, G. E., Bras, R. L., 1999. Downstream fining through selective particle sorting in an equilibrium drainage network. Geology, 27(12), 10791082.Google Scholar
Gaudet, J. M., Roy, A. G., 1995. Effect of bed morphology on flow mixing length at river confluences. Nature, 373(6510), 138139.Google Scholar
Gaudio, R., Miglio, A., Dey, S., 2010. Non-universality of von Karman’s κ in fluvial streams. Journal of Hydraulic Research, 48(5), 658663.Google Scholar
Gaurav, K., Tandon, S. K., Devauchelle, O., Sinha, R., Metivier, F., 2017. A single width-discharge regime relationship for individual threads of braided and meandering rivers from the Himalayan Foreland. Geomorphology, 295, 126133.Google Scholar
Gay, G. R., Gay, H. H., Gay, W. H., et al., 1998. Evolution of cutoffs across meander necks in Powder River, Montana, USA. Earth Surface Processes and Landforms, 23(7), 651662.Google Scholar
Gee, G. W., Bauder, J. W., 1986. Particle-size analysis. In: Klute, A. (ed.), Methods of Soil Analysis. Part 1, 2nd ed., Agronomy Monograph 9. Soil Science Society of America and American Society of Agronomy, Madison, WI, pp. 383411.Google Scholar
Gellis, A. C., 2013. Factors influencing storm-generated suspended-sediment concentrations and loads in four basins of contrasting land use, humid-tropical Puerto Rico. Catena, 104, 3957.Google Scholar
Gellis, A. C., Myers, M. K., Noe, G. B., et al., 2017. Storms, channel changes, and a sediment budget for an urban-suburban stream, Difficult Run, Virginia, USA. Geomorphology, 278, 128148.Google Scholar
Germanoski, D., Schumm, S. A., 1993. Changes in braided river morphology resulting from aggradation and degradation. Journal of Geology, 101(4), 451466.Google Scholar
Ghinassi, M., Ielpi, A., Aldinucci, M., Fustic, M., 2016. Downstream-migrating fluvial point bars in the rock record. Sedimentary Geology, 334, 6696.Google Scholar
Ghoshal, S., James, L. A., Singer, M. B., Aalto, R., 2010. Channel and floodplain change analysis over a 100-year period: Lower Yuba River, California. Remote Sensing, 2(7), 17971825.Google Scholar
Giachetta, E., Refice, A., Capolongo, D., Gasparini, N. M., Pazzaglia, F. J., 2014. Orogen-scale drainage network evolution and response to erodibility changes: insights from numerical experiments. Earth Surface Processes and Landforms, 39(9), 12591268.Google Scholar
Giere, R. N., 1988. Explaining Science: A Cognitive Approach. University of Chicago Press, Chicago.Google Scholar
Gilbert, G. K., 1877. Report on the geology of the Henry Mountains. U.S. Geographical and Geological Survey of the Rocky Mountains Region. U.S. Government Printing Office, Washington, DC.Google Scholar
Gilbert, G. K., 1907. Rate of Recession of Niagara Falls, U.S. Geological Survey Bulletin No. 36. U.S. Government Printing Office, Washington, DC.Google Scholar
Gilbert, G. K., 1914. The transportation of debris by running water. U.S. Geological Survey Professional Paper 86. U.S. Government Printing Office, Washington, DC.Google Scholar
Gilbert, G. K., 1917. Hydraulic-mining debris in the Sierra Nevada. U.S. Geological Survey Professional Paper 105, U.S. Government Printing Office, Washington, DC.Google Scholar
Gilvear, D., Bryant, R., 2016. Analysis of remotely sensed data for fluvial geomorphology and river science. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology, 2nd Edition. Wiley, Chichester, UK, pp. 103132.Google Scholar
Gleason, C. J., Smith, L. C., 2014. Toward global mapping of river discharge using satellite images and at-many-stations hydraulic geometry. Proceedings of the National Academy of Sciences of the United States of America, 111(13), 47884791.Google Scholar
Gleason, C. J., Wang, J., 2015. Theoretical basis for at-many-stations hydraulic geometry. Geophysical Research Letters, 42(17), 71077114. 10.1002/2015gl064935.Google Scholar
Gleason, C. J., Smith, L. C., Lee, J., 2014. Retrieval of river discharge solely from satellite imagery and at-many-stations hydraulic geometry: Sensitivity to river form and optimization parameters. Water Resources Research, 50(12), 96049619. 10.1002/2014wr016109.Google Scholar
Glock, W. S., 1931. The development of drainage systems: A synoptic view. Geographical Review, 21(3), 475482. 10.2307/209434.Google Scholar
Glymph, L. M., 1954. Studies of sediment yields from watersheds. IAHS Publication 36, IAHS Press, Wallingford, UK, 178191.Google Scholar
Goff, J. R., Ashmore, P., 1994. Gravel transport and morphological change in braided Sunwapta River, Alberta, Canada. Earth Surface Processes and Landforms, 19(3), 195212.Google Scholar
Goldrick, G., Bishop, P., 2007. Regional analysis of bedrock stream long profiles: evaluation of Hack’s SL form, and formulation and assessment of an alternative (the DS form). Earth Surface Processes and Landforms, 32(5), 649671.Google Scholar
Gomez, B., 1983. Temporal variations in the particle-size distribution of surficial bed material – the effect of progressive bed armoring. Geografiska Annaler Series A Physical Geography, 65(3–4), 183192.Google Scholar
Gomez, B., 1984. Typology of segregated (armored paved) surfaces – some comments. Earth Surface Processes and Landforms, 9(1), 1924.Google Scholar
Gomez, B., 1995. Bedload transport and changing grain size distributions. In: Gurnell, A., Petts, G. (eds.), Changing River Channels. Wiley, Chichester, UK, pp. 177199.Google Scholar
Gomez, B., Church, M., 1989. An assessment of bedload sediment transport formulas for gravel bed rivers. Water Resources Research, 25(6), 11611186.Google Scholar
Gomez, B., Mullen, V. T., 1992. An experimental study of sapped drainage network development. Earth Surface Processes and Landforms, 17(5), 465476.Google Scholar
Gomez, B., Mertes, L. A. K., Phillips, J. D., Magilligan, F. J., James, L. A., 1995. Sediment characteristics of an extreme flood – 1993 upper Mississippi River Valley. Geology, 23(11), 963966.Google Scholar
Gomez, B., Phillips, J. D., Magilligan, F. J., James, L. A., 1997. Floodplain sedimentation and sensitivity: summer 1993 flood, upper Mississippi River valley. Earth Surface Processes and Landforms, 22(10), 923936.Google Scholar
Gomez, B., Eden, D. N., Peacock, D. H., Pinkney, E. J., 1998. Floodplain construction by recent, rapid vertical accretion: Waipaoa River, New Zealand. Earth Surface Processes and Landforms, 23(5), 405413.Google Scholar
Gomez, B., Coleman, S. E., Sy, V. W. K., Peacock, D. H., Kent, M., 2007. Channel change, bankfull and effective discharges on a vertically accreting, meandering, gravel-bed river. Earth Surface Processes and Landforms, 32(5), 770785. 10.1002/esp.1424.Google Scholar
Gomez, J. A., Darboux, F., Nearing, M. A., 2003. Development and evolution of rill networks under simulated rainfall. Water Resources Research, 39(6), 14. 10.1029/2002wr001437.Google Scholar
Gomi, T., Sidle, R. C., Woodsmith, R. D., Bryant, M. D., 2003. Characteristics of channel steps and reach morphology in headwater streams, southeast Alaska. Geomorphology, 51(1–3), 225242.Google Scholar
Gonzalez, E., Sher, A. A., Tabacchi, E., Masip, A., Poulin, M., 2015. Restoration of riparian vegetation: A global review of implementation and evaluation approaches in the international, peer-reviewed literature. Journal of Environmental Management, 158, 8594.Google Scholar
Gonzalez, R. L., Pasternack, G. B., 2015. Reenvisioning cross-sectional at-a-station hydraulic geometry as spatially explicit hydraulic topography. Geomorphology, 246, 394406.Google Scholar
Goode, J. R., Wohl, E., 2007. Relationships between land-use and forced-pool characteristics in the Colorado Front Range. Geomorphology, 83(3–4), 249265.Google Scholar
Goodwell, A. E., Zhu, Z. D., Dutta, D., et al., 2014. Assessment of floodplain vulnerability during extreme Mississippi River flood 2011. Environmental Science & Technology, 48(5), 26192625.Google Scholar
Goodwin, P., 2004. Analytical solutions for estimating effective discharge. Journal of Hydraulic Engineering-ASCE, 130(8), 729738.Google Scholar
Gorrick, S., Rodriguez, J. F., 2014. Flow and force-balance relations in a natural channel with bank vegetation. Journal of Hydraulic Research, 52(6), 794810.Google Scholar
Gorycki, M. A., 1973. Hydraulic drag – a meander-initiating mechanism. Geological Society of America Bulletin, 84(1), 175186.Google Scholar
Grabowski, R. C., Gurnell, A. M., 2016. Using historical data in fluvial geomorphology. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology, 2nd Edition. Wiley, Chichester, UK, pp. 5675.Google Scholar
Graf, W. H., Cellino, M., 2002. Suspension flows in open channels; experimental study. Journal of Hydraulic Research, 40(4), 435447.Google Scholar
Graf, W. L., 1975. Impact of suburbanization on fluvial geomorphology. Water Resources Research, 11(5), 690692.Google Scholar
Graf, W. L., 1977a. Rate law in fluvial geomorphology. American Journal of Science, 277(2), 178191.Google Scholar
Graf, W. L., 1977b. Network characteristics in suburbanizing streams. Water Resources Research, 13(2), 459463.Google Scholar
Graf, W. L., 1983a. Downstream changes in stream power in the Henry Mountains, Utah. Annals of the Association of American Geographers, 73(3), 373387.Google Scholar
Graf, W. L., 1983b. Flood-related change in an arid-region river. Earth Surface Processes and Landforms, 8(2), 125139.Google Scholar
Graf, W. L., 1983c. The arroyo problem: Paleohydrology and paleohydraulics in the short term. In: Gregory, K. (ed.), Background to Paleohydrology. Wiley, New York, pp. 279302.Google Scholar
Graf, W. L., 1988a. Fluvial Processes in Dryland Rivers. Springer-Verlag, Berlin.Google Scholar
Graf, W. L., 1988b. Definition of flood plains along arid-region rivers. In: Baker, V. R., Kochel, R. C., Patton, P. C. (eds.), Flood Geomorphology. Wiley, New York, pp. 231242.Google Scholar
Graf, W. L., 1990. Fluvial dynamics of Thorium-230 in the Church Rock Event, Puerco River, New Mexico. Annals of the Association of American Geographers, 80(3), 327342.Google Scholar
Graf, W. L., 1993. Landscapes, Commodities, and Ecosystems: The Relationship between Science and Policy for American Rivers, Sustaining Our Water Resources. National Academy Press, Washington, DC, pp. 1142.Google Scholar
Graf, W. L., 1999. Dam nation: A geographic census of American dams and their large-scale hydrologic impacts. Water Resources Research, 35(4), 13051311.Google Scholar
Graf, W. L., 2006. Downstream hydrologic and geomorphic effects of large dams on American rivers. Geomorphology, 79(3–4), 336360.Google Scholar
Graf, W. L., 2008. Sources of uncertainty in river restoration research. In: Darby, S. E., Sear, D. A. (eds.), River Restoration: Managing the Uncertainty in Restoring Physical Habitat. Wiley, Chichester, UK, pp. 1519.Google Scholar
Grams, P. E., Schmidt, J. C., 2002. Streamflow regulation and multi-level flood plain formation: Channel narrowing on the aggrading Green River in the eastern Uinta Mountains, Colorado and Utah. Geomorphology, 44(3–4), 337360.Google Scholar
Grams, P. E., Schmidt, J. C., 2005. Equilibrium or indeterminate? Where sediment budgets fail: Sediment mass balance and adjustment of channel form, Green River downstream from Flaming Gorge Dam, Utah and Colorado. Geomorphology, 71(1–2), 156181.Google Scholar
Grams, P. E., Schmidt, J. C., Topping, D. J., 2007. The rate and pattern of bed incision and bank adjustment on the Colorado River in Glen Canyon downstream from Glen Canyon Dam, 1956–2000. Geological Society of America Bulletin, 119(5–6), 556575.Google Scholar
Gran, K. B., Czuba, J. A., 2017. Sediment pulse evolution and the role of network structure. Geomorphology, 277, 1730.Google Scholar
Gran, K. B., Paola, C., 2001. Riparian vegetation controls on braided stream dynamics. Water Resources Research, 37(12), 32753283.Google Scholar
Gran, K. B., Tal, M., Wartman, E. D., 2015. Co-evolution of riparian vegetation and channel dynamics in an aggrading braided river system, Mount Pinatubo, Philippines. Earth Surface Processes and Landforms, 40(8), 11011115. 10.1002/esp.3699.Google Scholar
Granger, D. E., Schaller, M., 2014. Cosmogenic nuclides and erosion at the watershed scale. Elements, 10(5), 369373.Google Scholar
Grant, G. E., Swanson, F. J., Wolman, M. G., 1990. Pattern and origin of stepped-bed morphology in high-gradient streams, western Cascades, Oregon. Geological Society of America Bulletin, 102(3), 340352.Google Scholar
Grant, G. E., Schmidt, J. C., Lewis, S., 2003. A geological framework for interpreting downstream effects of dams on rivers. In: O’Connor, J. E., Grant, G. E. (eds.), A Peculiar River. American Geophysical Union, Washington, DC., pp. 203219.Google Scholar
Grant, G. E., O’Connor, J. E., Wolman, M. G., 2013. A river runs through it: conceptual models in fluvial geomorphology. In: Schroder, J. W. (ed.), Treatise on Geomorphology, Vol. 9, Fluvial geomorphology, Wohl, E. (vol. ed.). Academic Press, San Diego, CA, pp. 621.Google Scholar
Gray, A. B., Warrick, J. A., Pasternack, G. B., Watson, E. B., Goni, M. A., 2014. Suspended sediment behavior in a coastal dry-summer subtropical catchment: Effects of hydrologic preconditions. Geomorphology, 214, 485501.Google Scholar
Gray, R. J., Glysson, G. D., Turcios, L. M., Schwartz, G. E., 2000. Comparability of suspended-sediment concentration and total suspended solids data. U.S. Geological Survey Water-Resources Investigation Report 00–4191. U.S. Geological Survey, Reston, VA.Google Scholar
Gregoretti, C., 2008. Inception sediment transport relationships at high slopes. Journal of Hydraulic Engineering, 134(11), 16201629.Google Scholar
Gregory, K. J., 1976. Drainage networks and climate. In: Derbyshire, E. (ed.), Geomorphology and Climate. Wiley and Sons, Chichester, UK, pp. 289315.Google Scholar
Gregory, K. J., 2006. The human role in changing river channels. Geomorphology, 79(3–4), 172191.Google Scholar
Gregory, K. J., Downs, P., 2008. The sustainability of restored rivers: catchment-scale perspectives on long-term response. In: Darby, S. E., Sear, D. A. (eds.), River Restoration: Managing the Uncertainty in Restoring Physical Habitat. Wiley, Chichester, UK, pp. 253286.Google Scholar
Gregory, K. J., Gurnell, A. M., Hill, C. T., Tooth, S., 1994. Stability of the pool riffle sequence in changing river channels. Regulated Rivers-Research & Management, 9(1), 3543.Google Scholar
Grenfell, M., Aalto, R., Nicholas, A., 2012. Chute channel dynamics in large, sand-bed meandering rivers. Earth Surface Processes and Landforms, 37(3), 315331.Google Scholar
Grenfell, M. C., Nicholas, A. P., Aalto, R., 2014. Mediative adjustment of river dynamics: The role of chute channels in tropical sand-bed meandering rivers. Sedimentary Geology, 301, 93106.Google Scholar
Griffiths, G. A., 1984. Extremal hypotheses for river regime – an illusion of progress. Water Resources Research, 20(1), 113118.Google Scholar
Grill, G., Lehner, B., Thieme, M., et al., 2019. Mapping the world’s free-flowing rivers. Nature, 569 (7755), 215221. 10.1038/s41586-019–1111-9.Google Scholar
Grimaldi, S., Petroselli, A., Nardi, F., 2012. A parsimonious geomorphological unit hydrograph for rainfall-runoff modelling in small ungauged basins. Hydrological Sciences Journal, 57(1), 7383. 10.1080/02626667.2011.636045.Google Scholar
Grimaud, J. L., Paola, C., Voller, V., 2016. Experimental migration of knickpoints: Influence of style of base-level fall and bed lithology. Earth Surface Dynamics, 4(1), 1123.Google Scholar
Grimley, D. A., Anders, A. M., Bettis, E. A., III, et al., 2017. Using magnetic fly ash to identify post-settlement alluvium and its record of atmospheric pollution, central USA. Anthropocene, 17, 8498.Google Scholar
Grissinger, E. H., 1996. Rill and gullies erosion. In: Agassi, M. (ed.), Soil Erosion, Conservation and Rehabilitation. Marcel Dekker, New York, pp. 153167.Google Scholar
Grissinger, E. H., Murphey, J. B., 1984. Morphometric evolution of man-modified channels. In: Elliott, C. M. (ed.), River Meandering. American Society of Civil Engineers, New York, pp. 273283.Google Scholar
Grudzinski, B. P., Daniels, M. D., 2018. Bison and cattle grazing impacts on grassland stream morphology in the Flint Hills of Kansas. Rangeland Ecology & Management, 71(6), 783791.Google Scholar
Grudzinski, B. P., Daniels, M. D., Anibas, K., Spencer, D., 2016. Bison and cattle grazing management, bare ground coverage, and links to suspended sediment concentrations in grassland streams. Journal of the American Water Resources Association, 52(1), 1630.Google Scholar
Gualtieri, C., Filizola, N., de Oliveira, M., Santos, A. M., Ianniruberto, M., 2018. A field study of the confluence between Negro and Solimoes Rivers. Part 1: Hydrodynamics and sediment transport. Comptes Rendus Geoscience, 350(1–2), 3142.Google Scholar
Guillen-Ludena, S., Franca, M. J., Cardoso, A. H., Schleiss, A. J., 2015. Hydro-morphodynamic evolution in a 90 degrees movable bed discordant confluence with low discharge ratio. Earth Surface Processes and Landforms, 40(14), 19271938.Google Scholar
Guillen-Ludena, S., Franca, M. J., Cardoso, A. H., Schleiss, A. J., 2016. Evolution of the hydromorphodynamics of mountain river confluences for varying discharge ratios and junction angles. Geomorphology, 255, 115.Google Scholar
Guillen-Ludena, S., Franca, M. J., Alegria, F., Schleiss, A. J., Cardoso, A. H., 2017a. Hydromorphodynamic effects of the width ratio and local tributary widening on discordant confluences. Geomorphology, 293, 289304.Google Scholar
Guillen-Ludena, S., Cheng, Z., Constantinescu, G., Franca, M. J., 2017b. Hydrodynamics of mountain-river confluences and its relationship to sediment transport. Journal of Geophysical Research – Earth Surface, 122(4), 901924. 10.1002/2016jf004122.Google Scholar
Guinoiseau, D., Bouchez, J., Gelabert, A., et al., 2016. The geochemical filter of large river confluences. Chemical Geology, 441, 191203.Google Scholar
Guneralp, I., Marston, R. A., 2012. Process-form linkages in meander morphodynamics: Bridging theoretical modeling and real world complexity. Progress in Physical Geography, 36(6), 718746.Google Scholar
Guneralp, I., Rhoads, B. L., 2008. Continuous characterization of the planform geometry and curvature of meandering rivers. Geographical Analysis, 40(1), 125.Google Scholar
Guneralp, I., Rhoads, B. L., 2009. Empirical analysis of the planform curvature-migration relation of meandering rivers. Water Resources Research, 45. 10.1029/2008wr007533.Google Scholar
Guneralp, I., Rhoads, B. L., 2010. Spatial autoregressive structure of meander evolution revisited. Geomorphology, 120(3–4), 91106.Google Scholar
Guneralp, I., Rhoads, B. L., 2011. Influence of floodplain erosional heterogeneity on planform complexity of meandering rivers. Geophysical Research Letters, 38. 10.1029/2011gl048134.Google Scholar
Guo, J., 2002. Hunter Rouse and Shields diagram advances in hydraulics and water engineering, Proceedings of the 13th IAHR-APD Congress. World Scientific Singapore, pp. 10961098.Google Scholar
Guo, J. K., 2013. Modified log-wake law for smooth rectangular open channel flow. Proceedings of the 35th IAHR World Congress, Vols I and II, 20102019.Google Scholar
Guo, J. K., Julien, P. Y., 2001. Turbulent velocity profiles in sediment-laden flows. Journal of Hydraulic Research, 39(1), 1123.Google Scholar
Guo, J., Julien, P. Y., 2005. Shear stress in smooth rectangular open-channel flows. Journal of Hydraulic Engineering, 131(1), 3037.Google Scholar
Gupta, N., Kleinhans, M. G., Addink, E. A., Atkinson, P. M., Carling, P. A., 2014. One-dimensional modeling of a recent Ganga avulsion: Assessing the potential effect of tectonic subsidence on a large river. Geomorphology, 213, 2437.Google Scholar
Gurnell, A. M., Petts, G. E., Hannah, D. M., et al., 2001. Riparian vegetation and island formation along the gravel-bed Fiume Tagliamento, Italy. Earth Surface Processes and Landforms, 26(1), 3162.Google Scholar
Gurnell, A., Lee, M., Souch, C., 2007. Urban rivers: hydrology, geomorphology, ecology and opportunities for change. Geography Compass, 1, 11181137.Google Scholar
Gurnell, A., Surian, N., Zanoni, L., 2009. Multi-thread river channels: a perspective on changing European alpine river systems. Aquatic Sciences, 71(3), 253265.Google Scholar
Gurram, S. K., Karki, K. S., Hager, W. H., 1997. Subcritical junction flow. Journal of Hydraulic Engineering, 123(5), 447455.Google Scholar
Guy, H. P., 1969. Laboratory theory and methods for sediment analysis, U.S. Geological Survey Techniques of Water Resources Investigations, Book 5, Chapter C1. Reston, VA, pp. 158.Google Scholar
Habersack, H. M., 2001. Radio-tracking gravel particles in a large braided river in New Zealand: A field test of the stochastic theory of bed load transport proposed by Einstein. Hydrological Processes, 15(3), 377391.Google Scholar
Hack, J. T., 1957. Studies of longitudinal profiles in Virginia and Maryland. U.S. Geological Survey Professional Paper 294-B. U.S. Government Printing Office, Washington, DC.Google Scholar
Hack, J. T., 1960. Interpretation of erosional topography in humid temperate regions. American Journal of Science, 258, 8097.Google Scholar
Hackney, C., Best, J., Leyland, J., et al., 2015. Modulation of outer bank erosion by slump blocks: disentangling the protective and destructive role of failed material on the three-dimensional flow structure. Geophysical Research Letters, 42(24), 1066310670. 10.1002/2015gl066481.Google Scholar
Hackney, C. R., Darby, S. E., Parsons, D. R., et al., 2018. The influence of flow discharge variations on the morphodynamics of a diffluence-confluence unit on a large river. Earth Surface Processes and Landforms, 43(2), 349362.Google Scholar
Hadley, R. F., Schumm, S. A., 1961. Sediment sources and drainage basin characteristics in the upper Cheyenne River basin, Wyoming. U.S. Geological Survey Water-Supply Paper 1531-B. U.S. Government Printing Office, Washington, DC.Google Scholar
Hagerty, D. J., 1991. Piping sapping erosion. 1. Basic considerations. Journal of Hydraulic Engineering, 117(8), 9911008.Google Scholar
Halbert, K., Nguyen, C. C., Payrastre, O., Gaume, E., 2016. Reducing uncertainty in flood frequency analyses: a comparison of local and regional approaches involving information on extreme historical floods. Journal of Hydrology, 541, 9098.Google Scholar
Hall, P., 2004. Alternating bar instabilities in unsteady channel flows over erodible beds. Journal of Fluid Mechanics, 499, 4973.Google Scholar
Hammer, T. R., 1972. Stream channel enlargement due to urbanization. Water Resources Research, 8, 15301540.Google Scholar
Han, B., Endreny, T. A., 2014. Detailed river stage mapping and head gradient analysis during meander cutoff in a laboratory river. Water Resources Research, 50(2), 16891703. 10.1002/2013wr013580.Google Scholar
Hancock, G. R., Evans, K. G., 2006. Channel head location and characteristics using digital elevation models. Earth Surface Processes and Landforms, 31(7), 809824.Google Scholar
Hancock, G., Willgoose, G., 2001a. The interaction between hydrology and geomorphology in a landscape simulator experiment. Hydrological Processes, 15(1), 115133.Google Scholar
Hancock, G., Willgoose, G., 2001b. Use of a landscape simulator in the validation of the SIBERIA catchment evolution model: Declining equilibrium landforms. Water Resources Research, 37(7), 19811992.Google Scholar
Hancock, G. R., Willgoose, G. R., 2002. The use of a landscape simulator in the validation of the SIBERIA landscape evolution model: Transient landforms. Earth Surface Processes and Landforms, 27(12), 13211334.Google Scholar
Hancock, G. R., Crawter, D., Fityus, S. G., Chandler, J., Wells, T., 2008. The measurement and modelling of rill erosion at angle of repose slopes in mine spoil. Earth Surface Processes and Landforms, 33(7), 10061020.Google Scholar
Hancock, G. R., Willgoose, G. R., Lowry, J., 2014a. Transient landscapes: gully development and evolution using a landscape evolution model. Stochastic Environmental Research and Risk Assessment, 28(1), 8398.Google Scholar
Hancock, G. J., Wilkinson, S. N., Hawdon, A. A., Keen, R. J., 2014b. Use of fallout tracers 7Be, 210Pb and 137Cs to distinguish the form of sub-surface soil erosion delivering sediment to rivers in large catchments. Hydrological Processes, 28(12), 38553874.Google Scholar
Hanrahan, B. R., Tank, J. L., Dee, M. M., et al., 2018. Restored floodplains enhance denitrification compared to naturalized floodplains in agricultural streams. Biogeochemistry, 141(3), 419437.Google Scholar
Happ, S. C., 1945. Sedimentation in South Carolina Piedmont valleys. American Journal of Science, 243(3), 113126.Google Scholar
Happ, S. C., Rittenhouse, G., Dobson, G. C., 1940. Some principles of accelerated stream and valley sedimentation. U.S. Department of Agriculture Technical Bulletin 695. U.S. Government Printing Office, Washington, DC.Google Scholar
Harden, C. P., 2013. Impacts of vegetation clearance on channel change: Historical perspective. In: Shroder, J. C. (ed.), Treatise on Geomorphology, Vol. 13, Geomorphology of Human Disturbances, Climate Change, and Natural Hazards, James, L. A., Harden, C. P., Clague, J. J. (vol. eds.). Academic Press, San Diego, CA, pp. 1427.Google Scholar
Harden, D. R., 1990. Controlling factors in the distribution and development of incised meanders in the central Colorado Plateau. Geological Society of America Bulletin, 102(2), 233242.Google Scholar
Hardy, R. J., Bates, P. D., Anderson, M. G., 2000. Modelling suspended sediment deposition on a fluvial floodplain using a two-dimensional dynamic finite element model. Journal of Hydrology, 229(3–4), 202218.Google Scholar
Hardy, R. J., Best, J. L., Lane, S. N., Carbonneau, P. E., 2009. Coherent flow structures in a depth-limited flow over a gravel surface: The role of near-bed turbulence and influence of Reynolds number. Journal of Geophysical Research – Earth Surface, 114. 10.1029/2007jf000970.Google Scholar
Hardy, R. J., Best, J. L., Parsons, D. R., Marjoribanks, T. I., 2016. On the evolution and form of coherent flow structures over a gravel bed: Insights from whole flow field visualization and measurement. Journal of Geophysical Research – Earth Surface, 121(8), 14721493. 10.1002/2015jf003753.Google Scholar
Harman, C., Stewardson, M., DeRose, R., 2008. Variability and uncertainty in reach bankfull hydraulic geometry. Journal of Hydrology, 351(1–2), 1325.Google Scholar
Harman, W., Starr, R., 2011. Natural Channel Design Review Checklist. US Fish and Wildlife Service, Chesapeake Bay Field Office, US Environmental Protection Agency, Office of Wetlands, Oceans, and Watersheds, Wetlands Division, Washington, DC.Google Scholar
Harrison, L. R., Keller, E. A., 2007. Modeling forced pool-riffle hydraulics in a boulder-bed stream, southern California. Geomorphology, 83(3–4), 232248.Google Scholar
Harrison, L. R., Dunne, T., Fisher, G. B., 2015. Hydraulic and geomorphic processes in an overbank flood along a meandering, gravel-bed river: Implications for chute formation. Earth Surface Processes and Landforms, 40(9), 12391253.Google Scholar
Harvey, A. M., 2002. Effective timescales of coupling within fluvial systems. Geomorphology, 44(3–4), 175201.Google Scholar
Harwood, K., Brown, A. G., 1993. Fluvial processes in a forested anastomosing river – flood partitioning and change flow patterns. Earth Surface Processes and Landforms, 18(8), 741748.Google Scholar
Haschenburger, J. K., 2011. Vertical mixing of gravel over a long flood series. Earth Surface Processes and Landforms, 36(8), 10441058.Google Scholar
Haschenburger, J. K., 2013a. Bedload kinematics and fluxes. In: Schroder, J. (ed.), Treatise on Geomorphology, Vol. 9, Fluvial Geomorphology, Wohl, E. (vol. ed.) Academic Press, San Diego, CA, pp. 103123.Google Scholar
Haschenburger, J. K., 2013b. Tracing river gravels: insights into dispersion from a long-term field experiment. Geomorphology, 200, 121131.Google Scholar
Haschenburger, J. K., 2017. Streambed disturbance over a long flood series. River Research and Applications, 33(5), 753765.Google Scholar
Haschenburger, J. K., Church, M., 1998. Bed material transport estimated from the virtual velocity of sediment. Earth Surface Processes and Landforms, 23(9), 791808.Google Scholar
Haschenburger, J. K., Cowie, M., 2009. Floodplain stages in the braided Ngaruroro River, New Zealand. Geomorphology, 103(3), 466475.Google Scholar
Hassan, M. A., Bradley, D. N., 2017. Geomorphic controls on tracer particle dispersion in gravel-bed rivers. In: Tsutsumi, D., Laronne, J. B. (eds.), Gravel-Bed Rivers: Processes and Disasters. Wiley, Chichester, UK, pp. 159184.Google Scholar
Hassan, M. A., Church, M., 1994. Vertical mixing of coarse particles in gravel-bed rivers – a kinematic model. Water Resources Research, 30(4), 11731185. 10.1029/93wr03351.Google Scholar
Hassan, M. A., Reid, I., 1990. The influence of microform bed roughness elements on flow and sediment transport in gravel bed rivers. Earth Surface Processes and Landforms, 15(8), 739750.Google Scholar
Hassan, M. A., Roy, A. G., 2016. Coarse particle tracing in fluvial geomorphology. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology, 2nd Edition. Wiley and Sons, Chichester, UK, pp. 306323.Google Scholar
Hassan, M. A., Schick, A. P., Laronne, J. B., 1984. The recovery of flood dispersed coarse sediment particles – a 3-dimensional magnetic tracing method. Catena, 153162.Google Scholar
Hassan, M. A., Church, M., Ashworth, P. J., 1992. Virtual rate and mean distance of travel of individual clasts in gravel-bed rivers. Earth Surface Processes and Landforms, 17(6), 617627.Google Scholar
Hassan, M. A., Brayshaw, D., Alila, Y., Andrews, E., 2014. Effective discharge in small formerly glaciated mountain streams of British Columbia: Limitations and implications. Water Resources Research, 50(5), 44404458. 10.1002/2013wr014529.Google Scholar
Hauer, C., Habersack, H., 2009. Morphodynamics of a 1000-year flood in the Kamp River, Austria, and impacts on floodplain morphology. Earth Surface Processes and Landforms, 34(5), 654682.Google Scholar
Haviv, I., Enzel, Y., Whipple, K. X., et al., 2010. Evolution of vertical knickpoints (waterfalls) with resistant caprock: Insights from numerical modeling. Journal of Geophysical Research – Earth Surface, 115. 10.1029/2008jf001187.Google Scholar
Haw, M. D., 2002. Colloidal suspensions, Brownian motion, molecular reality: A short history. Journal of Physics – Condensed Matter, 14(33), 77697779.Google Scholar
Hawley, R. J., 2018. Making stream restoration more sustainable: a geomorphically, ecologically, and socioeconomically principled approach to bridge the practice with the science. Bioscience, 68(7), 517528.Google Scholar
Hawley, R. J., Bledsoe, B. P., 2011. How do flow peaks and durations change in suburbanizing semi-arid watersheds? A southern California case study. Journal of Hydrology, 405(1–2), 6982.Google Scholar
Hawley, R. J., Bledsoe, B. P., 2013. Channel enlargement in semiarid suburbanizing watersheds: A southern California case study. Journal of Hydrology, 496, 1730.Google Scholar
Hawley, R. J., Bledsoe, B. P., Stein, E. D., Haines, B. E., 2012. Channel evolution model of semiarid stream response to urban-induced hydromodification. Journal of the American Water Resources Association, 48(4), 722744.Google Scholar
Hawley, R. J., MacMannis, K. R., Wooten, M. S., 2013. Bed coarsening, riffle shortening, and channel enlargement in urbanizing watersheds, northern Kentucky, USA. Geomorphology, 201, 111126.Google Scholar
Hayakawa, Y., Matsukura, Y., 2003. Recession rates of waterfalls in Boso Peninsula, Japan, and a predictive equation. Earth Surface Processes and Landforms, 28(6), 675684.Google Scholar
Hayakawa, Y. S., Oguchi, T., 2014. Spatial correspondence of knickzones and stream confluences along bedrock rivers in Japan: implications for hydraulic formation of knickzones. Geografiska Annaler Series A Physical Geography, 96(1), 919.Google Scholar
He, L., Wilkerson, G. V., 2011. Improved bankfull channel geometry prediction using two-year return-period discharge. Journal of the American Water Resources Association, 47(6), 12981316.Google Scholar
Heidel, S. G., 1956. The progressive lag of sediment concentration with flood waves. Transactions, American Geophysical Union, 37, 5666.Google Scholar
Hejazi, M. I., Markus, M., 2009. Impacts of urbanization and climate variability on floods in northeastern Illinois. Journal of Hydrologic Engineering, 14(6), 606616.Google Scholar
Henck, A. C., Montgomery, D. R., Huntington, K. W., Liang, C., 2010. Monsoon control of effective discharge, Yunnan and Tibet. Geology, 38(11), 975978.Google Scholar
Henderson, F. M., 1966. Open Channel Flow. Macmillan, New York.Google Scholar
Henkle, J. E., Wohl, E., Beckman, N., 2011. Locations of channel heads in the semiarid Colorado Front Range, USA. Geomorphology, 129(3–4), 309319.Google Scholar
Heritage, G., Hetherington, D., 2007. Towards a protocol for laser scanning in fluvial geomorphology. Earth Surface Processes and Landforms, 32(1), 6674.Google Scholar
Herrero, H. S., Garcia, C. M., Pedocchi, F., et al., 2016. Flow structure at a confluence: Experimental data and the bluff body analogy. Journal of Hydraulic Research, 54(3), 263274.Google Scholar
Herrero, H. S., Diaz Lozada, J. M., Garcia, C. M., et al., 2018. The influence of tributary flow density differences on the hydrodynamic behavior of a confluent meander bend and implications for flow mixing. Geomorphology, 304, 99112.Google Scholar
Hey, R. D., 1978. Determinate hydraulic geometry of river channels. Journal of the Hydraulics Division – ASCE, 104(6), 869885.Google Scholar
Hey, R. D., Thorne, C. R., 1986. Stable channels with mobile gravel beds. Journal of Hydraulic Engineering, 112(8), 671689.Google Scholar
Hickin, E. J., 1979. Concave-bank benches on the Squamish River, British Columbia, Canada. Canadian Journal of Earth Sciences, 16(1), 200203.Google Scholar
Hickin, E. J., 1986. Concave-bank benches in the floodplains of Muskwa and Fort Nelson Rivers, British Columbia. Canadian Geographer, 30(2), 111122.Google Scholar
Hickin, E. J., Nanson, G. C., 1984. Lateral migration rates of river bends. Journal of Hydraulic Engineering, 110(11), 15571567.Google Scholar
Hickin, A. S., Kerr, B., Barchyn, T. E., Paulen, R. C., 2009. Using ground-penetrating radar and capacitively coupled resistivity to investigate 3-D fluvial architecture and grain-size distribution of a gravel floodplain in northeast British Columbia, Canada. Journal of Sedimentary Research, 79(5–6), 457477.Google Scholar
Hicks, D. M., Gomez, B., 2016. Sediment transport. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology, 2nd Edition. Wiley, Chichester, UK, pp. 324356.Google Scholar
Hicks, D. M., Mason, P. D., 1991. Roughness Characteristics of New Zealand Rivers. National Institute of Water and Atmospheric Research, Wellington, New Zealand.Google Scholar
Hicks, D. M., Hill, J., Shankar, U., 1996. Variation of suspended sediment yields around New Zealand: The relative importance of rainfall and geology. In: Walling, D. E. and Webb, B. E. (eds.), Erosion and Sediment Yield: Global and Regional Perspectives. IAHS Publication No. 236, IAHS Press, Wallingford, UK, pp. 149156.Google Scholar
Hicks, D. M., Duncan, M. J., Lane, S. N., Tal, M., Westaway, R., 2008. Contemporary morphological change in braided gravel-bed rivers: New developments from field and laboratory studies, with particular reference to the influence of riparian vegetation. In: Habersack, H., Piegay, H., Rinaldi, M. (eds.), Gravel-Bed Rivers VI: From Process Understanding to River Restoration, Vol. 11. Elsevier, Amsterdam, the Netherlands, pp. 557586.Google Scholar
Hinderer, M., 2012. From gullies to mountain belts: A review of sediment budgets at various scales. Sedimentary Geology, 280, 2159.Google Scholar
Hirsch, P. J., Abrahams, A. D., 1981. The properties of bed sediments in pools and riffles. Journal of Sedimentary Petrology, 51(3), 757760.Google Scholar
Hobbs, W. H., 1921. Studies of the cycle of glaciation. Journal of Geology, 29, 370386.Google Scholar
Hodge, R. A., Hoey, T. B., Sklar, L. S., 2011. Bed load transport in bedrock rivers: The role of sediment cover in grain entrainment, translation, and deposition. Journal of Geophysical Research – Earth Surface, 116. 10.1029/2011jf002032.Google Scholar
Hodge, R. A., Sear, D. A., Leyland, J., 2013. Spatial variations in surface sediment structure in riffle-pool sequences: a preliminary test of the Differential Sediment Entrainment Hypothesis (DSEH). Earth Surface Processes and Landforms, 38(5), 449465.Google Scholar
Hoegh-Guldberg, O., Jacob, D., Taylor, M., et al., 2018. Impacts of 1.5°C global warming on natural and human systems. In: Masson-Delmotte, V., Zhai, P., Pörtner, H.-O., et al. (eds.), Global Warming of 1.5°C. Intergovernmental Panel On Climate Change, www.ipcc.ch/sr15/chapter/chapter-3/.Google Scholar
Hoey, T. B., Bluck, B. J., 1999. Identifying the controls over downstream fining of river gravels. Journal of Sedimentary Research, 69(1), 4050.Google Scholar
Hoey, T. B., Ferguson, R., 1994. Numerical simulation of downstream fining by selective transport in gravel bed rivers – model development and illustration. Water Resources Research, 30(7), 22512260.Google Scholar
Hoey, T. B., Ferguson, R. I., 1997. Controls of strength and rate of downstream fining above a river base level. Water Resources Research, 33(11), 26012608.Google Scholar
Hoey, T. B., Sutherland, A. J., 1991. Channel morphology and bedload pulses in braided river – a laboratory study. Earth Surface Processes and Landforms, 16(5), 447462.Google Scholar
Hoffmann, T., 2015. Sediment residence time and connectivity in non-equilibrium and transient geomorphic systems. Earth-Science Reviews, 150, 609627.Google Scholar
Holeman, J. N., 1968. Sediment yield of major rivers of the world. Water Resources Research, 4(4), 737747.Google Scholar
Hollis, G. E., 1975. The effect of urbanization on floods of different recurrence interval. Water Resources Research, 11(3), 431435.Google Scholar
Holtan, H. N., Kirkpatrick, M. H., 1950. Rainfall, infiltration, and hydraulics of flow in runoff computation. Transactions, American Geophysical Union, 31(5), 771779.Google Scholar
Hong, L. B., Davies, T. R. H., 1979. A study of stream braiding – summary. Geological Society of America Bulletin, 90(12), 10941095.Google Scholar
Hooke, J. M., 1977. The distribution and nature of changes in river channel patterns. In: Gregory, K. J. (ed.), River Channel Changes. Wiley, Chichester, UK, pp. 265280.Google Scholar
Hooke, J. M., 1979. Analysis of the processes of river bank erosion. Journal of Hydrology, 42(1–2), 3962.Google Scholar
Hooke, J. M., 1980. Magnitude and distribution of rates of river bank erosion. Earth Surface Processes and Landforms, 5(2), 143157.Google Scholar
Hooke, J. M., 1986. Applicable and applied geomorphology of rivers. Geography, 71(310), 113.Google Scholar
Hooke, J. M., 1995a. Processes of channel planform change on meandering channels in the UK. In: Gurnell, A. M., Gregory, K. J. (eds.), Changing River Channels. Wiley, Chichester, UK, pp. 87115.Google Scholar
Hooke, J. M., 1995b. River channel adjustment to meander cutoffs on the River Bollin and River Dane, northwest England. Geomorphology, 14(3), 235253.Google Scholar
Hooke, J. M., 2003a. Coarse sediment connectivity in river channel systems: a conceptual framework and methodology. Geomorphology, 56(1–2), 7994.Google Scholar
Hooke, J. M., 2003b. River meander behaviour and instability: a framework for analysis. Transactions of the Institute of British Geographers, 28(2), 238253.Google Scholar
Hooke, J. M., 2004a. Analysis of coarse sediment connectivity in semiarid river channels. In: Golosov, V., Belyaev, V., Walling, D. E. (eds.), Sediment Transfer through the Fluvial System. IAHS Publication No. 288, IAHS Press, Wallingford, UK, pp. 269275.Google Scholar
Hooke, J. M., 2004b. Cutoffs galore!: Occurrence and causes of multiple cutoffs on a meandering river. Geomorphology, 61(3–4), 225238.Google Scholar
Hooke, J. M., 2007. Complexity, self-organisation and variation in behaviour in meandering rivers. Geomorphology, 91(3–4), 236258.Google Scholar
Hooke, J. M., 2016a. Morphological impacts of flow events of varying magnitude on ephemeral channels in a semiarid region. Geomorphology, 252, 128143.Google Scholar
Hooke, J. M., 2016b. Geomorphological impacts of an extreme flood in SE Spain. Geomorphology, 263, 1938.Google Scholar
Hooke, J. M., Harvey, M. D., 1983. Meander changes in relation to bed morphology and secondary flows. In: Collinson, J. D., Lewin, J. (eds.), Modern and Ancient Fluvial Systems. International Association of Sedimentologists, Blackwell, Oxford, UK, pp. 121132.Google Scholar
Hooke, J. M., Mant, J., 2015. Morphological and vegetation variations in response to flow events in rambla channels of SE Spain. In: Dykes, A. P., Mulligan, M., Wainwright, J. (eds.), Monitoring and Modeling of Dynamic Environments. Wiley, Chichester, UK, pp. 6196.Google Scholar
Hooke, R. L. B., 1975. Distribution of sediment transport and shear stress in a meander bend. Journal of Geology, 83(5), 543565.Google Scholar
Hooke, R. L., 1994. On the efficacy of humans as geomorphic agents. GSA Today, 4, 217, 224225.Google Scholar
Hooke, R. L., 2000. On the history of humans as geomorphic agents. Geology, 28(9), 843846.Google Scholar
Hooke, R. L., 2012. Land transformation by humans: A review. GSA Today, 22, 410.Google Scholar
Hooshyar, M., Singh, A., Wang, D., 2017. Hydrologic controls on junction angle of river networks. Water Resources Research, 53(5), 40734083.Google Scholar
Hopkinson, L. C., Wynn-Thompson, T. M., 2016. Comparison of direct and indirect boundary shear stress measurements along vegetated streambanks. River Research and Applications, 32(8), 17551764.Google Scholar
Horn, J. D., Joeckel, R. M., Fielding, C. R., 2012. Progressive abandonment and planform changes of the central Platte River in Nebraska, central USA, over historical timeframes. Geomorphology, 139, 372383.Google Scholar
Horowitz, A. J., 2003. An evaluation of sediment rating curves for estimating suspended sediment concentrations for subsequent flux calculations. Hydrological Processes, 17(17), 33873409.Google Scholar
Horton, A. J., Constantine, J. A., Hales, T. C., et al., 2017. Modification of river meandering by tropical deforestation. Geology, 45(6), 511514.Google Scholar
Horton, R. E., 1945. Erosional development of streams and their drainage basins: A hydrophysical approach to quantitative morphology. Geological Society of America Bulletin, 56, 275370.Google Scholar
Hovius, N., 1996. Regular spacing of drainage outlets from linear mountain belts. Basin Research, 8(1), 2944.Google Scholar
Hovius, N., Meunier, P., Ching-Weei, L., et al., 2011. Prolonged seismically induced erosion and the mass balance of a large earthquake. Earth and Planetary Science Letters, 304(3–4), 347355.Google Scholar
Howard, A. D., 1967. Drainage analysis in geologic interpretation: A summation. American Association of Petroleum Geologists Bulletin, 51, 22462259.Google Scholar
Howard, A. D., 1971a. Simulation of stream networks by headward growth and branching. Geographical Analysis, 3(1), 2950.Google Scholar
Howard, A. D., 1971b. Simulation model of stream capture. Geological Society of America Bulletin, 82(5), 13551375.Google Scholar
Howard, A. D., 1971c. Optimal angles of stream junction: Geometric, stability to capture, and minimum power criteria. Water Resources Research, 7(4), 863873.Google Scholar
Howard, A. D., 1982. Equilibrium and time scales in geomorphology – application to sand-bed alluvial streams. Earth Surface Processes and Landforms, 7(4), 303325.Google Scholar
Howard, A. D., 1990. Theoretical model of optimal drainage networks. Water Resources Research, 26(9), 21072117.Google Scholar
Howard, A. D., 1992. Modelling channel migration and floodplain sedimentation in meandering streams. In: Carling, P. A., Petts, G. E. (eds.), Lowland Floodplain Rivers. Wiley, New York, pp. 141.Google Scholar
Howard, A. D., 1994. A detachment-limited model of drainage basin evolution. Water Resources Research, 30(7), 22612285.Google Scholar
Howard, A., Dolan, R., 1981. Geomorphology of the Colorado River in the Grand Canyon. Journal of Geology, 89(3), 269298.Google Scholar
Howard, A. D., Knutson, T. R., 1984. Sufficient conditions for river meandering – a simulation approach. Water Resources Research, 20(11), 16591667. 10.1029/WR020i011p01659.Google Scholar
Howard, A. D., McLane, C. F., 1988. Erosion of cohesionless sediment by groundwater seepage. Water Resources Research, 24(10), 16591674. 10.1029/WR024i010p01659.Google Scholar
Hsu, C.-C., Lee, W.-J., Chang, C.-H., 1998a. Subcritical open-channel junction flow. Journal of Hydraulic Engineering, 24, 847855.Google Scholar
Hsu, C.-C., Wu, F.-S., Lee, W.-J., 1998b. Flow at 90° equal-width open-channel junction. Journal of Hydraulic Engineering, 124, 186191.Google Scholar
Huang, H. Q., 2010. Reformulation of the bed load equation of Meyer-Peter and Muller in light of the linearity theory for alluvial channel flow. Water Resources Research, 46. 10.1029/2009wr008974.Google Scholar
Huang, H. Q., Nanson, G. C., 1997. Vegetation and channel variation; a case study of four small streams in southeastern Australia. Geomorphology, 18 (3–4), 237249. 10.1016/s0169-555x(96)00028–1.Google Scholar
Huang, H. Q., Nanson, G. C., 1998. The influence of bank strength on channel geometry: An integrated analysis of some observations. Earth Surface Processes and Landforms, 23(10), 865876.Google Scholar
Huang, H. Q., Nanson, G. C., 2000. Hydraulic geometry and maximum flow efficiency as products of the principle of least action. Earth Surface Processes and Landforms, 25(1), 116.Google Scholar
Huang, H. Q., Nanson, G. C., 2002. A stability criterion inherent in laws governing alluvial channel flow. Earth Surface Processes and Landforms, 27(9), 929944.Google Scholar
Huang, H. Q., Nanson, G. C., 2007. Why some alluvial rivers develop an anabranching pattern. Water Resources Research, 43(7). 10.1029/2006wr005223.Google Scholar
Huang, H. Q., Warner, R. F., 1995. The multivariate controls of hydraulic geometry – a causal investigation in terms of boundary shear distribution. Earth Surface Processes and Landforms, 20(2), 115130.Google Scholar
Huang, H., Chang, H. H., Nanson, G. C., 2004. Minimum energy as the general form of critical flow and maximum flow efficiency and for explaining variations in river channel pattern. Water Resources Research, 40(4), W04502. 10.1029/2003wr002539.Google Scholar
Huang, J. C., Weber, L. J., Lai, Y. G., 2002. Three-dimensional numerical study of flows in open-channel junctions. Journal of Hydraulic Engineering, 128(3), 268280.Google Scholar
Huang, S. Y., Cheng, S. J., Wen, J. C., Lee, J. H., 2008. Identifying peak-imperviousness-recurrence relationships on a growing-impervious watershed, Taiwan. Journal of Hydrology, 362(3–4), 320336.Google Scholar
Huckleberry, G., 1994. Contrasting channel response to floods on the middle Gila River, Arizona Geology, 22(12), 10831086.Google Scholar
Hudson, P. F., 2002. Pool-riffle morphology in an actively migrating alluvial channel: The Lower Mississippi River. Physical Geography, 23(2), 154169.Google Scholar
Hudson, P. F., Heitmuller, F. T., 2003. Local- and watershed-scale controls on the spatial variability of natural levee deposits in a large fine-grained floodplain: Lower Panuco Basin, Mexico. Geomorphology, 56(3–4), 255269.Google Scholar
Hudson-Edwards, K. A., Schell, C., Macklin, M. G., 1999. Mineralogy and geochemistry of alluvium contaminated by metal mining in the Rio Tinto area, southwest Spain. Applied Geochemistry, 14(8), 10151030.Google Scholar
Hundey, E. J., Ashmore, P. E., 2009. Length scale of braided river morphology. Water Resources Research, 45. 10.1029/2008wr007521.Google Scholar
Hung, C.-L.J., James, L. A., Carbone, G. J., 2018. Impacts of urbanization on stormflow magnitudes in small catchments in the Sandhills of South Carolina, USA. Anthropocene, 23, 1728.Google Scholar
Hupp, C. R., 1992. Riparian vegetation recovery patterns following stream channelization – a geomorphic perspective. Ecology, 73(4), 12091226.Google Scholar
Hupp, C. R., Simon, A., 1991. Bank accretion and the development of vegetation depositional surfaces along modified alluvial channels. Geomorphology, 4(2), 111124.Google Scholar
Hurther, D., Lemmin, U., 2000. Shear stress statistics and wall similarity analysis in turbulent boundary layers using a high-resolution 3-D ADVP. IEEE Journal of Oceanic Engineering, 25(4), 446457.Google Scholar
Hutton, J., 1795. Theory of the Earth, with Proofs and Illustrations. Vol. 1&2. Edinburgh; London: Printed for Cadell and Davies; William Creech.Google Scholar
Hyde, K. D., Wilcox, A. C., Jencso, K., Woods, S., 2014. Effects of vegetation disturbance by fire on channel initiation thresholds. Geomorphology, 214, 8496.Google Scholar
Ianniruberto, M., Trevethan, M., Pinheiro, A., et al., 2018. A field study of the confluence between Negro and Solimoes Rivers. Part 2: Bed morphology and stratigraphy. Comptes Rendus Geoscience, 350(1–2), 4354.Google Scholar
Ibbitt, R. P., Willgoose, G. R., Duncan, M. J., 1999. Channel network simulation models compared with data from the Ashley River, New Zealand. Water Resources Research, 35(12), 38753890.Google Scholar
Ijjasz-Vasquez, E. J., Bras, R. L., 1995. Scaling regimes of local slope versus contributing area in digital elevation models. Geomorphology, 12(4), 299311.Google Scholar
Ijjasz-Vasquez, E. J., Bras, R. L., Rodriguez-Iturbe, I., 1993. Hack’s relation and optimal channel networks – the elongation of river basins as a consequence of energy minimization. Geophysical Research Letters, 20(15), 15831586.Google Scholar
Ikeda, H., 1973. A study on the formation of sand bars in an experimental flume. Geographical Review of Japan, 46–7, 435451.Google Scholar
Ikeda, H., 1989. Sedimentary controls on channel migration and origin of point bars in sand-bedded meandering rivers. In: Ikeda, S., Parker, G. (eds.), River Meandering. Water Resources Monograph 12. American Geophysical Union, Washington, DC, pp. 5168.Google Scholar
Ikeda, S., 1984a. Flow and bed topography in channels with alternate bars. In: Elliott, C. M. (ed.), River Meandering. American Society of Civil Engineers, New York, pp. 733746.Google Scholar
Ikeda, S., 1984b. Prediction of alternate bar wavelength and height. Journal of Hydraulic Engineering, 110, 371386.Google Scholar
Ikeda, S., Izumi, N., 1990. Width and depth of self-formed straight gravel rivers with bank vegetation. Water Resources Research, 26(10), 23532364.Google Scholar
Ikeda, S., Parker, G., Sawai, K., 1981. Bend theory of river meanders. 1. Linear development. Journal of Fluid Mechanics, 112(NOV), 363377.Google Scholar
Ikeda, S., Parker, G., Kimura, Y., 1988. Stable width and depth of straight gravel rivers with heterogeneous bed materials. Water Resources Research, 24(5), 713722.Google Scholar
Imaizumi, F., Hattanji, T., Hayakawa, Y. S., 2010. Channel initiation by surface and subsurface flows in a steep catchment of the Akaishi Mountains, Japan. Geomorphology, 115(1–2), 3242.Google Scholar
Imhoff, K. S., Wilcox, A. C., 2016. Coarse bedload routing and dispersion through tributary confluences. Earth Surface Dynamics, 4(3), 591605.Google Scholar
Inoue, T., Parker, G., Stark, C. P., 2017. Morphodynamics of a bedrock-alluvial meander bend that incises as it migrates outward: approximate solution of permanent form. Earth Surface Processes and Landforms, 42(9), 13421354.Google Scholar
Ishii, Y., Hori, K., 2016. Formation and infilling of oxbow lakes in the Ishikari lowland, northern Japan. Quaternary International, 397, 136146.Google Scholar
Istanbulluoglu, E., Bras, R. L., 2005. Vegetation-modulated landscape evolution: effects of vegetation on landscape processes, drainage density, and topography. Journal of Geophysical Research – Earth Surface, 110(F2), 19. 10.1029/2004jf000249.Google Scholar
Istanbulluoglu, E., Bras, R. L., 2006. On the dynamics of soil moisture, vegetation, and erosion: Implications of climate variability and change. Water Resources Research, 42(6). 10.1029/2005wr004113.Google Scholar
Istanbulluoglu, E., Tarboton, D. G., Pack, R. T., Luce, C., 2002. A probabilistic approach for channel initiation. Water Resources Research, 38(12). 10.1029/2001wr000782.Google Scholar
Izenberg, N. R., Arvidson, R. E., Brackett, R. A., et al., 1996. Erosional and depositional patterns associated with the 1993 Missouri river floods inferred from SIR-C and TOPSAR radar data. Journal of Geophysical Research – Planets, 101(E10), 2314923167.Google Scholar
Izumi, N., Parker, G., 1995. Inception of channelization and drainage-basin formation – upstream-driven theory. Journal of Fluid Mechanics, 283, 341363.Google Scholar
Izumi, N., Parker, G., 2000. Linear stability analysis of channel inception: Downstream-driven theory. Journal of Fluid Mechanics, 419, 239262.Google Scholar
Jackson, C. R., Martin, J. K., Leigh, D. S., West, L. T., 2005. A southeastern piedmont watershed sediment budget: Evidence for a multi-millennial agricultural legacy. Journal of Soil and Water Conservation, 60(6), 298310.Google Scholar
Jackson, R. G., 1976. Largescale ripples of the lower Wabash River. Sedimentology, 23(5), 593623.Google Scholar
Jackson, W. L., Beschta, R. L., 1982. A model of 2-phase bedload transport in an Oregon Coast Range stream. Earth Surface Processes and Landforms, 7(6), 517527.Google Scholar
Jacobson, R. B., Femmer, S. R., McKenney, R. A., 2001. Land use changes and the physical habitat of streams – a review with emphasis on studies within the U.S. Geological Survey federal-state cooperative program. U.S. Geological Survey Circular 1175, Reston, VA.Google Scholar
Jacobson, R. B., O’Connor, J. E., Oguchi, T., 2016. Surficial geological tools in fluvial geomorphology. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology, 2nd Edition. Wiley, Chichester, UK, pp. 1539.Google Scholar
Jaeger, K. L., Montgomery, D. R., Bolton, S. M., 2007. Channel and perennial flow initiation in headwater streams: management implications of variability in source-area size. Environmental Management, 40(5), 775786.Google Scholar
Jaeggi, M. N. R., 1984. Formation and effects of alternate bars. Journal of Hydraulic Engineering, 110, 142156.Google Scholar
Jaehnig, S. C., Lorenz, A. W., Hering, D., et al., 2011. River restoration success: A question of perception. Ecological Applications, 21(6), 20072015.Google Scholar
Jain, S. C., 1990. Armor or pavement. Journal of Hydraulic Engineering, 116(3), 436440.Google Scholar
Jain, V., Fryirs, K., Brierley, G., 2008. Where do floodplains begin? The role of total stream power and longitudinal profile form on floodplain initiation processes. Geological Society of America Bulletin, 120(1–2), 127141.Google Scholar
James, C. S., 1985. Sediment transfer to overbank sections. Journal of Hydraulic Research, 23(5), 435452.Google Scholar
James, C. S., 1990. Prediction of entrainment conditions for nonuniform, noncohesive sediments. Journal of Hydraulic Research, 28(1), 2541.Google Scholar
James, L. A., 1989. Sustained storage and transport of hydraulic gold mining sediment in the Bear River, California. Annals of the Association of American Geographers, 79(4), 570592.Google Scholar
James, L. A., 1991. Incision and morphological evolution of an alluvial channel recovering from hydraulic mining sediment. Geological Society of America Bulletin, 103(6), 723736.Google Scholar
James, L. A., 1997. Channel incision on the lower American river, California, from streamflow gage records. Water Resources Research, 33(3), 485490.Google Scholar
James, L. A., 2004. Tailings fans and valley-spur cutoffs created by hydraulic mining. Earth Surface Processes and Landforms, 29(7), 869882.Google Scholar
James, L.A., 2006. Bed waves at the basin scale: Implications for river management and restoration. Earth Surface Processes and Landforms, 31(13), 16921706.Google Scholar
James, L. A., 2013. Impact of early agriculture and deforestation on geomorphic systems. In: Shroder, J. C. (ed.), Treatise on Geomorphology, Vol. 13, Geomorphology of Human Disturbances, Climate Change, and Natural Hazards, James, L. A., Harden, C. P., Clague, J. J. (vol. eds.). Academic Press, San Diego, CA, pp. 4867.Google Scholar
James, L. A., 2018. Ten conceptual models of large-scale legacy sedimentation – a review. Geomorphology, 317, 199217.Google Scholar
James, L. A., Lecce, S. A., 2013. Impacts of land-use and land-cover change on river systems. In: Shroder, J. C. (ed.), Treatise on Geomorphology, Vol. 9, Fluvial Geomorphology, Wohl, E. (vol. ed.). Academic Press, San Diego, CA, pp. 768793.Google Scholar
James, L. A., Phillips, J. D., Lecce, S. A., 2017. A centennial tribute to GK Gilbert’s Hydraulic Mining Debris in the Sierra Nevada. Geomorphology, 294, 419.Google 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.Google Scholar
James, W. R., Krumbein, W. C., 1969. Frequency distributions of stream link lengths. Journal of Geology, 77(5), 544565.Google Scholar
Jansen, J. D., Nanson, G. C., 2004. Anabranching and maximum flow efficiency in Magela Creek, northern Australia. Water Resources Research, 40(4). 10.1029/2003wr002408.Google Scholar
Jansen, J. D., Nanson, G. C., 2010. Functional relationships between vegetation, channel morphology, and flow efficiency in an alluvial (anabranching) river. Journal of Geophysical Research – Earth Surface, 115. 10.1029/2010jf001657.Google Scholar
Jansen, J. M. L., Painter, R. B., 1974. Predicting sediment yield from climate and topography. Journal of Hydrology, 21, 371380.Google Scholar
Jansson, M., 1985. A comparison of detransformed logarithmic regressions and power function regressions. Geografiska Annaler Series A Physical Geography, 67(1–2), 6170.Google Scholar
Jansson, M. B., 1996. Estimating a sediment rating curve of the Reventazon River at Palomo using logged mean loads within discharge classes. Journal of Hydrology, 183(3–4), 227241.Google Scholar
Jarrett, R. D., 1984. Hydraulics of high-gradient streams. Journal of Hydraulic Engineering, 110(11), 15191539.Google Scholar
Jarvis, N., Koestel, J., Messing, I., Moeys, J., Lindahl, A., 2013. Influence of soil, land use and climatic factors on the hydraulic conductivity of soil. Hydrology and Earth System Sciences, 17(12), 51855195.Google Scholar
Javemick, L., Brasington, J., Caruso, B., 2014. Modeling the topography of shallow braided rivers using structure-from-motion photogrammetry. Geomorphology, 213, 166182.Google Scholar
Jefferson, A. J., McGee, R. W., 2013. Channel network extent in the context of historical land use, flow generation processes, and landscape evolution in the North Carolina Piedmont. Earth Surface Processes and Landforms, 38(6), 601613.Google Scholar
Jeffries, R., Darby, S. E., Sear, D. A., 2003. The influence of vegetation and organic debris on flood-plain sediment dynamics: Case study of a low-order stream in the New Forest, England. Geomorphology, 51(1–3), 6180.Google Scholar
Jeong, A., Dorn, R. I., 2019. Soil erosion from urbanization processes in the Sonoran Desert, Arizona, USA. Land Degradation & Development, 30(2), 226238.Google Scholar
Jerolmack, D. J., Brzinski, T. A., III, 2010. Equivalence of abrupt grain-size transitions in alluvial rivers and eolian sand seas: a hypothesis. Geology, 38(8), 719722.Google Scholar
Jerolmack, D. J., Mohrig, D., 2007. Conditions for branching in depositional rivers. Geology, 35(5), 463466.Google Scholar
Jha, S., Bombardelli, F., 2009. Two-phase modeling of turbulence in dilute sediment-laden, open-channel flows. Environmental Fluid Mechanics, 9(2), 237266.Google Scholar
Jia, Y. F., 1990. Minimum Froude number and the equilibrium of alluvial sand rivers. Earth Surface Processes and Landforms, 15(3), 199209.Google Scholar
Joeckel, R. M., Tucker, S. T., McMullin, J. D., 2016. Morphosedimentary features from a major flood on a small, lower-sinuosity, single-thread river: The unknown quantity of overbank deposition, historical-change context, and comparisons with a multichannel river. Sedimentary Geology, 343, 1837.Google Scholar
Johannesson, H., Parker, G., 1989. Linear theory of river meanders. In: Ikeda, S., Parker, G. (eds.), River Meandering. Water Resources Monograph 12. American Geophysical Union, Washington, DC, pp. 181212.Google Scholar
Johnson, J. P. L., Aronovitz, A. C., Kim, W., 2015. Coarser and rougher: Effects of fine gravel pulses on experimental step-pool channel morphodynamics. Geophysical Research Letters, 42(20), 84328440. 10.1002/2015gl066097.Google Scholar
Johnson, P. A., Heil, T. M., 1996. Uncertainty in estimating bankfull conditions. Water Resources Bulletin, 32(6), 12831291.Google Scholar
Johnson, W. C., 1994. Woodland expansion in the Platte River, Nebraska – patterns and causes. Ecological Monographs, 64(1), 4584.Google Scholar
Johnson, W.C., 1997. Equilibrium response of riparian vegetation to flow regulation in the Platte River, Nebraska. Regulated Rivers – Research & Management, 13(5), 403415.Google Scholar
Johnston, C. E., Andrews, E. D., Pitlick, J., 1998. In situ determination of particle friction angles of fluvial gravels. Water Resources Research, 34(8), 20172030.Google Scholar
Jones, J. A., Grant, G. E., 1996. Peak flow responses to clear-cutting and roads in small and large basins, western Cascades, Oregon. Water Resources Research, 32(4), 959974.Google Scholar
Jones, L. S., Humphrey, N. F., 1997. Weathering-controlled abrasion in a coarse-grained, meandering reach of the Rio Grande: implications for the rock record. Geological Society of America Bulletin, 109(9), 10801088.Google Scholar
Jones, L. S., Schumm, S. A., 1999. Causes of avulsion: an overview. In: Smith, N.D., Rodgers, J. (eds.), Fluvial Sedimentology VI, International Association of Sedimentologists Special Publication 28, Blackwell, Malden, MA, pp.171178.Google Scholar
Jones, O. T., 1924. The Upper Towry drainage system. Quarterly Journal of the Geological Society of London, 80, 560609.Google Scholar
Jordan, B. A., Annable, W. K., Watson, C. C., Sen, D., 2010. Contrasting stream stability characteristics in adjacent urban watersheds: Santa Clara Valley, California. River Research and Applications, 26(10), 12811297.Google Scholar
Jothiprakash, V., Garg, V., 2008. Re-look to conventional techniques for trapping efficiency estimation of a reservoir. International Journal of Sediment Research, 23(1), 7684.Google Scholar
Joy, D. M., Townsend, R. D., 1981. Improved flow characteristics at a 90° channel confluence, 5th Canadian Hydrotechnical Conference. Canadian Society of Civil Engineers, Fredericton, New Brunswick, pp. 781799.Google Scholar
Julian, J. P., Elmore, A. J., Guinn, S. M., 2012. Channel head locations in forested watersheds across the mid-Atlantic United States: A physiographic analysis. Geomorphology, 177, 194203.Google Scholar
Jung, K., Marpu, P. R., Ouarda, T., 2015. Improved classification of drainage networks using junction angles and secondary tributary lengths. Geomorphology, 239, 4147.Google Scholar
Junker, B., Buchecker, M., Mueller-Boeker, U., 2007. Objectives of public participation: Which actors should be involved in the decision making for river restorations? Water Resources Research, 43(10). 10.1029/2006WR005584.Google Scholar
Juracek, K. E., 2004. Historical channel-bed elevation change as a result of multiple disturbances, Soldier Creek, Kansas. Physical Geography, 25(4), 269290.Google Scholar
Juracek, K. E., Fitzpatrick, F. A., 2003. Limitations and implications of stream classification. Journal of the American Water Resources Association, 39(3), 659670.Google Scholar
Kabiri-Samani, A., Farshi, F., Chamani, M. R., 2013. Boundary shear stress in smooth trapezoidal open channel flows. Journal of Hydraulic Engineering, 139(2), 205212.Google Scholar
Kail, J., Brabec, K., Poppe, M., Januschke, K., 2015. The effect of river restoration on fish, macroinvertebrates and aquatic macrophytes: A meta-analysis. Ecological Indicators, 58, 311321.Google Scholar
Karcz, I., 1966. Secondary currents and the configuration of a natural stream bed. Journal of Geophysical Research, 71, 31093117.Google Scholar
Karlstrom, L., Gajjar, P., Manga, M., 2013. Meander formation in supraglacial streams. Journal of Geophysical Research – Earth Surface, 118(3), 18971907.Google Scholar
Karr, J. R., 1981. Assessment of biotic integrity using fish communities. Fisheries, 6(6), 2127.Google Scholar
Kasprak, A., Magilligan, F. J., Nislow, K. H., et al., 2013. Differentiating the relative importance of land cover change and geomorphic processes on fine sediment sequestration in a logged watershed. Geomorphology, 185, 6777.Google Scholar
Kasprak, A., Wheaton, J. M., Ashmore, P. E., Hensleigh, J. W., Peirce, S., 2015. The relationship between particle travel distance and channel morphology: Results from physical models of braided rivers. Journal of Geophysical Research – Earth Surface, 120(1), 5574.Google Scholar
Kasprak, A., Hough-Snee, N., Beechie, T., et al., 2016. The blurred line between form and process: A comparison of stream channel classification frameworks. PLoS ONE, 11(3), 31. 10.1371/journal.pone.0150293.Google Scholar
Kasvi, E., Vaaja, M., Kaartinen, H., et al., 2015. Sub-bend scale flow-sediment interaction of meander bends – a combined approach of field observations, close-range remote sensing and computational modelling. Geomorphology, 238, 119134.Google Scholar
Kasvi, E., Hooke, J., Kurkela, M., et al., 2017. Modern empirical and modelling study approaches in fluvial geomorphology to elucidate sub-bend-scale meander dynamics. Progress in Physical Geography, 41(5), 533569.Google Scholar
Kean, J. W., Smith, J. D., 2004. Flow and boundary shear stress in channels with woody bank vegetation. In: Bennett, S. J., Simon, A. (eds.), Riparian Vegetation and Fluvial Geomorphology. American Geophysical Union, Washington, DC, pp. 237252.Google Scholar
Kean, J. W., Smith, J. D., 2006a. Form drag in rivers due to small-scale natural topographic features: 1. regular sequences. Journal of Geophysical Research – Earth Surface, 111(F4). 10.1029/2006jf000467.Google Scholar
Kean, J. W., Smith, J. D., 2006b. Form drag in rivers due to small-scale natural topographic features: 2. irregular sequences. Journal of Geophysical Research – Earth Surface, 111(F4). 10.1029/2006jf000490.Google Scholar
Kean, J. W., Kuhnle, R. A., Smith, J. D., Alonso, C. V., Langendoen, E. J., 2009. Test of a method to calculate near-bank velocity and boundary shear stress. Journal of Hydraulic Engineering, 135(7), 588601.Google Scholar
Kearsley, L. H., Schmidt, J. C., Warren, K. D., 1994. Effects of Glen Canyon Dam on Colorado River sand deposits used as campsites in Grand Canyon National Park, USA. Regulated Rivers-Research & Management, 9(3), 137149.Google Scholar
Keast, D., Ellison, J., 2013. Magnitude frequency analysis of small floods using the annual and partial series. Water, 5(4), 18161829.Google Scholar
Keller, E. A., 1971. Areal sorting of bed-load material: the hypothesis of velocity reversal. Geological Society of America Bulletin, 82, 753756.Google Scholar
Keller, E. A., 1972. Development of alluvial stream channels – a five-stage model. Geological Society of America Bulletin, 83(5), 15311536.Google Scholar
Keller, E. A., 1975. Channelization – search for a better way. Geology, 3(5), 246248.Google Scholar
Keller, E. A., Florsheim, J. L., 1993. Velocity-reversal hypothesis – a model approach. Earth Surface Processes and Landforms, 18(8), 733740.Google Scholar
Keller, E. A., Melhorn, W. N., 1978. Rhythmic spacing and origin of pools and riffles. Geological Society of America Bulletin, 89(5), 723730.Google Scholar
Kelly, S., 2006. Scaling and hierarchy in braided rivers and their deposits: examples and implications for reservoir modelling. In: Sambrook Smith, G. H., Best, J. B., Bristow, C. S., Petts, G. E. (eds.), Braided Rivers: Process, Deposits, Ecology and Management, International Association of Sedimentologists Special Publication 36. Blackwell, Malden, MA, pp. 75106.Google Scholar
Kelly, S. A., Takbiri, Z., Belmont, P., Foufoula-Georgiou, E., 2017. Human amplified changes in precipitation-runoff patterns in large river basins of the midwestern United States. Hydrology and Earth System Sciences, 21(10), 50655088.Google Scholar
Kemp, J., 2010. Downstream channel changes on a contracting, anabranching river: The Lachlan, southeastern Australia. Geomorphology, 121(3–4), 231244.Google Scholar
Kennedy, B. A., 1984. On Playfair’s Law of accordant junctions. Earth Surface Processes and Landforms, 9(2), 153173.Google Scholar
Kennedy, J. F., 1995. The Albert Shields story. Journal of Hydraulic Engineering, 121(11), 766772.Google Scholar
Kenworthy, S. T., Rhoads, B. L., 1995. Hydrologic control of spatial patterns of suspended sediment concentration at a stream confluence. Journal of Hydrology, 168(1–4), 251263.Google Scholar
Kesel, R. H., Yodis, E. G., 1992. Some effects of human modifications on sand-bed channels in southwestern Mississippi, USA. Environmental Geology and Water Sciences, 20(2), 93104.Google Scholar
Kesseli, T. E., 1941. The concept of the graded river. Journal of Geology, 49(6), 561588.Google Scholar
Keylock, C. J., 2015. Flow resistance in natural, turbulent channel flows: the need for a fluvial fluid mechanics. Water Resources Research, 51(6), 43744390. 10.1002/2015wr016989.Google Scholar
Keylock, C. J., Constantinescu, G., Hardy, R. J., 2012. The application of computational fluid dynamics to natural river channels: Eddy resolving versus mean flow approaches. Geomorphology, 179, 120.Google Scholar
Keylock, C. J., Singh, A., Foufoula-Georgiou, E., 2013. The influence of migrating bed forms on the velocity-intermittency structure of turbulent flow over a gravel bed. Geophysical Research Letters, 40(7). 10.1002/grl.50337.Google Scholar
Keylock, C. J., Singh, A., Venditti, J. G., Foufoula-Georgiou, E., 2014a. Robust classification for the joint velocity-intermittency structure of turbulent flow over fixed and mobile bedforms. Earth Surface Processes and Landforms, 39(13), 17171728.Google Scholar
Keylock, C. J., Lane, S. N., Richards, K. S., 2014b. Quadrant/octant sequencing and the role of coherent structures in bed load sediment entrainment. Journal of Geophysical Research – Earth Surface, 119(2), 264286. 10.1002/2012jf002698.Google Scholar
Khodashenas, S. R., Paquier, A., 1999. A geometrical method for computing the distribution of boundary shear stress across irregular straight open channels. Journal of Hydraulic Research, 37(3), 381388.Google Scholar
Kidson, R., Richards, K. S., 2005. Flood frequency analysis: assumptions and alternatives. Progress in Physical Geography, 29(3), 392410.Google Scholar
Kiffney, P. M., Greene, C. M., Hall, J. E., Davies, J. R., 2006. Tributary streams create spatial discontinuities in habitat, biological productivity, and diversity in mainstem rivers. Canadian Journal of Fisheries and Aquatic Sciences, 63(11), 25182530.Google Scholar
Kim, S. C., Friedrichs, C. T., Maa, J. P. Y., Wright, L. D., 2000. Estimating bottom stress in tidal boundary layer from Acoustic Doppler Velocimeter data. Journal of Hydraulic Engineering, 126(6), 399406.Google Scholar
Kincey, M., Warburton, J., Brewer, P., 2018. Contaminated sediment flux from eroding abandoned historical metal mines: spatial and temporal variability in geomorphological drivers. Geomorphology, 319, 199215.Google Scholar
King, L. C., 1953. Canons of landscape evolution. Geological Society of America Bulletin, 64(7), 721752.Google Scholar
Kinnell, P. I. A., 2004. Sediment delivery ratios: a misaligned approach to determining sediment delivery from hillslopes. Hydrological Processes, 18(16), 31913194.Google Scholar
Kinnell, P. I. A., 2008a. Sediment delivery from hillslopes and the Universal Soil Loss Equation: some perceptions and misconceptions. Hydrological Processes, 22(16), 31683175.Google Scholar
Kinnell, P. I. A., 2008b. Discussion: misrepresentation of the USLE in “Is sediment delivery a fallacy?”. Earth Surface Processes and Landforms, 33(10), 16271629.Google Scholar
Kirby, E., Whipple, K., 2001. Quantifying differential rock-uplift rates via stream profile analysis. Geology, 29(5), 415418.Google Scholar
Kirchner, J. W., 1993. Statistical inevitability of Horton’s Laws and the apparent randomness of stream channel networks. Geology, 21(7), 591594.Google Scholar
Kirchner, J. W., Dietrich, W. E., Iseya, F., Ikeda, H., 1990. The variability of critical shear stress, friction angle, and grain protrusion in water-worked sediments. Sedimentology, 37(4), 647672.Google Scholar
Kirchner, J. W., Finkel, R. C., Riebe, C. S., et al., 2001. Mountain erosion over 10 yr, 10 k.y., and 10 m.y. time scales. Geology, 29(7), 591594.Google Scholar
Kirkby, M. J., 1976. Tests of random network model and its application to basin hydrology. Earth Surface Processes and Landforms, 1(3), 197212.Google Scholar
Kirkby, M. J., 1980. The streamhead as a significant geomorphic threshold. In: Coates, D. R., Vitek, J. D. (eds.), Thresholds in Geomorphology. Allen and Unwin, Boston, MA, pp. 5373.Google Scholar
Kirkby, M. J., 1996. A role for theoretical models in geomorphology? In: Rhoads, B. L., Thorn, C. E. (eds.), The Scientific Nature of Geomorphology. Wiley and Sons, Chichester, UK, pp. 257272.Google Scholar
Kirkby, M. J., Bracken, L. J., 2009. Gully processes and gully dynamics. Earth Surface Processes and Landforms, 34(14), 18411851.Google Scholar
Kisling-Moller, J., 1992. Lateral sediment transport by bedforms in a meander bend. Earth Surface Processes and Landforms, 17(5), 501513.Google Scholar
Kitanidis, P. K., Kennedy, J. F., 1984. Secondary current and river-meander formation. Journal of Fluid Mechanics, 144(JUL), 217229.Google Scholar
Klavon, K., Fox, G., Guertault, L., et al., 2017. Evaluating a process-based model for use in streambank stabilization: Insights on the Bank Stability and Toe Erosion Model (BSTEM). Earth Surface Processes and Landforms, 42(1), 191213.Google Scholar
Klein, M., 1984. Anticlockwise hysteresis in suspended sediment concentration during individual storms – Holbeck Catchment – Yorkshire, England. Catena, 11(2–3), 251257. 10.1016/0341-8162(84)90014-6.Google Scholar
Kleinhans, M. G., 2004. Sorting in grain flows at the lee side of dunes. Earth-Science Reviews, 65(1–2), 75102.Google Scholar
Kleinhans, M. G., van den Berg, J. H., 2011. River channel and bar patterns explained and predicted by an empirical and a physics-based method. Earth Surface Processes and Landforms, 36(6), 721738.Google Scholar
Kleinhans, M. G., Jagers, H. R. A., Mosselman, E., Sloff, C. J., 2008. Bifurcation dynamics and avulsion duration in meandering rivers by one-dimensional and three-dimensional models. Water Resources Research, 44(8). 10.1029/2007wr005912.Google Scholar
Kleinhans, M. G., de Haas, T., Lavooi, E., Makaske, B., 2012. Evaluating competing hypotheses for the origin and dynamics of river anastomosis. Earth Surface Processes and Landforms, 37(12), 13371351.Google Scholar
Kleinhans, M. G., Ferguson, R. I., Lane, S. N., Hardy, R. J., 2013. Splitting rivers at their seams: bifurcations and avulsion. Earth Surface Processes and Landforms, 38(1), 4761.Google Scholar
Kleinhans, M. G., van Dijk, W. M., van de Lageweg, W. I., et al., 2014. Quantifiable effectiveness of experimental scaling of river-and delta morphodynamics and stratigraphy. Earth-Science Reviews, 133, 4361.Google Scholar
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. 10.1002/esp.4437.Google Scholar
Kline, M., Cahoon, B., 2010. Protecting river corridors in Vermont. Journal of the American Water Resources Association, 46(2), 227236. 10.1111/j.1752–1688.2010.00417.x.Google Scholar
Klonsky, L., Vogel, R. M., 2011. Effective measures of “effective” discharge. Journal of Geology, 119(1), 114.Google Scholar
Knaapen, M. A. F., Hulscher, S., De Vriend, H. J., Van Harten, A., 2001. Height and wavelength of alternate bars in rivers: Modelling vs. laboratory experiments. Journal of Hydraulic Research, 39(2), 147153.Google Scholar
Knight, D. W., Shiono, K., 1990. Turbulence measurements in a shear-layer region of a compound channel. Journal of Hydraulic Research, 28(2), 175196.Google Scholar
Knighton, A. D., 1975. Variations in at-a-station hydraulic geometry. American Journal of Science, 275(2), 186218.Google Scholar
Knighton, A. D., 1980. Longitudinal changes in size and sorting of stream-bed material in four English rivers. Geological Society of America Bulletin, 91(1), 5562.Google Scholar
Knighton, A. D., 1982a. Asymmetry of river channel cross-sections. 2. Mode of development and local variation. Earth Surface Processes and Landforms, 7(2), 117131.Google Scholar
Knighton, A. D., 1982b. Longitudinal changes in the size and shape of stream bed material: Evidence of variable transport conditions. Catena, 9, 2534.Google Scholar
Knighton, A. D., 1987. River channel adjustment – the downstream dimension. In: Richards, K. S. (ed.), River Channels: Environment and Process. Blackwell, New York, pp. 95128.Google Scholar
Knighton, A. D., 1989. River adjustment to changes in sediment load – the effects of tin mining on the Ringarooma River, Tasmania, 1875–1984. Earth Surface Processes and Landforms, 14(4), 333359.Google Scholar
Knighton, A. D., 1991. Channel bed adjustment along mine-affected rivers of northeast Tasmania. Geomorphology, 4(3–4), 205219.Google Scholar
Knighton, A. D., 1998. Fluvial Forms and Processes. Arnold, London.Google Scholar
Knighton, A. D., 1999a. Downstream variation in stream power. Geomorphology, 29(3–4), 293306.Google Scholar
Knighton, A. D., 1999b. The gravel-sand transition in a disturbed catchment. Geomorphology, 27(3–4), 325341.Google Scholar
Knighton, A. D., Nanson, G. C., 1993. Anastomosis and the continuum of channel pattern. Earth Surface Processes and Landforms, 18(7), 613625.Google Scholar
Knox, J. C., 1972. Valley alluviation in southwestern Wisconsin. Annals of the Association of American Geographers, 62(3), 401410.Google Scholar
Knox, J. C., 1976. Concept of the graded stream. In: Melhorn, W. N., Flemal, R. C. (eds.), Theories of Landform Development. Binghamton State University of New York, Binghamton, NY, pp. 169198.Google Scholar
Knox, J. C., 1977. Human impacts on Wisconsin stream channels. Annals of the Association of American Geographers, 67(3), 323342.Google Scholar
Knox, J. C., 1987. Historical valley floor sedimentation in the upper Mississippi Valley. Annals of the Association of American Geographers, 77(2), 224244.Google Scholar
Knox, J. C., 1993. Large increases in flood magnitude in response to modest changes in climate. Nature, 361(6411), 430432.Google Scholar
Knox, J. C., 2000. Sensitivity of modern and Holocene floods to climate change. Quaternary Science Reviews, 19(1–5), 439457.Google Scholar
Knox, J. C., 2001. Agricultural influence on landscape sensitivity in the Upper Mississippi River Valley. Catena, 42(2–4), 193224.Google Scholar
Knox, J. C., 2006. Floodplain sedimentation in the Upper Mississippi Valley: Natural versus human accelerated. Geomorphology, 79(3–4), 286310.Google Scholar
Kochel, R. C., 1988. Geomorphic impact of large floods: Review and new perspectives on frequency and magnitude. In: Baker, V. R., Kochel, R. C., Patton, P. C. (eds.), Flood Geomorphology. Wiley, New York, pp. 169187.Google Scholar
Kochel, R. C., Piper, J. F., 1986. Morphology of large valleys on Hawaii – evidence for groundwater sapping and comparisons with Martian valleys. Journal of Geophysical Research – Solid Earth and Planets, 91(B13), E175E192. 10.1029/JB091iB13p0E175.Google Scholar
Kochel, R. C., Howard, A. D., McLane, C. F., 1985. Channel networks developed in fine-grained sediments: analogs to Martian valleys. In: Woldenberg, M. J. (ed.), Models in Geomorphology. Allen and Unwin, London, pp. 313341.Google Scholar
Kochel, R. C., Simmons, D. W., Piper, J. F., 1988. Groundwater sapping experiments in weakly consolidated layered sediments: A summary. In: Howard, A. D., Kochel, R. C., Holt, H. E. (eds.), Sapping Features of the Colorado Plateau: A Comparative Planetary Geology Field Guide. National Aeronautics and Space Administration, Washington, DC, pp. 8493.Google Scholar
Kodama, Y., 1994a. Experimental study of abrasion and its role in producing downstream fining in gravel-bed rivers. Journal of Sedimentary Research, 64(1), 7685.Google Scholar
Kodama, Y., 1994b. Downstream changes in the lithology and grain size of fluvial gravels, the Watarase River, Japan – evidence of the role of abrasion in downstream fining. Journal of Sedimentary Research, 64(1), 6875.Google Scholar
Koiter, A. J., Owens, P. N., Petticrew, E. L., Lobb, D. A., 2013. The behavioural characteristics of sediment properties and their implications for sediment fingerprinting as an approach for identifying sediment sources in river basins. Earth-Science Reviews, 125, 2442.Google Scholar
Komar, P. D., 1983. Shapes of streamlined islands on Earth and Mars – experiments and analyses of the minimum-drag form. Geology, 11(11), 651654.Google Scholar
Komar, P. D., Li, Z., 1986. Pivoting analysis of the selective entrainment of sediments by shape and size with application to gravel threshold. Sedimentology, 33(3), 425436.Google Scholar
Komar, P. D., Li, Z., 1988. Applications of grain-pivoting and sliding analyses to selective entrainment of gravel and to flow-competence evaluations. Sedimentology, 35(4), 681695.Google Scholar
Komar, P. D., Wang, C., 1984. Processes of selective grain transport and the formation of placers on beaches. Journal of Geology, 92(6), 637655.Google Scholar
Kondolf, G. M., 1994. Geomorphic and environmental effects of instream gravel mining. Landscape and Urban Planning, 28(2–3), 225243.Google Scholar
Kondolf, G. M., 1997. Hungry water: effects of dams and gravel mining on river channels. Environmental Management, 21(4), 533551.Google Scholar
Kondolf, G. M., 2000. Assessing salmonid spawning gravel quality. Transactions of the American Fisheries Society, 129(1), 262281.Google Scholar
Kondolf, G. M., 2006. River restoration and meanders. Ecology and Society, 11(2). www.ecologyandsociety.org/vol11/iss12/art42/.Google Scholar
Kondolf, G. M., 2011. Setting goals in river restoration: when and where can the river “heal itself”? In: Simon, A., Bennett, S. J., Castro, J. M. (eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools. American Geophysical Union, Washington, DC, pp. 2943.Google Scholar
Kondolf, G. M., 2012. The Espace de Liberte and restoration of fluvial process: when can the river restore itself and when must we intervene? In: Boon, P. J., Raven, P. J. (eds.), River Conservation and Management. Wiley, Chichester, UK, pp. 225241.Google Scholar
Kondolf, G. M., Lisle, T. E., 2016. Measuring bed sediment. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology. Wiley, Chichester, UK, pp. 278305.Google Scholar
Kondolf, G. M., Matthews, W. V. G., 1991. Unmeasured residuals in sediment budgets – a cautionary note. Water Resources Research, 27(9), 24832486.Google Scholar
Kondolf, G. M., Piegay, H. (eds.), 2016. Tools in Fluvial Geomorphology. Wiley, Chichester, UK.Google Scholar
Kondolf, G. M., Swanson, M. L., 1993. Channel adjustments to reservoir construction and gravel extraction along Stony Creek, California. Environmental Geology, 21(4), 256269.Google Scholar
Kondolf, G. M., Wilcock, P. R., 1996. The flushing flow problem: defining and evaluating objectives. Water Resources Research, 32(8), 25892599.Google Scholar
Kondolf, G. M., Smeltzer, M. W., Railsback, S. F., 2001. Design and performance of a channel reconstruction project in a coastal California gravel-bed stream. Environmental Management, 28(6), 761776.Google Scholar
Kondolf, G. M., Boulton, A. J., O’Daniel, S., et al., 2006. Process-based ecological river restoration: Visualizing three-dimensional connectivity and dynamic vectors to recover lost linkages. Ecology and Society, 11 (2). www.ecologyandsociety.org/vol11/iss12/art15.Google Scholar
Kondolf, G. M., Piegay, H., Schmitt, L., Montgomery, D. R., 2016. Geomorphic classification of rivers and streams. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology, 2nd Edition. Wiley, Chichester, UK, pp. 133158.Google Scholar
Kong, D., Miao, C., Wu, J., et al., 2017. Environmental impact assessments of the Xiaolangdi Reservoir on the most hyperconcentrated laden river, Yellow River, China. Environmental Science and Pollution Research, 24(5), 43374351.Google Scholar
Konrad, C. P., Booth, D. B., Burges, S. J., 2005. Effects of urban development in the Puget Lowland, Washington, on interannual streamflow patterns: Consequences for channel form and streambed disturbance. Water Resources Research, 41(7), 115. 10.1029/2005WR004097.Google Scholar
Konsoer, K. M., Rhoads, B. L., 2014. Spatial-temporal structure of mixing interface turbulence at two large river confluences. Environmental Fluid Mechanics, 14(5), 10431070.Google Scholar
Konsoer, K. M., Rhoads, B. L., Langendoen, E. J., et al., 2016a. Spatial variability in bank resistance to erosion on a large meandering, mixed bedrock-alluvial river. Geomorphology, 252, 8097.Google Scholar
Konsoer, K. M., Rhoads, B. L., Best, J. L., et al., 2016b. Three-dimensional flow structure and bed morphology in large elongate meander loops with different outer bank roughness characteristics. Water Resources Research, 52(12), 96219641. 10.1002/2016wr019040.Google Scholar
Konsoer, K., Rhoads, B., Best, J., et al., 2017. Length scales and statistical characteristics of outer bank roughness for large elongate meander bends: The influence of bank material properties, floodplain vegetation and flow inundation. Earth Surface Processes and Landforms, 42(13), 20242037.Google Scholar
Kooi, H., Beaumont, C., 1996. Large-scale geomorphology: Classical concepts reconciled and integrated with contemporary ideas via a surface processes model. Journal of Geophysical Research – Solid Earth, 101(B2), 33613386.Google 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.Google Scholar
Korup, O., 2012. Earth’s portfolio of extreme sediment transport events. Earth-Science Reviews, 112(3–4), 115125.Google Scholar
Krider, L., Magner, J., Hansen, B., et al., 2017. Improvements in fluvial stability associated with two-stage ditch construction in Mower County, Minnesota. Journal of the American Water Resources Association, 53(4), 886902.Google Scholar
Krumbein, W. C., 1941. The effects of abrasion on the size, shape and roundness of rock fragments. Journal of Geology, 49(5), 482520.Google Scholar
Kuenen, P. H., 1956. Experimental abrasion of pebbles. 2. Rolling by current. Journal of Geology, 64(4), 336368.Google Scholar
Kuhnle, R. A., 1993a. Incipient motion of sand-gravel sediment mixtures. Journal of Hydraulic Engineering, 119(12), 14001415.Google Scholar
Kuhnle, R. A., 1993b. Fluvial transport of sand and gravel mixtures with bimodal size distributions. Sedimentary Geology, 85(1–4), 1724.Google Scholar
Kumar, P., Le, P. V. V., Papanicolaou, A. N. T., et al., 2018. Critical transition in critical zone of intensively managed landscapes. Anthropocene, 22, 1019.Google Scholar
La Barbera, P., Rosso, R., 1989. On the fractal dimension of stream networks. Water Resources Research, 25(4), 735741.Google Scholar
La Barbera, P., Rosso, R., 1990. On the fractal dimension of stream networks – reply. Water Resources Research, 26(9), 22452248.Google Scholar
Lacey, R. W. J., Roy, A. G., 2008. The spatial characterization of turbulence around large roughness elements in a gravel-bed river. Geomorphology, 102(3–4), 542553.Google Scholar
Lackey, R. T., 2001. Values, policy, and ecosystem health. Bioscience, 51(6), 437443.Google Scholar
Lague, D., 2014. The stream power river incision model: Evidence, theory and beyond. Earth Surface Processes and Landforms, 39(1), 3861.Google Scholar
Laity, J. E., Malin, M. C., 1985. Sapping processes and the development of theater-headed valley networks on the Colorado Plateau. Geological Society of America Bulletin, 96(2), 203217.Google Scholar
Lajeunesse, E., Malverti, L., Charru, F., 2010. Bed load transport in turbulent flow at the grain scale: experiments and modeling. Journal of Geophysical Research – Earth Surface, 115. 10.1029/2009jf001628.Google Scholar
Lam, D., Thompson, C., Croke, J., Sharma, A., Macklin, M., 2017. Reducing uncertainty with flood frequency analysis: the contribution of paleoflood and historical flood information. Water Resources Research, 53(3), 23122327. 10.1002/2016wr019959.Google Scholar
Lam, M. Y., Ghidaoui, M. S., Kolyshkin, A. A., 2016. The roll-up and merging of coherent structures in shallow mixing layers. Physics of Fluids, 28(9). 10.1063/1.4960391.Google Scholar
Lamarre, H., Roy, A. G., 2008. The role of morphology on the displacement of particles in a step-pool river system. Geomorphology, 99(1–4), 270279.Google Scholar
Lamarre, H., MacVicar, B., Roy, A. G., 2005. Using passive integrated transponder (PIT) tags to investigate sediment transport in gravel-bed rivers. Journal of Sedimentary Research, 75(4), 736741.Google Scholar
Lamb, M. P., Dietrich, W. E., 2009. The persistence of waterfalls in fractured rock. Geological Society of America Bulletin, 121(7–8), 11231134.Google Scholar
Lamb, M. P., Venditti, J. G., 2016. The grain size gap and abrupt gravel-sand transitions in rivers due to suspension fallout. Geophysical Research Letters, 43(8), 37773785. 10.1002/2016gl068713.Google Scholar
Lamb, M. P., Howard, A. D., Johnson, J., et al., 2006. Can springs cut canyons into rock? Journal of Geophysical Research-Planets, 111(E7), 18. 10.1029/2005je002663.Google Scholar
Lamb, M. P., Dietrich, W. E., Aciego, S. M., DePaolo, D. J., Manga, M., 2008a. Formation of Box Canyon, Idaho, by megaflood: implications for seepage erosion on Earth and Mars. Science, 320(5879), 10671070.Google Scholar
Lamb, M. P., Dietrich, W. E., Sklar, L. S., 2008b. A model for fluvial bedrock incision by impacting suspended and bed load sediment. Journal of Geophysical Research-Earth Surface, 113(F3). 10.1029/2007jf000915.Google Scholar
Lamb, M. P., Dietrich, W. E., Venditti, J. G., 2008c. Is the critical Shields stress for incipient sediment motion dependent on channel-bed slope? Journal of Geophysical Research-Earth Surface, 113(F2). 10.1029/2007jf000831.Google Scholar
Lana-Renault, N., Regues, D., 2009. Seasonal patterns of suspended sediment transport in an abandoned farmland catchment in the Central Spanish Pyrenees. Earth Surface Processes and Landforms, 34(9), 12911301.Google Scholar
Lancaster, S. T., Underwood, E. F., Frueh, W. T., 2010. Sediment reservoirs at mountain stream confluences: dynamics and effects of tributaries dominated by debris-flow and fluvial processes. Geological Society of America Bulletin, 122(11–12), 17751786.Google Scholar
Landemaine, V., Gay, A., Cerdan, O., Salvador-Blanes, S., Rodrigues, S., 2015. Morphological evolution of a rural headwater stream after channelisation. Geomorphology, 230, 125137.Google Scholar
Landwehr, K., Rhoads, B. L., 2003. Depositional response of a headwater stream to channelization, east central Illinois, USA. River Research and Applications, 19(1), 77100.Google Scholar
Lane, E. W., 1954. The Importance of Fluvial Morphology in Hydraulic Engineering. Bureau of Reclamation, Engineering Laboratory Division, Hydraulic Laboratories Report No. 372, U.S. Department of Interior, Denver, CO.Google Scholar
Lane, S. N., 2006. Approaching the system-scale understanding of braided river behaviour. In: Smith, G. H. S., Best, J. L., Bristow, C. S., Petts, G. E. (eds.), Braided Rivers: Process, Deposits, Ecology and Management. Special Publications of the International Association of Sedimentologists, 36, Blackwell, Malden, MA, pp. 107135.Google Scholar
Lane, S. N., Chandler, J. H., Richards, K. S., 1994. Developments in monitoring and modeling small-scale river bed topography. Earth Surface Processes and Landforms, 19(4), 349368.Google Scholar
Lane, S. N., Richards, K. S., Chandler, J. H., 1995. Morphological estimation of the time-integrated bed-load transport rate. Water Resources Research, 31(3), 761772.Google Scholar
Lane, S. N., Westaway, R. M., Hicks, D. M., 2003. Estimation of erosion and deposition volumes in a large, gravel-bed, braided river using synoptic remote sensing. Earth Surface Processes and Landforms, 28(3), 249271.Google Scholar
Lane, S. N., Parsons, D. R., Best, J. L., et al., 2008. Causes of rapid mixing at a junction of two large rivers: Rio Parana and Rio Paraguay, Argentina. Journal of Geophysical Research-Earth Surface, 113(F2). 10.1029/2006jf000745.Google Scholar
Lang, M., Ouarda, T., Bobee, B., 1999. Towards operational guidelines for over-threshold modeling. Journal of Hydrology, 225(3–4), 103117.Google Scholar
Lang, A., Bork, H. R., Mackel, R., et al., 2003. Changes in sediment flux and storage within a fluvial system: Some examples from the Rhine catchment. Hydrological Processes, 17(16), 33213334.Google Scholar
Langbein, W. B., 1949. Annual floods and the partial duration series. Transactions, American Geophysical Union, 30(6), 879881.Google Scholar
Langbein, W. B., 1964. Geometry of river channels. American Society of Civil Engineers Proceedings, 90, 301312.Google Scholar
Langbein, W. B., Leopold, L. B., 1964. Quasi-equilibrium states in channel morphology. American Journal of Science, 262(6), 782794.Google Scholar
Langbein, W. B., Leopold, L. B., 1966. River meanders – theory of minimum variance. U.S. Geological Survey Professional Paper 422-H. U.S. Government Printing Office, Washington, DC.Google Scholar
Langbein, W. B., Schumm, S. A., 1958. Yield of sediment in relation to mean annual precipitation. Transactions, American Geophysical Union, 39(6), 10761084.Google Scholar
Langendoen, E. J., 2011. Application of the CONCEPTS channel evolution model in stream restoration strategies. In: Simon, A., Bennett, S. J., Castro, J. M. (eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools. American Geophysical Union, Washington, DC, pp. 487502.Google Scholar
Langendoen, E. J., Alonso, C. V., 2008. Modeling the evolution of incised streams: I. Model formulation and validation of flow and streambed evolution components. Journal of Hydraulic Engineering, 134(6), 749762.Google Scholar
Lanz, B., Dietz, S., Swanson, T., 2018. The expansion of modern agriculture and global biodiversity decline: An integrated assessment. Ecological Economics, 144, 260277.Google Scholar
Lanzoni, S., 2000a. Experiments on bar formation in a straight flume 1. Uniform sediment. Water Resources Research, 36(11), 33373349.Google Scholar
Lanzoni, S., 2000b. Experiments on bar formation in a straight flume 2. Graded sediment. Water Resources Research, 36(11), 33513363.Google Scholar
Lanzoni, S., Seminara, G., 2006. On the nature of meander instability. Journal of Geophysical Research-Earth Surface, 111(F4). 10.1029/2005jf000416.Google Scholar
Lara Pinto Coelho, M. M., 2015. Experimental determination of free surface levels at open-channel junctions. Journal of Hydraulic Research, 53(3), 394399.Google Scholar
Laraque, A., Guyot, J. L., Filizola, N., 2009. Mixing processes in the Amazon River at the confluences of the Negro and Solimoes Rivers, Encontro das Aguas, Manaus, Brazil. Hydrological Processes, 23(22), 31313140.Google Scholar
Laronne, J. B., Reid, I., Yitshak, Y., Frostick, L. E., 1994. The non-layering of gravel streambeds under ephemeral flood regimes. Journal of Hydrology, 159(1–4), 353363.Google Scholar
Larsen, I. J., Montgomery, D. R., Greenberg, H. M., 2014. The contribution of mountains to global denudation. Geology, 42(6), 527530.Google Scholar
Latosinski, F. G., Szupiany, R. N., Garcia, C. M., Guerrero, M., Amsler, M. L., 2014. Estimation of concentration and load of suspended bed sediment in a large river by means of acoustic Doppler technology. Journal of Hydraulic Engineering, 140(7). 10.1061/(asce)hy.1943–7900.0000859.Google Scholar
Latosinski, F. G., Szupiany, R. N., Guerrero, M., Amsler, M. L., Vionnet, C., 2017. The ADCP’s bottom track capability for bedload prediction: Evidence on method reliability from sandy river applications. Flow Measurement and Instrumentation, 54, 124135.Google Scholar
Latrubesse, E. M., 2008. Patterns of anabranching channels: the ultimate end-member adjustment of mega rivers. Geomorphology, 101(1–2), 130145.Google Scholar
Latrubesse, E. M., Franzinelli, E., 2005. The late Quaternary evolution of the Negro River, Amazon, Brazil: Implications for island and floodplain formation in large anabranching tropical systems. Geomorphology, 70(3–4), 372397.Google Scholar
Latrubesse, E. M., Amsler, M. L., de Morais, R. P., Aquino, S., 2009. The geomorphologic response of a large pristine alluvial river to tremendous deforestation in the South American tropics: the case of the Araguaia River. Geomorphology, 113(3–4), 239252.Google Scholar
Latrubesse, E. M., Arima, E. Y., Dunne, T., et al., 2017. Damming the rivers of the Amazon basin. Nature, 546(7658), 363369.Google Scholar
Lau, J. K., Lauer, T. E., Weinman, M. L., 2006. Impacts of channelization on stream habitats and associated fish assemblages in east central Indiana. American Midland Naturalist, 156(2), 319330.Google Scholar
Lauer, J. W., Parker, G., 2008. Net local removal of floodplain sediment by river meander migration. Geomorphology, 96(1–2), 123149.Google Scholar
Lave, R., 2009. The controversy over natural channel design: substantive explanations and potential avenues for resolution. Journal of the American Water Resources Association, 45(6), 15191532.Google Scholar
Lave, R., 2012. Bridging political ecology and STS: A field analysis of the Rosgen wars. Annals of the Association of American Geographers, 102(2), 366382.Google Scholar
Lave, R., 2014. Freedom and constraint: Generative expectations in the US stream restoration field. Geoforum, 52, 236244.Google Scholar
Lave, R., 2016. Stream restoration and the surprisingly social dynamics of science. Wiley Interdisciplinary Reviews-Water, 3(1), 7581.Google Scholar
Lave, R., 2018. Stream mitigation banking. Wiley Interdisciplinary Reviews-Water, 5(3). 10.1002/wat2.1279.Google Scholar
Lave, R., Doyle, M., Robertson, M., 2010. Privatizing stream restoration in the US. Social Studies of Science, 40(5), 677703.Google Scholar
Lavelle, J. W., Mofjeld, H. O., 1987. Do critical stresses for incipient motion and erosion really exist? Journal of Hydraulic Engineering, 113(3), 370385.Google Scholar
Lawler, D. M., 1993. Needle ice processes and sediment mobilization on river banks – the River Ilston, West-Glamorgan, UK. Journal of Hydrology, 150(1), 81114.Google Scholar
Le Coz, J., Michalkova, M., Hauet, A., et al., 2010. Morphodynamics of the exit of a cutoff meander: Experimental findings from field and laboratory studies. Earth Surface Processes and Landforms, 35(3), 249261.Google Scholar
Lecce, S. A., 1997a. Nonlinear downstream changes in stream power on Wisconsin’s Blue River. Annals of the Association of American Geographers, 87(3), 471486.Google Scholar
Lecce, S. A., 1997b. Spatial patterns of historical overbank sedimentation and floodplain evolution: Blue River, Wisconsin. Geomorphology, 18(3–4), 265277.Google Scholar
Lecce, S. A., Pease, P. P., Gares, P. A., Wang, J. Y., 2006. Seasonal controls on sediment delivery in a small coastal plain watershed, North Carolina, USA. Geomorphology, 73(3–4), 246260.Google Scholar
Lecce, S. A., Pavlowsky, R. T., Bassett, G. S., Martin, D. J., 2011. Metal contamination from gold mining in the Cid District, North Carolina. Physical Geography, 32(5), 469495.Google Scholar
Leclerc, R. F., Hickin, E. J., 1997. The internal structure of scrolled floodplain deposits based on ground-penetrating radar, North Thompson River, British Columbia. Geomorphology, 21(1), 1738.Google Scholar
Leddy, J. O., Ashworth, P. J., Best, J. L., 1993. Mechanisms of anabranch avulsion within gravel-bed braided rivers: observations from a scaled physical model. In: Best, J. L., Bristow, C. S. (eds.), Braided Rivers. Geological Society of London Special Publication No. 75. Geological Society of London, London, pp. 119127.Google Scholar
Lee, A. J., Ferguson, R. I., 2002. Velocity and flow resistance in step-pool streams. Geomorphology, 46(1–2), 5971.Google Scholar
Lee, H. G., Kim, J., 2015. Two-dimensional Kelvin-Helmholtz instabilities of multi-component fluids. European Journal of Mechanics B-Fluids, 49, 7788.Google Scholar
Lee, H. Y., Fu, D. T., Song, M. H., 1993. Migration of rectangular mining pit composed of uniform sediment. Journal of Hydraulic Engineering, 119(1), 6480.Google Scholar
Lee, J.-S., Julien, P. Y., 2006. Downstream hydraulic geometry of alluvial channels. Journal of Hydraulic Engineering, 132(12), 13471352.Google Scholar
Leeder, M. R., Bridges, P. H., 1975. Flow separation in meander bends. Nature, 253(5490), 338339.Google Scholar
Legg, N., Heimburg, C., Collins, B. D., Olsen, P. L., 2014. The Channel Migration Toolbox. Washington State Department of Ecology, Olympia, WA.Google Scholar
Legleiter, C. J., Harrison, L. R., Dunne, T., 2011. Effect of point bar development on the local force balance governing flow in a simple, meandering gravel bed river. Journal of Geophysical Research-Earth Surface, 116, 29. 10.1029/2010jf001838.Google Scholar
Leigh, D. S., 1994. Mercury contamination and floodplain sedimentation from former gold mines in north Georgia. Water Resources Bulletin, 30(4), 739748.Google Scholar
Leigh, D. S., 2008. Late Quaternary climates and river channels of the Atlantic Coastal Plain, Southeastern USA. Geomorphology, 101(1–2), 90108.Google Scholar
Leighly, J., 1932. Toward a theory of the morphological significance of turbulence in the flow of streams. University of California Publications in Geography, 6(1), 122.Google Scholar
Leite Ribeiro, M., Blanckaert, K., Roy, A. G., Schleiss, A. J., 2012a. Flow and sediment dynamics in channel confluences. Journal of Geophysical Research-Earth Surface, 117. 10.1029/2011jf002171.Google Scholar
Leite Ribeiro, M., Blanckaert, K., Roy, A. G., Schleiss, A. J., 2012b. Hydromorphological implications of local tributary widening for river rehabilitation. Water Resources Research, 48. 10.1029/2011wr011296.Google Scholar
Lenzi, M. A., 2001. Step-pool evolution in the Rio Cordon, northeastern Italy. Earth Surface Processes and Landforms, 26(9), 9911008.Google Scholar
Lenzi, M. A., Mao, L., Comiti, F., 2006. Effective discharge for sediment transport in a mountain river: computational approaches and geomorphic effectiveness. Journal of Hydrology, 326(1–4), 257276.Google Scholar
Leopold, L. B., 1953. Downstream change of velocity in rivers. American Journal of Science, 251(8), 606624.Google Scholar
Leopold, L. B., 1973. River channel change with time – an example. Geological Society of America Bulletin, 84(6), 18451860.Google Scholar
Leopold, L. B., 1977. A reverence for rivers. Geology, 5(7), 429430.Google Scholar
Leopold, L. B., 1990. Lag times for small drainage basins. Catena, 18, 157171.Google Scholar
Leopold, L. B., Bull, W. B., 1979. Base level, aggradation, and grade. Proceedings of the American Philosophical Society, 123(3), 168202.Google Scholar
Leopold, L. B., Langbein, W. B., 1962. The concept of entropy in landscape evolution. U.S. Geological Survey Professional Paper 500-A. U.S. Government Printing Office, Washington, DC.Google Scholar
Leopold, L. B., Maddock, T., 1953. The hydraulic geometry of stream channels and some physiographic implications. U.S. Geological Survey Professional Paper 252. U.S. Government Printing Office, Washington, DC.Google Scholar
Leopold, L. B., Miller, J., 1956. Ephemeral streams – hydraulic factors and their relation to the drainage net. U.S. Geological Survey Professional Paper 282-A. U.S. Government Printing Office, Washington, DC.Google Scholar
Leopold, L. B., Wolman, M. G., 1957. River channel patterns: braided, meandering and straight. U.S. Geological Survey Professional Paper 282-B. U.S. Government Printing Office, Washington, DC.Google Scholar
Leopold, L. B., Wolman, M. G., 1960. River meanders. Geological Society of America Bulletin, 71, 769794.Google Scholar
Leopold, L., Wolman, M. G., Miller, J. P., 1964. Fluvial Processes in Geomorphology. Freeman, San Francisco, CA.Google Scholar
Leopold, L. B., Huppman, R., Miller, A., 2005. Geomorphic effects of urbanization in forty-one years of observation. Proceedings of the American Philosophical Society, 149(3), 349371.Google Scholar
Lewin, J., 1976. Initiation of bed forms and meanders in coarse-grained sediment. Geological Society of America Bulletin, 87(2), 281285.Google Scholar
Lewin, J., 1978. Meander development and floodplain sedimentation: a case study from mid-Wales. Geological Journal, 13(1), 2536.Google Scholar
Lewin, J., Ashworth, P. J., 2014a. Defining large river channel patterns: alluvial exchange and plurality. Geomorphology, 215, 8398.Google Scholar
Lewin, J., Ashworth, P. J., 2014b. The negative relief of large river floodplains. Earth-Science Reviews, 129, 123.Google Scholar
Lewin, J., Brewer, P. A., 2001. Predicting channel patterns. Geomorphology, 40(3–4), 329339.Google Scholar
Lewin, J., Brewer, P. A., 2002. Laboratory simulation of clast abrasion. Earth Surface Processes and Landforms, 27(2), 145164.Google Scholar
Lewin, J., Brewer, P. A., 2003. Reply to Van den Berg and Bledsoe’s comment on Lewin and Brewer (2001) “Predicting Channel Patterns”. Geomorphology, 53(3–4), 339342.Google Scholar
Lewin, J., Macklin, M. G., 2010. Floodplain catastrophes in the UK Holocene: messages for managing climate change. Hydrological Processes, 24(20), 29002911.Google Scholar
Lewin, J., Macklin, M. G., 2014. Marking time in geomorphology: should we try to formalise an Anthropocene definition? Earth Surface Processes and Landforms, 39(1), 133137.Google Scholar
Lewin, J., Ashworth, P. J., Strick, R. J. P., 2017. Spillage sedimentation on large river floodplains. Earth Surface Processes and Landforms, 42(2), 290305.Google Scholar
Lewis, G. W., Lewin, J., 1983. Alluvial cutoffs in Wales and the Borderlands. In: Collinson, J.D., Lewin, J. (eds.), Modern and Ancient Fluvial Systems. International Association of Sedimentologists Special Publications 6, Blackwell, Oxford, UK, pp. 145154.Google Scholar
Lewis, Q. W., Rhoads, B. L., 2015. Rates and patterns of thermal mixing at a small stream confluence under variable incoming flow conditions. Hydrological Processes, 29(20), 44424456.Google Scholar
Lewis, Q. W., Rhoads, B. L., 2018. LSPIV measurements of two-dimensional flow structure in streams using small unmanned aerial systems: 2. hydrodynamic mapping at river confluences. Water Resources Research, 54(10), 79817999. 10.1029/2018wr022551.Google Scholar
Lewis, S. L., Maslin, M. A., 2015. Defining the Anthropocene. Nature, 519(7542), 171180.Google Scholar
Li, C., Czapiga, M. J., Eke, E. C., Viparelli, E., Parker, G., 2015a. Variable Shields number model for river bankfull geometry: bankfull shear velocity is viscosity-dependent but grain size-independent. Journal of Hydraulic Research, 53(1), 3648.Google Scholar
Li, J., Bristow, C. S., 2015. Crevasse splay morphodynamics in a dryland river terminus: Rio Colorado in Salar de Uyuni Bolivia. Quaternary International, 377, 7182.Google Scholar
Li, J., Bristow, C. S., Luthi, S. M., Donselaar, M. E., 2015b. Dryland anabranching river morphodynamics: Rio Capilla, Salar de Uyuni, Bolivia. Geomorphology, 250, 282297.Google Scholar
Li, M. Z., Komar, P. D., 1992. Selective entrainment and transport of mixed size and density sands – flume experiments simulating the formation of black-sand placers. Journal of Sedimentary Petrology, 62(4), 584590.Google Scholar
Li, Z., Komar, P. D., 1986. Laboratory measurements of pivoting angles for applications to selective entrainment of gravel in a current. Sedimentology, 33, 413423.Google Scholar
Liebault, F., Bellot, H., Chapuis, M., Klotz, S., Deschatres, M., 2012. Bedload tracing in a high-sediment-load mountain stream. Earth Surface Processes and Landforms, 37(4), 385399.Google Scholar
Limerinos, J. T., 1970. Determination of the Manning coefficient from measured bed roughness in natural channels. U.S. Geological Survey Water-Supply Paper 1898-B. U.S. Government Printing Office, Washington, DC.Google Scholar
Lin, J. D., Soong, H. K., 1979. Junction losses in open channel flows. Water Resources Research, 15, 414418.Google Scholar
Lin, Z., Oguchi, T., 2004. Drainage density, slope angle, and relative basin position in Japanese bare lands from high-resolution DEMs. Geomorphology, 63(3–4), 159173.Google Scholar
Lindow, N., Fox, G. A., Evans, R. O., 2009. Seepage erosion in layered stream bank material. Earth Surface Processes and Landforms, 34(12), 16931701.Google Scholar
Lininger, K. B., Wohl, E., Sutfin, N. A., Rose, J. R., 2017. Floodplain downed wood volumes: a comparison across three biomes. Earth Surface Processes and Landforms, 42(8), 12481261.Google Scholar
Lisenby, P. E., Fryirs, K. A., 2016. Catchment- and reach-scale controls on the distribution and expectation of geomorphic channel adjustment. Water Resources Research, 52(5), 34083427. 10.1002/2015wr017747.Google Scholar
Lisenby, P. E., Croke, J., Fryirs, K. A., 2018. Geomorphic effectiveness: a linear concept in a non-linear world. Earth Surface Processes and Landforms, 43(1), 420.Google Scholar
Lisle, T. E., 1979. A sorting mechanism for a riffle-pool sequence. Geological Society of America Bulletin, Part II, 90(7), 11421157.Google Scholar
Lisle, T. E., 1986. Stabilization of a gravel channel by large streamside obstructions and bedrock bends, Jacoby Creek, northwestern California. Geological Society of America Bulletin, 97(8), 9991011.Google Scholar
Lisle, T. E., Hilton, S., 1992. The volume of fine sediment in pools – an index of sediment supply in gravel-bed streams. Water Resources Bulletin, 28(2), 371383.Google Scholar
Lisle, T. E., Ikeda, H., Iseya, F., 1991. Formation of stationary alternate bars in a steep channel with mixed-size sediment – a flume experiment. Earth Surface Processes and Landforms, 16(5), 463469.Google Scholar
Lisle, T. E., Iseya, F., Ikeda, H., 1993. Response of a channel with alternate bars to a decrease in supply of mixed-size bedload – a flume experiment. Water Resources Research, 29(11), 36233629.Google Scholar
Lisle, T. E., Cui, Y. T., Parker, G., Pizzuto, J. E., Dodd, A. M., 2001. The dominance of dispersion in the evolution of bed material waves in gravel-bed rivers. Earth Surface Processes and Landforms, 26(13), 14091420.Google Scholar
Liu, T.-h., Chen, L., Fan, B.-l., 2012. Experimental study on flow pattern and sediment transportation at a 90° open-channel confluence. International Journal of Sediment Research, 27(2), 178187.Google Scholar
Lobkovsky, A. E., Jensen, B., Kudrolli, A., Rothman, D. H., 2004. Threshold phenomena in erosion driven by subsurface flow. Journal of Geophysical Research-Earth Surface, 109(F4). 10.1029/2004jf000172.Google Scholar
Lobkovsky, A. E., Smith, B. E., Kudrolli, A., Mohrig, D. C., Rothman, D. H., 2007. Erosive dynamics of channels incised by subsurface water flow. Journal of Geophysical Research-Earth Surface, 112(F3). 10.1029/2006jf000517.Google Scholar
Loewenherz, D. S., 1991. Stability and the initiation of channelized surface drainage – a reassessment of the short wavelength limit. Journal of Geophysical Research-Solid Earth and Planets, 96(B5), 84538464.Google Scholar
Loewenherz-Lawrence, D. S., 1994. Hydrodynamic description for advective sediment transport processes and rill initiation. Water Resources Research, 30(11), 32033212.Google Scholar
Lofthouse, C., Robert, A., 2008. Riffle-pool sequences and meander morphology. Geomorphology, 99(1–4), 214223.Google Scholar
Loget, N., Van Den Driessche, J., 2009. Wave train model for knickpoint migration. Geomorphology, 106(3–4), 376382.Google Scholar
Long, D. G. F., 2011. Architecture and depositional style of fluvial systems before land plants: a comparison of Precambrian, early Paleozoic, and modern river deposits. In: Davidson, S. K., Leleu, S., North, C. P. (eds.), From River to Rock Record: The Preservation of Fluvial Sediments and Their Subsequent Interpretation. Society for Sedimentary Geology Special Publication, SEPM, Tulsa, OK, pp. 3761.Google Scholar
Lopez, F., Garcia, M. H., 1999. Wall similarity in turbulent open-channel flow. Journal of Engineering Mechanics, 125(7), 789796.Google Scholar
Lopez-Tarazon, J. A., Estrany, J., 2017. Exploring suspended sediment delivery dynamics of two Mediterranean nested catchments. Hydrological Processes, 31(3), 698715.Google Scholar
Lotsari, E., Vaaja, M., Flener, C., et al., 2014. Annual bank and point bar morphodynamics of a meandering river determined by high-accuracy multitemporal laser scanning and flow data. Water Resources Research, 50(7), 55325559. 10.1002/2013wr014106.Google Scholar
Lotsari, E., Thorndycraft, V., Alho, P., 2015. Prospects and challenges of simulating river channel response to future climate change. Progress in Physical Geography, 39(4), 483513.Google Scholar
Lu, H., Moran, C. J., Sivapalan, M., 2005. A theoretical exploration of catchment-scale sediment delivery. Water Resources Research, 41(9). 10.1029/2005wr004018.Google Scholar
Lu, S. S., Willmarth, W. W., 1973. Measurements of structure of Reynolds stress in a turbulent boundary layer. Journal of Fluid Mechanics, 60(SEP18), 481511.Google Scholar
Lubowe, J. K., 1964. Steam junction angles in the dendritic drainage pattern. American Journal of Science, 262(3), 325339.Google Scholar
Lucas, R. W., Baker, T. T., Wood, M. K., Allison, C. D., VanLeeuwen, D. M., 2009. Streambank morphology and cattle grazing in two montane riparian areas in western New Mexico. Journal of Soil and Water Conservation, 64(3), 183189.Google Scholar
Ludwig, W., Probst, J. L., 1998. River sediment discharge to the oceans: present-day controls and global budgets. American Journal of Science, 298(4), 265295.Google Scholar
Lunt, I. A., Bridge, J. S., Tye, R. S., 2004. A quantitative, three-dimensional depositional model of gravelly braided rivers. Sedimentology, 51(3), 377414.Google Scholar
Luo, H., Fytanidis, D. K., Schmidt, A. R., Garcia, M. H., 2018. Comparative 1D and 3D numerical investigation of open-channel junction flows and energy losses. Advances in Water Resources, 117, 120139. 10.1016/j.advwatres.2018.05.012.Google Scholar
Lyubimova, T., Lepikhin, A., Konovalov, V., Parshakova, Y., Tiunov, A., 2014. Formation of the density currents in the zone of confluence of two rivers. Journal of Hydrology, 508, 328342.Google Scholar
Ma, Y., Huang, H. Q., Xu, J., Brierley, G. J., Yao, Z., 2010. Variability of effective discharge for suspended sediment transport in a large semi-arid river basin. Journal of Hydrology, 388(3–4), 357369.Google Scholar
MacFarlane, W. A., Wohl, E., 2003. Influence of step composition on step geometry and flow resistance in step-pool streams of the Washington Cascades. Water Resources Research, 39(2), 10.1029/2001wr001238.Google Scholar
Mackay, J. R., 1970. Lateral mixing of the Liard and Mackenzie rivers downstream from their confluence. Canadian Journal of Earth Sciences, 7, 111124.Google Scholar
Mackin, J. H., 1948. Concept of the graded river. Geological Society of America Bulletin, 59(5), 463511.Google Scholar
Mackin, J., 1963. Methods of investigation in geology. In: Albritton, C. C. (ed.), The Fabric of Geology. Addison-Wesley, Reading, MA, pp. 135163.Google Scholar
Macklin, M. G., Lewin, J., Woodward, J. C., 2012. The fluvial record of climate change. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 370(1966), 21432172.Google Scholar
Macklin, M. G., Lewin, J., Jones, A. F., 2014. Anthropogenic alluvium: an evidence-based meta-analysis for the UK Holocene. Anthropocene, 6, 2638.Google Scholar
MacVicar, B., Best, J., 2013. A flume experiment on the effect of channel width on the perturbation and recovery of flow in straight pools and riffles with smooth boundaries. Journal of Geophysical Research-Earth Surface, 118(3), 18501863.Google Scholar
MacVicar, B. J., Roy, A. G., 2007a. Hydrodynamics of a forced riffle pool in a gravel bed river: 1. mean velocity and turbulence intensity. Water Resources Research, 43(12). 10.1029/2006wr005272.Google Scholar
MacVicar, B. J., Roy, A. G., 2007b. Hydrodynamics of a forced riffle pool in a gravel bed river: 2. scale and structure of coherent turbulent events. Water Resources Research, 43(12). 10.1029/2006wr005274.Google Scholar
MacVicar, B. J., Roy, A. G., 2011. Sediment mobility in a forced riffle-pool. Geomorphology, 125(3), 445456.Google Scholar
MacWilliams, M. L., Jr., Wheaton, J. M., Pasternack, G. B., Street, R. L., Kitanidis, P. K., 2006. Flow convergence routing hypothesis for pool-riffle maintenance in alluvial rivers. Water Resources Research, 42(10). 10.1029/2005wr004391.Google Scholar
Maddock, T., Jr., 1970. Indeterminate hydraulics of alluvial channels. Journal of the Hydraulics Division – ASCE, 96(HY11), 23092323.Google Scholar
Madej, M. A., Ozaki, V., 1996. Channel response to sediment wave propagation and movement, Redwood Creek, California, USA. Earth Surface Processes and Landforms, 21(10), 911927.Google Scholar
Madej, M. A., Ozaki, V., 2009. Persistence of effects of high sediment loading in a salmon-bearing river, northern California. In: James, L. A., Rathburn, S. L., Whittecar, G. R. (eds.), Management and Restoration of Fluvial Systems with Broad Historical Changes and Human Impacts. Geological Society of America Special Papers, 451, pp. 4355.Google Scholar
Magilligan, F. J., 1985. Historical floodplain sedimentation in the Galena River Basin, Wisconsin and Illinois. Annals of the Association of American Geographers, 75(4), 583594.Google Scholar
Magilligan, F. J., 1992a. Thresholds and the spatial variability of flood power during extreme floods. Geomorphology, 5(3–5), 373390.Google Scholar
Magilligan, F. J., 1992b. Sedimentology of a fine-grained aggrading floodplain. Geomorphology, 4(6), 393408.Google Scholar
Magilligan, F. J., Nislow, K. H., 2005. Changes in hydrologic regime by dams. Geomorphology, 71(1–2), 6178.Google Scholar
Magilligan, F. J., Phillips, J. D., James, L. A., Gomez, B., 1998. Geomorphic and sedimentological controls on the effectiveness of an extreme flood. Journal of Geology, 106(1), 8795.Google Scholar
Magilligan, F. J., Nislow, K. H., Graber, B. E., 2003. Scale-independent assessment of discharge reduction and riparian disconnectivity following flow regulation by dams. Geology, 31(7), 569572.Google Scholar
Magilligan, F. J., Buraas, E. M., Renshaw, C. E., 2015. The efficacy of stream power and flow duration on geomorphic responses to catastrophic flooding. Geomorphology, 228, 175188.Google Scholar
Magilligan, F. J., Graber, B. E., Nislow, K. H., et al., 2016. River restoration by dam removal: enhancing connectivity at watershed scales. Elementa-Science of the Anthropocene, 4, 114.Google Scholar
Mahowald, N. M., Baker, A. R., Bergametti, G., et al., 2005. Atmospheric global dust cycle and iron inputs to the ocean. Global Biogeochemical Cycles, 19(4). 10.1029/2004gb002402.Google Scholar
Major, J. J., O’Connor, J. E., Podolak, C. J., et al., 2012. Geomorphic response of the Sandy River, Oregon, to removal of the Marmot Dam. U.S. Geological Survey Professional Paper 1792, Reston, VA.Google Scholar
Major, J. J., East, A. E., O’Connor, J. E., et al., 2017. Geomorphic responses to dam removal in the United States – a two-decade perspective. In: Tsutsumi, D., Laronne, J. B. (eds.), Gravel-Bed Rivers: Processes and Disasters. Wiley, Chichester, UK, pp. 355381.Google Scholar
Makaske, B., 2001. Anastomosing rivers: a review of their classification, origin and sedimentary products. Earth-Science Reviews, 53(3–4), 149196.Google Scholar
Makaske, B., Smith, D. G., Berendsen, H. J. A., 2002. Avulsions, channel evolution and floodplain sedimentation rates of the anastomosing upper Columbia River, British Columbia, Canada. Sedimentology, 49(5), 10491071.Google Scholar
Makaske, B., Smith, D. G., Berendsen, H. J. A., et al., 2009. Hydraulic and sedimentary processes causing anastomosing morphology of the upper Columbia River, British Columbia, Canada. Geomorphology, 111 (3–4),194205.Google Scholar
Makaske, B., Lavooi, E., de Haas, T., Kleinhans, M. G., Smith, D. G., 2017. Upstream control of river anastomosis by sediment overloading, upper Columbia River, British Columbia, Canada. Sedimentology, 64(6), 14881510.Google Scholar
Malamud, B. D., Turcotte, D. L., 2006. The applicability of power-law frequency statistics to floods. Journal of Hydrology, 322(1–4), 168180.Google Scholar
Maloney, E. M., 2019. How do we take the pulse of an aquatic ecosystem? Current and historical approaches to measuring ecosystem integrity. Environmental Toxicology and Chemistry, 38(2), 289301.Google Scholar
Mandelbrot, B., 1982. The Fractal Geometry of Nature. W.H. Freeman and Company, San Francisco, CA.Google Scholar
Mannering, J. V., Fenster, C. R., 1983. What is conservation tillage? Journal of Soil and Water Conservation, 38(3), 141143.Google Scholar
Mao, L., Picco, L., Lenzi, M. A., Surian, N., 2017. Bed material transport estimate in large gravel-bed rivers using the virtual velocity approach. Earth Surface Processes and Landforms, 42(4), 595611.Google Scholar
Marchi, L., Cavalli, M., Amponsah, W., Borga, M., Crema, S., 2016. Upper limits of flash flood stream power in Europe. Geomorphology, 272, 6877.Google Scholar
Marcinkowski, P., Grabowski, R. C., Okruszko, T., 2017. Controls on anastomosis in lowland river systems: towards process-based solutions to habitat conservation. Science of the Total Environment, 609, 15441555.Google Scholar
Mark, D. M., Church, M., 1980. On size and scale in geomorphology. Progress in Physical Geography, 4, 342390.Google Scholar
Markham, A. J., Thorne, C. R., 1992. Geomorphology of gravel-bed river bends. In: Billi, P., Hey, R. D., Thorne, C. R., Taconi, P. (eds.), Dynamics of Gravel-Bed Rivers. Wiley and Sons, Chichester, UK, pp. 433456.Google Scholar
Maroulis, J. C., Nanson, G. C., 1996. Bedload transport of aggregated muddy alluvium from Cooper Creek, central Australia: a flume study. Sedimentology, 43(5), 771790.Google Scholar
Marra, W. A., Parsons, D. R., Kleinhans, M. G., Keevil, G. M., Thomas, R. E., 2014. Near-bed and surface flow division patterns in experimental river bifurcations. Water Resources Research, 50(2), 15061530.Google Scholar
Marra, W. A., McLelland, S. J., Parsons, D. R., et al., 2015. Groundwater seepage landscapes from distant and local sources in experiments and on Mars. Earth Surface Dynamics, 3(3), 389408.Google Scholar
Marriott, S., 1992. Textural analysis and modeling of a flood deposit – River Severn, UK. Earth Surface Processes and Landforms, 17(7), 687697.Google Scholar
Marsh, W., Dozier, J., 1981. Landscape: An Introduction to Physical Geography. Addison-Wesley, Reading, MA.Google Scholar
Marston, R. A., 1982. The geomorphic significance of log steps in forest streams. Annals of the Association of American Geographers, 72(1), 99108.Google Scholar
Marston, R. A., 1983. Supraglacial stream dynamics on the Juneau Icefield. Annals of the Association of American Geographers, 73(4), 597608.Google Scholar
Martin, R. L., Jerolmack, D. J., Schumer, R., 2012. The physical basis for anomalous diffusion in bed load transport. Journal of Geophysical Research-Earth Surface, 117. 10.1029/2011jf002075.Google Scholar
Martin, Y., Church, M., 1995. Bed-material transport estimated from channel surveys – Vedder River, British Columbia. Earth Surface Processes and Landforms, 20(4), 347361.Google Scholar
Martin, Y., Church, M., 2004. Numerical modelling of landscape evolution: geomorphological perspectives. Progress in Physical Geography, 28(3), 317339.Google Scholar
Martin-Vide, J. P., Ferrer-Boix, C., Ollero, A., 2010. Incision due to gravel mining: modeling a case study from the Gallego River, Spain. Geomorphology, 117(3–4), 261271.Google Scholar
Martin-Vide, J. P., Plana-Casado, A., Sambola, A., Capape, S., 2015. Bedload transport in a river confluence. Geomorphology, 250, 1528.Google Scholar
Masek, J. G., Turcotte, D. L., 1993. A diffusion-limited aggregation model for the evolution of drainage networks. Earth and Planetary Science Letters, 119(3), 379386.Google Scholar
Masteller, C. C., Finnegan, N. J., 2017. Interplay between grain protrusion and sediment entrainment in an experimental flume. Journal of Geophysical Research-Earth Surface, 122(1), 274289. 10.1002/2016jf003943.Google Scholar
Matsubara, Y., Howard, A. D., 2014. Modeling planform evolution of a mud-dominated meandering river: Quinn River, Nevada, USA. Earth Surface Processes and Landforms, 39(10), 13651377.Google Scholar
Matsubara, Y., Howard, A. D., Burr, D. M., et al., 2015. River meandering on Earth and Mars: a comparative study of Aeolis Dorsa meanders, Mars and possible terrestrial analogs of the Usuktuk River, AK, and the Quinn River, NV. Geomorphology, 240, 102120.Google Scholar
Mattingly, R. L., Herricks, E. E., Johnston, D. M., 1993. Channelization and levee construction in Illinois – review and implications for management. Environmental Management, 17(6), 781795.Google Scholar
McCarthy, T. S., Stanistreet, I. G., Cairncross, B., 1991. The sedimentary dynamics of active fluvial channels on the Okavango Fan, Botswana. Sedimentology, 38(3), 471487.Google Scholar
McCarthy, T. S., Ellery, W. N., Stanistreet, I. G., 1992. Avulsion mechanisms on the Okavango Fan, Botswana – the control of a fluvial system by vegetation. Sedimentology, 39(5), 779795.Google Scholar
McCarthy, T. S., Ellery, W. N., Bloem, A., 1998. Some observations on the geomorphological impact of hippopotamus (Hippopotamus amphibius L) in the Okavango Delta, Botswana. African Journal of Ecology, 36(1), 4456.Google Scholar
McCormick, J., 1989. Reclaiming Paradise: The Global Environmental Movement. Indiana University Press, Bloomington, IN.Google Scholar
McDonald, A., Lane, S. N., Haycock, N. E., Chalk, E. A., 2004. Rivers of dreams: on the gulf between theoretical and practical aspects of an upland river restoration. Transactions of the Institute of British Geographers, 29(3), 257281.Google Scholar
McEwan, I., Heald, J., 2001. Discrete particle modeling of entrainment from flat uniformly sized sediment beds. Journal of Hydraulic Engineering, 127(7), 588597.Google Scholar
McGuire, L. A., Pelletier, J. D., 2016. Controls on valley spacing in landscapes subject to rapid base-level fall. Earth Surface Processes and Landforms, 41(4), 460472.Google Scholar
McGuire, L. A., Pelletier, J. D., Gomez, J. A., Nearing, M. A., 2013. Controls on the spacing and geometry of rill networks on hillslopes: rain splash detachment, initial hillslope roughness, and the competition between fluvial and colluvial transport. Journal of Geophysical Research-Earth Surface, 118(1), 241256. 10.1002/jgrf.20028.Google Scholar
McLean, D. G., Church, M., 1999. Sediment transport along lower Fraser River – 2. Estimates based on the long-term gravel budget. Water Resources Research, 35(8), 25492559.Google Scholar
McLean, D. G., Church, M., Tassone, B., 1999. Sediment transport along lower Fraser River – 1. Measurements and hydraulic computations. Water Resources Research, 35(8), 25332548. 10.1029/1999wr900101.Google Scholar
McLelland, S. J., Ashworth, P. J., Best, J. L., 1996. The origin and downstream development of coherent flow structures at channel junctions. In: Ashworth, P. J., Bennett, S. J., Best, J. L., McLelland, S. J. (eds.), Coherent Flow Structures in Open Channels. John Wiley and Sons, Chichester, UK, pp. 459490.Google Scholar
McNamara, J. P., Borden, C., 2004. Observations on the movement of coarse gravel using implanted motion-sensing radio transmitters. Hydrological Processes, 18(10), 18711884.Google Scholar
McNamara, J. P., Ziegler, A. D., Wood, S. H., Vogler, J. B., 2006. Channel head locations with respect to geomorphologic thresholds derived from a digital elevation model: a case study in northern Thailand. Forest Ecology and Management, 224(1–2), 147156.Google Scholar
McParland, D., Eaton, B., Rosenfeld, J., 2016. At-a-station hydraulic geometry simulator. River Research and Applications, 32(3), 399410.Google Scholar
Meade, R. H., 1982. Sources, sinks, and storage of river sediment in the Atlantic Drainage of the United States. Journal of Geology, 90(3), 235252.Google Scholar
Mejia, A. I., Niemann, J. D., 2008. Identification and characterization of dendritic, parallel, pinnate, rectangular, and trellis networks based on deviations from planform self-similarity. Journal of Geophysical Research-Earth Surface, 113(F2), 21. 10.1029/2007jf000781.Google Scholar
Melton, M. A., 1957. An analysis of the relation among elements of climate, surface properties, and geomorphology. Office of Naval Research Technical Report No. 11, Columbia University, New York.Google Scholar
Melton, M. A., 1961. Discussion: the effect of sediment type on the shape and stratification of some modern fluvial deposits. American Journal of Science, 259, 231233.Google Scholar
Merritt, D. M., Wohl, E. E., 2003. Downstream hydraulic geometry and channel adjustment during a flood along an ephemeral, arid-region drainage. Geomorphology, 52(3–4), 165180.Google Scholar
Merz, B., Aerts, J., Arnbjerg-Nielsen, K., et al., 2014. Floods and climate: emerging perspectives for flood risk assessment and management. Natural Hazards and Earth System Sciences, 14(7), 19211942.Google Scholar
Messager, M. L., Lehner, B., Grill, G., Nedeva, I., Schmitt, O., 2016. Estimating the volume and age of water stored in global lakes using a geostatistical approach. Nature Communications, 7. 7:13603 | 10.1038/ncomms13603.Google Scholar
Metivier, F., Gaudemer, Y., 1999. Stability of output fluxes of large rivers in South and East Asia during the last 2 million years: implications on floodplain processes. Basin Research, 11(4), 293303.Google Scholar
Metivier, F., Paola, C., Kozarek, J. L., Tai, M., 2016. Experimental studies and practical challenges in fluvial geomorphology. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology, 2nd Edition. Wiley, London, pp. 454475.Google Scholar
Meybeck, M., 2003. Global analysis of river systems: from Earth system controls to Anthropocene syndromes. Philosophical Transactions of the Royal Society B-Biological Sciences, 358(1440), 19351955.Google Scholar
Meyer, J. L., 1997. Stream health: incorporating the human dimension to advance stream ecology. Journal of the North American Benthological Society, 16(2), 439447.Google Scholar
Meyer-Peter, E., Muller, R., 1948. Formulas for bedload transport. Proceedings of the Second Congress, International Association for Hydraulic Research Stockholm, pp. 3964.Google Scholar
Miall, A. D., 2006. The Geology of Fluvial Deposits. 4th Corrected Edition. Springer Verlag, Berlin.Google Scholar
Miao, C., Borthwick, A. G. L., Liu, H., Liu, J., 2015. China’s policy on dams at the crossroads: removal or further construction? Water, 7(5), 23492357.Google Scholar
Micheli, E. R., Larsen, E. W., 2011. River channel cutoff dynamics, Sacramento River, California, USA. River Research and Applications, 27(3), 328344.Google Scholar
Middleton, G. V., Wilcock, P. R., 1994. Mechanics in the Earth and Environmental Sciences. Cambridge University Press, Cambridge, UK.Google Scholar
Mignot, E., Vinkovic, I., Doppler, D., Riviere, N., 2014. Mixing layer in open-channel junction flows. Environmental Fluid Mechanics, 14(5), 10271041.Google Scholar
Milan, D. J., 2013a. Virtual velocity of tracers in a gravel-bed river using size-based competence duration. Geomorphology, 198, 107114.Google Scholar
Milan, D.J., 2013b. Sediment routing hypothesis for pool-riffle maintenance. Earth Surface Processes and Landforms, 38(14), 16231641.Google Scholar
Milan, D. J., Heritage, G. L., Large, A. R. G., Charlton, M. E., 2001. Stage dependent variability in tractive force distribution through a riffle-pool sequence. Catena, 44(2), 85109.Google Scholar
Milan, D. J., Heritage, G. L., Hetherington, D., 2007. Application of a 3D laser scanner in the assessment of erosion and deposition volumes and channel change in a proglacial river. Earth Surface Processes and Landforms, 32(11), 16571674.Google Scholar
Milan, D. J., Heritage, G. L., Large, A. R. G., Fuller, I. C., 2011. Filtering spatial error from DEMs: implications for morphological change estimation. Geomorphology, 125(1), 160171.Google Scholar
Millar, R.G., 2000. Influence of bank vegetation on alluvial channel patterns. Water Resources Research, 36(4), 11091118.Google Scholar
Millar, R. G., 2005. Theoretical regime equations for mobile gravel-bed rivers with stable banks. Geomorphology, 64(3–4), 207220.Google Scholar
Millar, R. G., Eaton, B. C., 2011. Bank vegetation, bank strength, and application of the University of British Columbia regime model to stream restoration. In: Simon, A., Bennett, S. J., Castro, J. M. (eds.), Stream Restoration in Dynamic Fluvial Systems. American Geophysical Union, Washington, DC, pp. 475485.Google Scholar
Millar, R. G., Quick, M. C., 1993. Effect of bank stability on geometry of gravel rivers. Journal of Hydraulic Engineering, 119(12), 13431363.Google Scholar
Millar, R. G., Quick, M. C., 1998. Stable width and depth of gravel-bed rivers with cohesive banks. Journal of Hydraulic Engineering, 124(10), 10051013.Google 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.Google Scholar
Millennium Ecosystem Assessment, 2005. Ecosystems and Human Well-being: A Framework for Assessment. Island Press, Washington, DC.Google Scholar
Miller, A. J., 1990. Flood hydrology and geomorphic effectiveness in the central Appalachians. Earth Surface Processes and Landforms, 15(2), 119134.Google Scholar
Miller, J. R., 1991b. Development of anastomosing channels in south-central Indiana. Geomorphology, 4(3–4), 221229.Google Scholar
Miller, J. R., 1991c. The influence of bedrock geology on knickpoint development and channel-bed degradation along downcutting streams in south-central Indiana. Journal of Geology, 99(4), 591605.Google Scholar
Miller, J. R., 1997. The role of fluvial geomorphic processes in the dispersal of heavy metals from mine sites. Journal of Geochemical Exploration, 58(2–3),101118.Google Scholar
Miller, J. R., 2017. Causality of historic arroyo incision in the southwestern United States. Anthropocene, 18, 6975. 10.1016/j.ancene.2017.06.003.Google Scholar
Miller, J. R., Kochel, R. C., 2010. Assessment of channel dynamics, in-stream structures and post-project channel adjustments in North Carolina and its implications to effective stream restoration. Environmental Earth Sciences, 59(8), 16811692. 10.1007/s12665-009–0150-1.Google Scholar
Miller, J. R., Ritter, J. B., 1996. An examination of the Rosgen classification of natural rivers. Catena, 27(3–4), 295299. 10.1016/0341–8162(96)00017–3.Google Scholar
Miller, J. R., Lechler, P. J., Desilets, M., 1998. The role of geomorphic processes in the transport and fate of mercury in the Carson River basin, west-central Nevada. Environmental Geology, 33(4), 249262. 10.1007/s002540050244.Google Scholar
Miller, K. L., Szabo, T., Jerolmack, D. J., Domokos, G., 2014. Quantifying the significance of abrasion and selective transport for downstream fluvial grain size evolution. Journal of Geophysical Research-Earth Surface, 119(11), 24122429. 10.1002/2014jf003156.Google Scholar
Miller, M. C., McCave, I. N., Komar, P. D., 1977. Threshold of sediment motion under unidirectional currents. Sedimentology, 24(4), 507527.Google Scholar
Miller, R. L., Byrne, R. J., 1966. The angle of repose for a single grain on a fixed rough bed. Sedimentology, 6, 303314.Google Scholar
Miller, S. O., Ritter, D. F., Kochel, R. C., Miller, J. R., 1993. Fluvial responses to land-use changes and climatic variations within the Drury Creek watershed, southern Illinois. Geomorphology, 6(4), 309329.Google Scholar
Miller, T. K., 1984. A system model of stream-channel shape and size. Geological Society of America Bulletin, 95(2), 237241.Google Scholar
Miller, T. K., 1991a. A model of stream channel adjustment – assessment of Rubey’s hypothesis. Journal of Geology, 99(5), 699710.Google Scholar
Miller, T. K., Onesti, L. J., 1979. Relationship between channel shape and sediment characteristics in the channel perimeter. Geological Society of America Bulletin, 90(3), 301304. 10.1130/0016–7606(1979)90<301:trbcsa>2.0.co;2.Google Scholar
Milliman, J. D., Farnsworth, K. L., 2011. River Discharge to the Coastal Ocean: A Global Synthesis. Cambridge University Press, Cambridge, UK.Google Scholar
Milliman, J. D., Meade, R. H., 1983. World-wide delivery of river sediment to the oceans. Journal of Geology, 91(1), 121.Google Scholar
Milliman, J. D., Syvitski, J. P. M., 1992. Geomorphic/tectonic control of sediment discharge to the ocean – the importance of small mountainous rivers. Journal of Geology, 100(5), 525544.Google Scholar
Milne, J. A., 1982. Bed-material size and the riffle-pool sequence. Sedimentology, 29(2), 267278.Google Scholar
Milzow, C., Molnar, P., McArdell, B. W., Burlando, P., 2006. Spatial organization in the step-pool structure of a steep mountain stream (Vogelbach, Switzerland). Water Resources Research, 42(4). 10.1029/2004WR003870.Google Scholar
Miori, S., Repetto, R., Tubino, M., 2006. A one-dimensional model of bifurcations in gravel bed channels with erodible banks. Water Resources Research, 42(11). 10.1029/2006wr004863.Google Scholar
Miori, S., Hardy, R. J., Lane, S. N., 2012. Topographic forcing of flow partition and flow structures at river bifurcations. Earth Surface Processes and Landforms, 37(6), 666679.Google Scholar
Mohajeri, S. H., Righetti, M., Wharton, G., Romano, G. P., 2016. On the structure of turbulent gravel bed flow: implications for sediment transport. Advances in Water Resources, 92, 90104.Google Scholar
Molnar, P., Anderson, R. S., Kier, G., Rose, J., 2006. Relationships among probability distributions of stream discharges in floods, climate, bed load transport, and river incision. Journal of Geophysical Research-Earth Surface, 111(F2). 10.1029/2005jf000310.Google Scholar
Molnar, P., Densmore, A. L., McArdell, B. W., Turowski, J. M., Burlando, P., 2010. Analysis of changes in the step-pool morphology and channel profile of a steep mountain stream following a large flood. Geomorphology, 124(1–2), 8594.Google Scholar
Montgomery, D. R., 2007a. Dirt: The Erosion of Civilizations. The University of California Press, Berkeley, CA.Google Scholar
Montgomery, D. R., 2007b. Soil erosion and agricultural sustainability. Proceedings of the National Academy of Sciences of the United States of America, 104(33), 1326813272.Google Scholar
Montgomery, D. R., Buffington, J. M., 1997. Channel-reach morphology in mountain drainage basins. Geological Society of America Bulletin, 109(5), 596611.Google Scholar
Montgomery, D. R., Dietrich, W. E., 1988. Where do channels begin? Nature, 336(6196), 232234.Google Scholar
Montgomery, D. R., Dietrich, W. E., 1989. Sources area, drainage density, and channel initiation. Water Resources Research, 25(8), 19071918.Google Scholar
Montgomery, D. R., Dietrich, W. E., 1992. Channel initiation and the problem of landscape scale. Science, 255(5046), 826830.Google Scholar
Montgomery, D.R., Dietrich, W.E., 1994. Landscape dissection and drainage area-slope thresholds. In: Kirkby, M.J. (ed.), Process Models and Theoretical Geomorphology. Wiley, Chichester, UK, pp. 221246.Google Scholar
Montgomery, D. R., Foufoula-Georgiou, E., 1993. Channel network source representation using digital elevation models. Water Resources Research, 29(12), 39253934.Google Scholar
Montgomery, D. R., Gran, K. B., 2001. Downstream variations in the width of bedrock channels. Water Resources Research, 37(6), 18411846.Google Scholar
Montgomery, D. R., Buffington, J. M., Smith, R. D., Schmidt, K. M., Pess, G., 1995. Pool spacing in forest channels. Water Resources Research, 31(4), 10971105.Google Scholar
Moody, J. A., Troutman, B. M., 2000. Quantitative model of the growth of floodplains by vertical accretion. Earth Surface Processes and Landforms, 25(2), 115133.Google Scholar
Moody, J. A., Pizzuto, J. E., Meade, R. H., 1999. Ontogeny of a flood plain. Geological Society of America Bulletin, 111(2), 291303.Google Scholar
Morgan, R. P. C., Quinton, J. N., Smith, R. E., et al., 1998. The European Soil Erosion Model (EUROSEM): a dynamic approach for predicting sediment transport from fields and small catchments. Earth Surface Processes and Landforms, 23(6), 527544.Google Scholar
Morisawa, M. E., 1962. Quantitative geomorphology of some watersheds in the Appalachian Plateau. Geological Society of America Bulletin, 73(9), 10251046.Google Scholar
Morisawa, M., 1964. Development of drainage systems on upraised lake floor. American Journal of Science, 262(3), 340354.Google Scholar
Morisawa, M., 1985. Rivers: Form and Process. Longman, London.Google Scholar
Moron, S., Edmonds, D. A., Amos, K., 2017. The role of floodplain width and alluvial bar growth as a precursor for the formation of anabranching rivers. Geomorphology, 278, 7890.Google Scholar
Morozova, G. S., Smith, N. D., 1999. Holocene avulsion history of the lower Saskatchewan fluvial system, Cumberland Marshes, Saskatchewan-Manitoba, Canada. In: Smith, N. D., Rogers, J. (eds.), Fluvial Sedimentology VI, International Association of Sedimentologists Special Publication No. 28, Blackwell, Malden, MA, pp. 231249.Google Scholar
Morozova, G. S., Smith, N. D., 2000. Holocene avulsion styles and sedimentation patterns of the Saskatchewan River, Cumberland Marshes, Canada. Sedimentary Geology, 130(1–2), 81105.Google Scholar
Morris, G. L., Annandale, G., Hotchkiss, R., 2008. Reservoir sedimentation. In: Garcia, M. H. (ed.), Sedimentation Engineering: Processes, Measurements, Modeling, and Practice. American Society of Civil Engineers, New York, pp. 579612.Google Scholar
Morris, P. H., Williams, D. J., 1999a. Worldwide correlations for subaerial aqueous flows with exponential longitudinal profiles. Earth Surface Processes and Landforms, 24(10), 867879.Google Scholar
Morris, P. H., Williams, D. J., 1999b. A worldwide correlation for exponential bed particle size variation in subaerial aqueous flows. Earth Surface Processes and Landforms, 24(9), 835847.Google Scholar
Mosher, S.-J., Martini, I. P., 2002. Coarse-grained flood bars formed at the confluence of two subarctic rivers affected by hydroelectric dams, Ontario, Canada. In: Martini, I. P., Baker, V. R., Garzon, G. (eds.), Flood and Megaflood Deposits: Recent and Ancient Examples. International Association of Sedimentologists Special Publication 32, Blackwell, Oxford, UK, pp. 213231.Google Scholar
Mosley, M. P., 1976. An experimental study of channel confluences. Journal of Geology, 84, 535562.Google Scholar
Mosley, M. P., 1982. Analysis of the effect of changing discharge on channel morphology and instream uses in a braided river, Ohua River, New Zealand. Water Resources Research, 18(4), 800812.Google Scholar
Moss, J. H., Kochel, R. C., 1978. Unexpected geomorphic effects of Hurricane Agnes storm and flood, Conestoga Drainage Basin, southeastern Pennsylvania. Journal of Geology, 86(1), 111.Google Scholar
Mossa, J., McLean, M., 1997. Channel planform and land cover changes on a mined river floodplain. Applied Geography, 17, 4354.Google Scholar
Mosselman, E., 2005. Basic equations for sediment transport in CFD for fluvial morphodynamics. In: Bates, P. D., Lane, S. N., Ferguson, R. I. (eds.), Computational Fluid Dynamics: Applications in Environmental Hydraulics. Wiley, Chichester, UK, pp. 7189.Google Scholar
Motta, D., Abad, J. D., Langendoen, E. J., Garcia, M. H., 2012a. A simplified 2D model for meander migration with physically-based bank evolution. Geomorphology, 163, 1025.Google Scholar
Motta, D., Abad, J. D., Langendoen, E. J., Garcia, M. H., 2012b. The effects of floodplain soil heterogeneity on meander planform shape. Water Resources Research, 48. 10.1029/2011wr011601.Google Scholar
Motta, D., Langendoen, E. J., Abad, J. D., Garcia, M. H., 2014. Modification of meander migration by bank failures. Journal of Geophysical Research-Earth Surface, 119(5), 10261042. 10.1002/2013jf002952.Google Scholar
Mueller, E. R., Pitlick, J., 2013. Sediment supply and channel morphology in mountain river systems: 1. relative importance of lithology, topography, and climate. Journal of Geophysical Research-Earth Surface, 118(4), 23252342. 10.1002/2013jf002843.Google Scholar
Mueller, E. R., Pitlick, J., 2014. Sediment supply and channel morphology in mountain river systems: 2. single thread to braided transitions. Journal of Geophysical Research-Earth Surface, 119(7), 15161541. 10.1002/2013jf003045.Google Scholar
Mulder, T., Syvitski, J. P. M., 1996. Climatic and morphologic relationships of rivers: implications of sea-level fluctuations on river loads. Journal of Geology, 104(5), 509523.Google Scholar
Murphy, P. J., Randle, T. J., Fotherby, L. M., Daraio, J. A., 2004. The Platte River Channel: History and Restoration. U.S. Department of Interior, Bureau of Reclamation, Denver, CO.Google Scholar
Murray, A. B., Paola, C., 1994. A cellular model of braided rivers. Nature, 371(6492), 5457. 10.1038/371054a0.Google Scholar
Murray, A. B., Paola, C., 1997. Properties of a cellular braided-stream model. Earth Surface Processes and Landforms, 22(11), 10011025.Google Scholar
Murray, A. B., Paola, C., 2003. Modelling the effect of vegetation on channel pattern in bedload rivers. Earth Surface Processes and Landforms, 28(2), 131143.Google Scholar
Muste, M., Yu, K., Fujita, I., Ettema, R., 2005. Two-phase versus mixed-flow perspective on suspended sediment transport in turbulent channel flows. Water Resources Research, 41(10). 10.1029/2004wr003595.Google Scholar
Naden, P. S., 1992. Spatial variability in flood estimation for large catchments – the exploitation of channel network structure. Hydrological Sciences Journal, 37(1), 5371.Google Scholar
Naiman, R. J., Turner, M. G., 2000. A future perspective on North America’s freshwater ecosystems. Ecological Applications, 10(4), 958970.Google Scholar
Nakato, T., 1990. Tests of selected sediment transport formulas. Journal of Hydraulic Engineering, 116(3), 362379.Google Scholar
Nanson, G. C., 1980. Point bar and floodplain formation of the meandering Beatton River, northeastern British Columbia, Canada. Sedimentology, 27(1), 329.Google Scholar
Nanson, G. C., 1986. Episodes of vertical accretion and catastrophic stripping – a model of disequilibrium floodplain development. Geological Society of America Bulletin, 97(12), 14671475.Google Scholar
Nanson, G., 2013. Anabranching and anastomosing rivers. In: Schroder, J. (ed.), Treatise on Geomorphology, Vol. 9, Fluvial Geomorphology, Wohl, E. (vol. ed.). Academic Press, San Diego, CA, pp. 330345.Google Scholar
Nanson, G. C., Croke, J. C., 1992. A genetic classification of floodplains. Geomorphology, 4(6), 459486.Google Scholar
Nanson, G. C., Hickin, E. J., 1983. Channel migration and incision on the Beatton River. Journal of Hydraulic Engineering, ASCE, 109(3), 327337.Google Scholar
Nanson, G. C., Hickin, E. J., 1986. A statistical analysis of bank erosion and channel migration in western Canada. Geological Society of America Bulletin, 97(4), 497504.Google Scholar
Nanson, G. C., Huang, H. Q., 2008. Least action principle, equilibrium states, iterative adjustment and the stability of alluvial channels. Earth Surface Processes and Landforms, 33(6), 923942.Google Scholar
Nanson, G. C., Huang, H. Q., 2017. Self-adjustment in rivers: evidence for least action as the primary control of alluvial-channel form and process. Earth Surface Processes and Landforms, 42(4), 575594.Google Scholar
Nanson, G. C., Huang, H. Q., 2018. A philosophy of rivers: equilibrium states, channel evolution, teleomatic change and least action principle. Geomorphology, 302, 319.Google Scholar
Nanson, G. C., Knighton, A. D., 1996. Anabranching rivers: their cause, character and classification. Earth Surface Processes and Landforms, 21(3), 217239.Google Scholar
Nanson, G. C., Page, K., 1983. Lateral accretion of fine-grained concave benches on meandering rivers. Special Publication of the International Association of Sedimentologists, 6, 133143.Google Scholar
Nanson, G. C., Rust, B. R., Taylor, G., 1986. Coexistent mud braids and anastomosing channels in an arid-zone river – Cooper Creek, central Australia. Geology, 14(2), 175178.Google Scholar
Nanson, R. A., 2010. Flow fields in tightly curving meander bends of low width-depth ratio. Earth Surface Processes and Landforms, 35(2), 119135. 10.1002/esp.1878.Google Scholar
Naot, D., Rodi, W., 1982. Calculation of secondary currents in channel flow. Journal of the Hydraulics Division – ASCE, 108, 948968.Google Scholar
Naqshband, S., McElroy, B., Mahon, R. C., 2017. Validating a universal model of particle transport lengths with laboratory measurements of suspended grain motions. Water Resources Research, 53(5), 41064123. 10.1002/2016wr020024.Google Scholar
Nash, D. B., 1994. Effective sediment-transporting discharge from magnitude-frequency analysis. Journal of Geology, 102(1), 7995.Google Scholar
National Research Council, 1992. Restoration of Aquatic Ecosystems. National Academy Press, Washington, DC.Google Scholar
Navratil, O., Albert, M. B., 2010. Non-linearity of reach hydraulic geometry relations. Journal of Hydrology, 388(3–4), 280290.Google Scholar
Navratil, O., Albert, M. B., Herouin, E., Gresillon, J. M., 2006. Determination of bankfull discharge magnitude and frequency: comparison of methods on 16 gravel-bed river reaches. Earth Surface Processes and Landforms, 31(11), 13451363.Google Scholar
Navratil, O., Breil, P., Schmitt, L., Grosprêtre, L., Albert, M. B., 2013. Hydrogeomorphic adjustments of stream channels disturbed by urban runoff (Yzeron River basin, France). Journal of Hydrology, 485, 2436.Google Scholar
Naylor, L. A., Spencer, T., Lane, S. N., et al., 2017. Stormy geomorphology: geomorphic contributions in an age of climate extremes. Earth Surface Processes and Landforms, 42(1), 166190.Google Scholar
Nazari-Giglou, A., Jabbari-Sahebari, A., Shakibaeinia, A., Borghei, S. M., 2016. An experimental study of sediment transport in channel confluences. International Journal of Sediment Research, 31(1), 8796.Google Scholar
Nearing, M. A., Nichols, M. H., Stone, J. J., Renard, K. G., Simanton, J. R., 2007. Sediment yields from unit-source semiarid watersheds at Walnut Gulch. Water Resources Research, 43(6). 10.1029/2006wr005692.Google Scholar
Nearing, M. A., Xie, Y., Liu, B., Ye, Y., 2017. Natural and anthropogenic rates of soil erosion. International Soil and Water Conservation Research, 5(2), 7784.Google Scholar
Neely, A. B., Bookhagen, B., Burbank, D. W., 2017. An automated knickzone selection algorithm (KZ-Picker) to analyze transient landscapes: calibration and validation. Journal of Geophysical Research – Earth Surface, 122(6), 12361261. 10.1002/2017jf004250.Google Scholar
Neitsch, S. L., Arnold, J. G., Kiniry, J. R., Williams, J. R., 2009. Soil and Water Assessment Tool: Theoretical Documentation Version 2009. Texas Water Resources Institute Technical Report No. 406, Texas A&M University College Station, Texas.Google Scholar
Nelson, A. D., Church, M., 2012. Placer mining along the Fraser River, British Columbia: the geomorphic impact. Geological Society of America Bulletin, 124(7–8), 12121228.Google Scholar
Nelson, J. M., 1990. The initial instability and finite-amplitude stability of alternate bars in straight channels. Earth-Science Reviews, 29(1–4), 97115.Google Scholar
Nelson, J. M., Smith, J. D., 1989. Evolution and stability of erodible channel beds. In: Ikeda, S., Parker, G. (eds.), River Meandering, Water Resources Monograph 12. American Geophysical Union, Washington, DC, pp. 321377.Google Scholar
Nelson, J. M., Shreve, R. L., McLean, S. R., Drake, T. G., 1995. Role of near-bed turbulence structure in bed load transport and bed form mechanics. Water Resources Research, 31(8), 20712086.Google Scholar
Nelson, J. M., McDonald, R. R., Shimizu, Y., et al., 2016. Modelling flow, sediment transport and morphodynamics in rivers. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology, 2nd Edition. Wiley, Chichester, UK, pp. 412441.Google Scholar
Nelson, P. A., Smith, J. A., Miller, A. J., 2006. Evolution of channel morphology and hydrologic response in an urbanizing drainage basin. Earth Surface Processes and Landforms, 31(9), 10631079. 10.1002/esp.1308.Google Scholar
Newson, M. D., Large, A. R. G., 2006. “Natural” rivers, “hydromorphological quality” and river restoration: a challenging new agenda for applied fluvial geomorphology. Earth Surface Processes and Landforms, 31(13), 16061624.Google Scholar
Nezu, I., Nakagawa, H., 1993. Turbulence in Open-Channel Flows. Balkema, Rotterdam.Google Scholar
Nezu, I., Nakagawa, H., Rodi, W., 1989. Significant difference between secondary currents in closed channels and narrow open channels, Proceedings of the 28th Congress, International Association of Hydraulic Research, Ottawa, Canada, pp. A-125–A132.Google Scholar
Nezu, I., Tominaga, A., Nakagawa, H., 1993. Field measurements of secondary currents in straight rivers. Journal of Hydraulic Engineering, 119(5), 598614.Google Scholar
Nicholas, A., 2013. Morphodynamic diversity of the world’s largest rivers. Geology, 41(4), 475478.Google Scholar
Nicholas, A. P., Mitchell, C. A., 2003. Numerical simulation of overbank processes in topographically complex floodplain environments. Hydrological Processes, 17(4), 727746.Google Scholar
Nicholas, A. P., Sambrook Smith, G. H., 1999. Numerical simulation of three-dimensional flow hydraulics in a braided channel. Hydrological Processes, 13(6), 913929.Google Scholar
Nicholas, A. P., Walling, D. E., 1997. Modelling flood hydraulics and overbank deposition on river floodplains. Earth Surface Processes and Landforms, 22(1), 5977.Google Scholar
Nicholas, A. P., Walling, D. E., 1998. Numerical modelling of floodplain hydraulics and suspended sediment transport and deposition. Hydrological Processes, 12(8), 13391355.Google Scholar
Nicholas, A. P., Thomas, R., Quine, T. A., 2006a. Cellular modelling of braided river form and process. In: Smith, G. H. S., Best, J. L., Bristow, C. S., Petts, G. E. (eds.), Braided Rivers: Process, Deposits, Ecology and Management. Special Publications of the International Association of Sedimentologists, 36, Blackwell, Malden, MA, pp. 137151.Google Scholar
Nicholas, A. P., Walling, D. E., Sweet, R. J., Fang, X., 2006b. Development and evaluation of a new catchment-scale model of floodplain sedimentation. Water Resources Research, 42(10), 13. 10.1029/2005wr004579.Google Scholar
Nicholas, A. P., Ashworth, P. J., Smith, G. H. S., Sandbach, S. D., 2013. Numerical simulation of bar and island morphodynamics in anabranching megarivers. Journal of Geophysical Research-Earth Surface, 118(4), 20192044. 10.1002/jgrf.20132.Google Scholar
Nichols, A. L., Viers, J. H., 2017. Not all breaks are equal: variable hydrologic and geomorphic responses to intentional levee breaches along the lower Cosumnes River, California. River Research and Applications, 33(7), 11431155.Google Scholar
Nichols, M. H., Nearing, M. A., Polyakov, V. O., Stone, J. J., 2013. A sediment budget for a small semiarid watershed in southeastern Arizona, USA. Geomorphology, 180, 137145.Google Scholar
Nichols, M. H., Nearing, M., Hernandez, M., Polyakov, V. O., 2016. Monitoring channel head erosion processes in response to an artificially induced abrupt base level change using time-lapse photography. Geomorphology, 265, 107116.Google Scholar
Nickolotsky, A., Pavlowsky, R. T., 2007. Morphology of step-pools in a wilderness headwater stream: the importance of standardizing geomorphic measurements. Geomorphology, 83(3–4), 294306.Google Scholar
Niemann, J. D., Bras, R. L., Veneziano, D., Rinaldo, A., 2001. Impacts of surface elevation on the growth and scaling properties of simulated river networks. Geomorphology, 40(1–2), 3755.Google Scholar
Nienhuis, J. H., Tornqvist, T. E., Esposito, C. R., 2018. Crevasse splays versus avulsions: a recipe for land building with levee breaches. Geophysical Research Letters, 45(9), 40584067.Google Scholar
Niiniluoto, I., 1993. The aim and structure of applied research. Erkenntnis, 38, 121.Google Scholar
Niiniluoto, I., 2014. Values in design sciences. Studies in History and Philosophy of Science, 46, 1115.Google Scholar
Nijssen, B., O’Donnell, G. M., Hamlet, A. F., Lettenmaier, D. P., 2001. Hydrologic sensitivity of global rivers to climate change. Climatic Change, 50(1–2), 143175.Google Scholar
Nikora, V., Goring, D., 2000. Flow turbulence over fixed and weakly mobile gravel beds. Journal of Hydraulic Engineering, 126(9), 679690.Google Scholar
Nikora, V., Habersack, H., Huber, T., McEwan, I., 2002. On bed particle diffusion in gravel bed flows under weak bed load transport. Water Resources Research, 38(6). 10.1029/2001wr000513.Google Scholar
Nilsson, C., Reidy, C. A., Dynesius, M., Revenga, C., 2005. Fragmentation and flow regulation of the world’s large river systems. Science, 308(5720), 405408.Google Scholar
Nino, Y., Garcia, M., 1998. Using Lagrangian particle saltation observations for bedload sediment transport modelling. Hydrological Processes, 12(8), 11971218.Google Scholar
Nino, Y., Garcia, M., Ayala, L., 1994. Gravel saltation: 1. experiments. Water Resources Research, 30(6), 19071914.Google Scholar
Nino, Y., Lopez, F., Garcia, M., 2003. Threshold for particle entrainment into suspension. Sedimentology, 50(2), 247263.Google Scholar
Noble, C. A., Palmquist, R. C., 1968. Meander growth in artificially straightened streams. Proceedings of the Iowa Academy of Science, 75, 234242.Google Scholar
Nolan, K. M., Kelsey, H. M., Marron, D. C. (eds.), 1995. Geomorphic processes and aquatic habitat in the Redwood Creek basin, northwestern California. U.S. Geological Survey Professional Paper 1454. U.S. Government Printing Office, Washington, DC.Google Scholar
Notebaert, B., Broothaerts, N., Verstraeten, G., 2018. Evidence of anthropogenic tipping points in fluvial dynamics in Europe. Global and Planetary Change, 164, 2738.Google Scholar
Nunnally, N. R., 1978. Stream renovation – an alternative to channelization. Environmental Management, 2(5), 403411.Google Scholar
Nykanen, D. K., Foufoula-Georgiou, E., Sapozhnikov, V. B., 1998. Study of spatial scaling in braided river patterns using synthetic aperture radar imagery. Water Resources Research, 34(7), 17951807.Google Scholar
O’Brien, P. E., Wells, A. T., 1986. A small, alluvial crevasse splay. Journal of Sedimentary Petrology, 56(6), 876879.Google Scholar
Odoni, N. A., Lane, S. N., 2011. The significance of models in geomorphology: from concepts to experiments. In: Gregory, K. J., Goudie, A. S. (eds.), The Sage Handbook on Geomorphology. Sage, Los Angeles, CA, pp. 154173.Google Scholar
Oeurng, C., Sauvage, S., Sanchez-Perez, J.-M., 2010. Dynamics of suspended sediment transport and yield in a large agricultural catchment, southwest France. Earth Surface Processes and Landforms, 35(11), 12891301.Google Scholar
Ohmori, H., 1991. Change in the mathematical function type describing the longitudinal profile of a river through an evolutionary process. Journal of Geology, 99(1), 97110.Google Scholar
Onda, Y., 1994. Seepage erosion and its implication to the formation of amphitheater valley heads – a case study at Obara, Japan. Earth Surface Processes and Landforms, 19(7), 627640.Google Scholar
O’Neill, M. P., Abrahams, A. D., 1984. Objective identification of pools and riffles. Water Resources Research, 20(7), 921926.Google Scholar
Orme, A. R., 2013. The scientific roots of geomorphology before 1830. In: Shroder, J. W. (ed.), Treatise on Geomorphology, Vol. 1, The Foundations of Geomorphology, Orme, A.R., Sack, D. (vol. eds.). Elsevier, New York, pp. 1136.Google Scholar
Osman, A. M., Thorne, C. R., 1988. Riverbank stability analysis. 1: theory. Journal of Hydraulic Engineering, 114(2), 134150.Google Scholar
Osterkamp, W. R., Costa, J. E., 1987. Changes accompanying an extraordinary flood on a sand-bed stream. In: Mayer, L., Nash, D. B. (eds.), Catastrophic Flooding. Allen and Unwin, London, pp. 201224.Google Scholar
Ottevanger, W., Blanckaert, K., Uijttewaal, W. S. J., 2012. Processes governing the flow redistribution in sharp river bends. Geomorphology, 163, 4555.Google Scholar
Ottevanger, W., Blanckaert, K., Uijttewaal, W. S. J., de Vriend, H. J., 2013. Meander dynamics: a reduced-order nonlinear model without curvature restrictions for flow and bed morphology. Journal of Geophysical Research – Earth Surface, 118(2), 11181131. 10.1002/jgrf.20080.Google Scholar
Oudin, L., Salavati, B., Furusho-Percot, C., Ribstein, P., Saadi, M., 2018. Hydrological impacts of urbanization at the catchment scale. Journal of Hydrology, 559, 774786.Google Scholar
Owen, L. A., 2013. Tectonic geomorphology: a perspective. In: Schroder, J. W. (ed.), Treatise on Geomorphology, Vol. 5, Tectonic Geomorphology, Owen, J. A. (vol. ed.). Elsevier, New York, pp. 412.Google Scholar
Owens, P. N., Batalla, R. J., Collins, A. J., et al., 2005. Fine-grained sediment in river systems: environmental significance and management issues. River Research and Applications, 21(7), 693717.Google Scholar
Page, K., Nanson, G., 1982. Concave-bank benches and associated floodplain formation. Earth Surface Processes and Landforms, 7(6), 529543.Google Scholar
Page, K. J., Nanson, G. C., Frazier, P. S., 2003. Floodplain formation and sediment stratigraphy resulting from oblique accretion on the Murrumbidgee River, Australia. Journal of Sedimentary Research, 73(1), 514.Google Scholar
Paiement-Paradis, G., Marquis, G., Roy, A., 2011. Effects of turbulence on the transport of individual particles as bedload in a gravel-bed river. Earth Surface Processes and Landforms, 36(1), 107116.Google Scholar
Pal, D., Ghoshal, K., 2016. Effect of particle concentration on sediment and turbulent diffusion coefficients in open-channel turbulent flow. Environmental Earth Sciences, 75(18). 10.1007/s12665-016-6045-z.Google Scholar
Palmer, M. A., Hondula, K. L., 2014. Restoration as mitigation: analysis of stream mitigation for coal mining impacts in Southern Appalachia. Environmental Science & Technology, 48(18), 1055210560.Google Scholar
Palmer, M. A., Bernhardt, E. S., Allan, J. D., et al., 2005. Standards for ecologically successful river restoration. Journal of Applied Ecology, 42(2), 208217.Google Scholar
Palmer, M. A., Bernhardt, E. S., Schlesinger, W. H., et al., 2010a. Mountaintop mining consequences. Science, 327(5962), 148149.Google Scholar
Palmer, M. A., Menninger, H. L., Bernhardt, E., 2010b. River restoration, habitat heterogeneity and biodiversity: a failure of theory or practice? Freshwater Biology, 55, 205222.Google Scholar
Palmer, M. A., Hondula, K. L., Koch, B. J., 2014. Ecological restoration of streams and rivers: shifting strategies and shifting goals. In: Futuyma, D. J. (ed.), Annual Review of Ecology, Evolution, and Systematics, 45, 247269.Google Scholar
Paola, C., 2001. Modelling stream braiding over a range of scales. In: Mosley, M. P. (ed.), Gravel Bed Rivers V. New Zealand Hydrological Society, Wellington, pp. 1146.Google Scholar
Paola, C., Seal, R., 1995. Grain-size patchiness as a cause of selective deposition and downstream fining. Water Resources Research, 31(5), 13951407.Google Scholar
Paola, C., Parker, G., Seal, R., et al., 1992. Downstream fining by selective deposition in a laboratory flume. Science, 258(5089), 17571760.Google Scholar
Paola, C., Straub, K., Mohrig, D., Reinhardt, L., 2009. The “unreasonable effectiveness” of stratigraphic and geomorphic experiments. Earth-Science Reviews, 97(1–4), 143.Google Scholar
Papangelakis, E., Hassan, M. A., 2016. The role of channel morphology on the mobility and dispersion of bed sediment in a small gravel-bed stream. Earth Surface Processes and Landforms, 41(15), 21912206.Google Scholar
Papanicolaou, A. N., Diplas, P., Dancey, C. L., Balakrishnan, M., 2001. Surface roughness effects in near-bed turbulence: implications to sediment entrainment. Journal of Engineering Mechanics, 127(3), 211218.Google Scholar
Park, C. C., 1977. World-wide variations in hydraulic geometry exponents of stream channels – analysis and some observations. Journal of Hydrology, 33(1–2), 133146.Google Scholar
Park, C. C., 1981. Man, river systems, and environmental impact. Progress in Physical Geography, 5, 131.Google Scholar
Park, E., Latrubesse, E. M., 2015. Surface water types and sediment distribution patterns at the confluence of mega rivers: the Solimoes-Amazon and Negro Rivers junction. Water Resources Research, 51(8), 61976213. 10.1002/2014wr016757.Google Scholar
Parker, C., Clifford, N. J., Thorne, C. R., 2011. Understanding the influence of slope on the threshold of coarse grain motion: revisiting critical stream power. Geomorphology, 126(1–2), 5165. 10.1016/j.geomorph.2010.10.027.Google Scholar
Parker, G., 1976. Cause and characteristic scales of meandering and braiding in rivers. Journal of Fluid Mechanics, 76(AUG11), 457480.Google Scholar
Parker, G., 1978a. Self-formed straight rivers with equilibrium banks and mobile bed. Part 1. The sand-silt river. Journal of Fluid Mechanics, 89(NOV), 109125. 10.1017/s0022112078002499.Google Scholar
Parker, G., 1978b. Self-formed straight rivers with equilibrium banks and mobile bed. Part 2. The gravel river. Journal of Fluid Mechanics, 89(NOV), 127146.Google Scholar
Parker, G., 1979. Hydraulic geometry of active gravel rivers. Journal of the Hydraulics Division – ASCE, 105(9), 11851201.Google Scholar
Parker, G., 1990. Surface-based bedload transport relation for gravel rivers. Journal of Hydraulic Research, 28(4), 417436.Google Scholar
Parker, G., 1991a. Selective sorting and abrasion of river gravel. 1. Theory. Journal of Hydraulic Engineering, 117(2), 131149.Google Scholar
Parker, G., 1991b. Selective sorting and abrasion of river gravel. 2. Applications. Journal of Hydraulic Engineering, 117(2), 150171.Google Scholar
Parker, G., 2008. Transport of gravel and sediment mixtures. In: Garcia, M. H. (ed.), Sedimentation Engineering: Processes, Measurements, Modeling, and Practice. American Society of Civil Engineers, Reston, VA, pp. 165251.Google Scholar
Parker, G., Andres, D., 1976. Detrimental effects of river channelization, Rivers ’76. American Society of Civil Engineers, New York, pp. 12481266.Google Scholar
Parker, G., Andrews, E. D., 1986. On the time development of meander bends. Journal of Fluid Mechanics, 162, 139156. 10.1017/s0022112086001970.Google Scholar
Parker, G., Cui, Y. T., 1998. The arrested gravel front: stable gravel-sand transitions in rivers – Part 1: simplified analytical solution. Journal of Hydraulic Research, 36(1), 75100.Google Scholar
Parker, G., Johannesson, H., 1989. Observations on some recent theories of resonance and overdeepening in meandering channels. In: Ikeda, S., Parker, G. (eds.), River Meandering. Water Resources Monograph 12. American Geophysical Union, Washington, DC, pp. 379415.Google Scholar
Parker, G., Klingeman, P. C., 1982. On why gravel bed streams are paved. Water Resources Research, 18(5), 14091423. 10.1029/WR018i005p01409.Google Scholar
Parker, G., Sutherland, A. J., 1990. Fluvial armor. Journal of Hydraulic Research, 28(5), 529544.Google Scholar
Parker, G., Toro-Escobar, C. M., 2002. Equal mobility of gravel in streams: the remains of the day. Water Resources Research, 38 (11),46/41–46/48. 10.1029/2001wr000669.Google Scholar
Parker, G., Klingeman, P. C., McLean, D. G., 1982a. Bedload and size distribution in paved gravel-bed streams. Journal of the Hydraulics Division-ASCE, 108(4), 544571.Google Scholar
Parker, G., Dhamotharan, S., Stefan, H., 1982b. Model experiments on mobile, paved gravel bed streams. Water Resources Research, 18(5), 13951408.Google Scholar
Parker, G., Sawai, K., Ikeda, S., 1982c. Bend theory of river meanders. 2. Nonlinear deformation of finite-amplitude bends. Journal of Fluid Mechanics, 115(FEB), 303314.Google Scholar
Parker, G., Diplas, P., Akiyama, J., 1983. Meander bends of high amplitude. Journal of Hydraulic Engineering, 109(10), 13231337.Google Scholar
Parker, G., Wilcock, P. R., Paola, C., Dietrich, W. E., Pitlick, J., 2007. Physical basis for quasi-universal relations describing bankfull hydraulic geometry of single-thread gravel bed rivers. Journal of Geophysical Research-Earth Surface, 112(F4). 10.1029/2006jf000549.Google Scholar
Parker, G., Hassan, M. A., Wilcock, P. R., 2008. Adjustment of the bed surface size distribution of gravel-bed rivers in response to cycled hydrographs. In: Habersack, H., Piegay, H., Rinaldi, M. (eds.), Gravel-bed Rivers VI: From Process Understanding to River Restoration. Elsevier, Amsterdam, the Netherlands, pp. 241285.Google Scholar
Parker, G., Shimizu, Y., Wilkerson, G. V., et al., 2011. A new framework for modeling the migration of meandering rivers. Earth Surface Processes and Landforms, 36(1), 7086.Google Scholar
Parker, R. S., Schumm, S. A., 1982. Experimental study of drainage networks. In: Bryan, R. B., Yair, A. (eds.), Badland: Geomorphology and Piping. Geobooks, Norwich, UK, pp. 153168.Google Scholar
Parsons, A. J., 2012. How useful are catchment sediment budgets? Progress in Physical Geography, 36(1), 6071.Google Scholar
Parsons, A. J., Wainwright, J., Brazier, R. E., Powell, D. M., 2006. Is sediment delivery a fallacy? Earth Surface Processes and Landforms, 31(10), 13251328.Google Scholar
Parsons, A. J., Wainwright, J., Brazier, R. E., Powell, D. M., 2008a. Is sediment delivery a fallacy? Reply. Earth Surface Processes and Landforms, 33(10), 16301631.Google Scholar
Parsons, D. R., Best, J. L., Orfeo, O., et al., 2005. Morphology and flow fields of three-dimensional dunes, Rio Parana, Argentina: results from simultaneous multibeam echo sounding and acoustic Doppler current profiling. Journal of Geophysical Research-Earth Surface, 110(F4). 10.1029/2004jf000231.Google Scholar
Parsons, D. R., Best, J. L., Lane, S. N., et al., 2007. Form roughness and the absence of secondary flow in a large confluence-diffluence, Rio Parana, Argentina. Earth Surface Processes and Landforms, 32(1), 155162.Google Scholar
Parsons, D. R., Best, J. L., Lane, S. N., et al., 2008b. Large river channel confluences. In: Rice, S. P., Roy, A. G., Rhoads, B. L. (eds.), River Confluences, Tributaries and the Fluvial Network. Wiley, Chichester, UK, pp. 7391.Google Scholar
Pasternack, G. B., Ellis, C. R., Marr, J. D., 2007. Jet and hydraulic jump near-bed stresses below a horseshoe waterfall. Water Resources Research, 43(7), 14. 10.1029/2006wr005774.Google Scholar
Pasternack, G. B., Bounrisavong, M. K., Parikh, K. K., 2008. Backwater control on riffle-pool hydraulics, fish habitat quality, and sediment transport regime in gravel-bed rivers. Journal of Hydrology, 357(1–2), 125139.Google Scholar
Patil, S., Sivapalan, M., Hassan, M. A., et al., 2012. A network model for prediction and diagnosis of sediment dynamics at the watershed scale. Journal of Geophysical Research – Earth Surface, 117. 10.1029/2012jf002400.Google Scholar
Patton, P. C., 1988. Geomorphic response of streams to floods in the glaciated terrain of southern New England. In: Baker, V. R., Kochel, R. C., Patton, P. C. (eds.), Flood Geomorphology. Wiley, New York, pp. 261277.Google Scholar
Pavlowsky, R. T., Lecce, S. A., Owen, M. R., Martin, D. J., 2017. Legacy sediment, lead, and zinc storage in channel and floodplain deposits of the Big River, Old Lead Belt Mining District, Missouri, USA. Geomorphology, 299, 5475.Google Scholar
Payne, B. A., Lapointe, M. F., 1997. Channel morphology and lateral stability: effects on distribution of spawning and rearing habitat for Atlantic salmon in a wandering cobble-bed river. Canadian Journal of Fisheries and Aquatic Sciences, 54(11), 26272636.Google Scholar
Peakall, J., Ashworth, P., Best, J., 1996. Physical modeling in fluvial geomorphology: principles, applications and unresolved issues. In: Rhoads, B. L., Thorn, C. E. (eds.), The Scientific Nature of Geomorphology. Wiley and Sons, Chichester, UK, pp. 221253.Google Scholar
Peakall, J., Ashworth, P. J., Best, J. L., 2007. Meander-bend evolution, alluvial architecture, and the role of cohesion in sinuous river channels: a flume study. Journal of Sedimentary Research, 77(3–4), 197212.Google Scholar
Pearson, A. J., Pizzuto, J., 2015. Bed load transport over run-of-river dams, Delaware, USA. Geomorphology, 248, 382395.Google Scholar
Peckham, S. D., 1995. New results for self-similar trees with applications to river networks. Water Resources Research, 31(4), 10231029.Google Scholar
Peckham, S. D., Gupta, V. K., 1999. A reformulation of Horton’s laws for large river networks in terms of statistical self-similarity. Water Resources Research, 35(9), 27632777.Google Scholar
Pelletier, J. D., 1999. Self-organization and scaling relationships of evolving river networks. Journal of Geophysical Research – Solid Earth, 104(B4), 73597375.Google Scholar
Pelletier, J. D., 2003. Drainage basin evolution in the Rainfall Erosion Facility: dependence on initial conditions. Geomorphology, 53(1–2), 183196.Google Scholar
Pelletier, J. D., 2004. Persistent drainage migration in a numerical landscape evolution model. Geophysical Research Letters, 31(20). 10.1029/2004gl020802.Google Scholar
Pelletier, J. D., 2012a. Fluvial and slope-wash erosion of soil-mantled landscapes: detachment- or transport-limited? Earth Surface Processes and Landforms, 37(1), 3751.Google Scholar
Pelletier, J. D., 2012b. A spatially distributed model for the long-term suspended sediment discharge and delivery ratio of drainage basins. Journal of Geophysical Research-Earth Surface, 117. 10.1029/2011jf002129.Google Scholar
Pelletier, J. D., 2013. A robust, two-parameter method for the extraction of drainage networks from high-resolution digital elevation models (DEMs): evaluation using synthetic and real-world DEMs. Water Resources Research, 49(1), 7589.Google Scholar
Pelosi, A., Schumer, R., Parker, G., Ferguson, R. I., 2016. The cause of advective slowdown of tracer pebbles in rivers: implementation of Exner-Based Master Equation for coevolving streamwise and vertical dispersion. Journal of Geophysical Research-Earth Surface, 121(3), 623637. 10.1002/2015jf003497.Google Scholar
Penck, W., 1972. Morphological Analysis of Land Forms: A Contribution to Physical Geology. Translated by Hella Czech and Cumming Boswell. Hafner Pub. Co., New York.Google Scholar
Penna, N., De Marchis, M., Canelas, O. B., et al., 2018. Effect of the junction angle on turbulent flow at a hydraulic confluence. Water, 10(4). 10.3390/w10040469.Google Scholar
Penning-Rowsell, E. C., Townshend, J. R. G., 1978. Influence of scale on the factors affecting stream channel slope. Transactions of the Institute of British Geographers, 3(4), 395415.Google Scholar
Perkins, H. J., 1970. The formation of vorticity in turbulent flow. Journal of Fluid Mechanics, 44, 721740.Google Scholar
Perron, J. T., Fagherazzi, S., 2012. The legacy of initial conditions in landscape evolution. Earth Surface Processes and Landforms, 37(1), 5263.Google Scholar
Perron, J. T., Royden, L., 2013. An integral approach to bedrock river profile analysis. Earth Surface Processes and Landforms, 38(6), 570576.Google Scholar
Perron, J. T., Kirchner, J. W., Dietrich, W. E., 2008a. Spectral signatures of characteristic spatial scales and nonfractal structure in landscapes. Journal of Geophysical Research – Earth Surface, 113(F4). 10.1029/2007jf000866.Google Scholar
Perron, J. T., Dietrich, W. E., Kirchner, J. W., 2008b. Controls on the spacing of first-order valleys. Journal of Geophysical Research – Earth Surface, 113(F4). 10.1029/2007jf000977.Google Scholar
Perron, J. T., Kirchner, J. W., Dietrich, W. E., 2009. Formation of evenly spaced ridges and valleys. Nature, 460(7254), 502505.Google Scholar
Perron, J. T., Richardson, P. W., Ferrier, K. L., Lapotre, M., 2012. The root of branching river networks. Nature, 492(7427), 100103.Google Scholar
Persendt, F. C., Gomez, C., 2016. Assessment of drainage network extractions in a low-relief area of the Cuvelai Basin (Namibia) from multiple sources: LiDAR, topographic maps, and digital aerial orthophotographs. Geomorphology, 260, 3250.Google Scholar
Perucca, E., Camporeale, C., Ridolfi, L., 2005. Nonlinear analysis of the geometry of meandering rivers. Geophysical Research Letters, 32(3). 10.1029/2004gl021966.Google Scholar
Perucca, E., Camporeale, C., Ridolfi, L., 2007. Significance of the riparian vegetation dynamics on meandering river morphodynamics. Water Resources Research, 43(3). 10.1029/2006wr005234.Google Scholar
Petit, F., 1994. Dimensionless critical shear stress evaluation from flume experiments using different gravel beds. Earth Surface Processes and Landforms, 19(6), 565576.Google Scholar
Petit, F., Pauquet, A., 1997. Bankfull discharge recurrence interval in gravel-bed rivers. Earth Surface Processes and Landforms, 22(7), 685693.Google Scholar
Petit, F., Gob, F., Houbrechts, G., Assani, A. A., 2005. Critical specific stream power in gravel-bed rivers. Geomorphology, 69(1–4), 92101.Google Scholar
Petit, F., Houbrechts, G., Peeters, A., et al., 2015. Dimensionless critical shear stress in gravel-bed rivers. Geomorphology, 250, 308320.Google Scholar
Petroff, A. P., Devauchelle, O., Seybold, H., Rothman, D. H., 2013. Bifurcation dynamics of natural drainage networks. Philosophical Transactions of the Royal Society A – Mathematical Physical and Engineering Sciences, 371(2004). 10.1098/rsta.2012.0365.Google Scholar
Petts, G. E., Gurnell, A. M., 2005. Dams and geomorphology: research progress and future directions. Geomorphology, 71(1–2), 2747.Google Scholar
Petts, G., Gurnell, A., 2013. Hydrogeomorphic effects of reservoirs, dams and diversions. In: Shroder, J. C. (ed.), Treatise on Geomorphology, Vol. 13, Geomorphology of Human Disturbances, Climate Change and Natural Hazards, James, L. A., Harden, C. P., Clague, C. C. (vol. eds). Academic Press, San Diego, CA, pp. 96114.Google Scholar
Petts, G. E., Thoms, M. C., 1987. Morphology and sedimentology of a tributary confluence bar in a regulated river – North Tyne, UK. Earth Surface Processes and Landforms, 12(4), 433440.Google Scholar
Petts, J., 2007. Learning about learning: lessons from public engagement and deliberation on urban river restoration. Geographical Journal, 173, 300311.Google Scholar
Pfeiffer, A. M., Finnegan, N. J., Willenbring, J. K., 2017. Sediment supply controls equilibrium channel geometry in gravel rivers. Proceedings of the National Academy of Sciences of the United States of America, 114(13), 33463351.Google Scholar
Phillips, C. B., Jerolmack, D. J., 2014. Dynamics and mechanics of bed-load tracer particles. Earth Surface Dynamics, 2(2), 513530.Google Scholar
Phillips, C. B., Jerolmack, D. J., 2016. Self-organization of river channels as a critical filter on climate signals. Science, 352(6286), 694697.Google Scholar
Phillips, C. B., Scatena, F. N., 2013. Reduced channel morphological response to urbanization in a flood-dominated humid tropical environment. Earth Surface Processes and Landforms, 38(9), 970982.Google Scholar
Phillips, C. B., Martin, R. L., Jerolmack, D. J., 2013. Impulse framework for unsteady flows reveals superdiffusive bed load transport. Geophysical Research Letters, 40(7). 10.1002/grl.50323.Google Scholar
Phillips, J. D., 1990. The instability of hydraulic geometry. Water Resources Research, 26(4), 739744.Google Scholar
Phillips, J. D., 1991. Fluvial sediment budgets in the North Carolina Piedmont. Geomorphology, 4(3–4), 231241.Google Scholar
Phillips, J. D., 1992. The end of equilibrium? Geomorphology, 5(3–5), 195201.Google Scholar
Phillips, J. D., 1993. Interpreting the fractal dimension of river networks. In: Lam, N. S., De Cola, L. (eds.), Fractals in Geography. Prentice Hall, Englewood Cliffs, NJ, pp. 142157.Google Scholar
Phillips, J. D., 1996. Deterministic complexity, explanation, and predictability in geomorphic systems. In: Rhoads, B. L., Thorn, C. E. (eds.), The Scientific Nature of Geomorphology. Wiley, Chichester, UK, pp. 315335.Google Scholar
Phillips, J. D., 2002. Erosion, isostatic response, and the missing peneplains. Geomorphology, 45(3–4), 225241.Google Scholar
Phillips, J. D., 2003. Alluvial storage and the long-term stability of sediment yields. Basin Research, 15(2), 153163.Google Scholar
Phillips, J. D., 2006a. Evolutionary geomorphology: thresholds and nonlinearity in landform response to environmental change. Hydrology and Earth System Sciences, 10(5), 731742.Google Scholar
Phillips, J. D., 2006b. Deterministic chaos and historical geomorphology: a review and look forward. Geomorphology, 76(1–2), 109121.Google Scholar
Phillips, J. D., 2010. The job of the river. Earth Surface Processes and Landforms, 35(3), 305313.Google Scholar
Phillips, J. D., 2012. Log-jams and avulsions in the San Antonio River Delta, Texas. Earth Surface Processes and Landforms, 37(9), 936950.Google Scholar
Phillips, J. D., 2014. Anastomosing channels in the lower Neches River valley, Texas. Earth Surface Processes and Landforms, 39(14), 18881899.Google Scholar
Phillips, J. D., Slattery, M. C., 2006. Sediment storage, sea level, and sediment delivery to the ocean by coastal plain rivers. Progress in Physical Geography, 30(4), 513530.Google Scholar
Phillips, J. D., Van Dyke, C., 2016. Principles of geomorphic disturbance and recovery in response to storms. Earth Surface Processes and Landforms, 41(7), 971979.Google Scholar
Phillips, J. D., Slattery, M. C., Musselman, Z. A., 2005. Channel adjustments of the lower Trinity River, Texas, downstream of Livingston Dam. Earth Surface Processes and Landforms, 30(11), 14191439.Google Scholar
Pickup, G., Rieger, W. A., 1979. Conceptual model of the relationship between channel characteristics and discharge. Earth Surface Processes and Landforms, 4(1), 3742.Google Scholar
Pickup, G., Warner, R. F., 1976. Effects of hydrological regime on magnitude and frequency of dominant discharge. Journal of Hydrology, 29(1–2), 5175.Google Scholar
Piegay, H., Vaudor, L., 2016. Statistics and fluvial geomorphology. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology, 2nd Edition. Wiley, Chichester, UK, pp. 476506.Google Scholar
Piegay, H., Darby, S. E., Mosselman, E., Surian, N., 2005. A review of techniques available for delimiting the erodible river corridor: a sustainable approach to managing bank erosion. River Research and Applications, 21(7), 773789.Google Scholar
Piegay, H., Grant, G., Nakamura, F., Trustrum, N., 2006. Braided river management: from assessment of river behavior to improved sustainable management. In: Sambrook Smith, G. H., Best, J. L., Bristow, C. S., Petts, G. E. (eds.), Braided Rivers: Process, Deposits, Ecology and Management, International Association of Sedimentologists Special Publication 36. Blackwell, Malden, MA, pp. 257274.Google Scholar
Piegay, H., Hupp, C. R., Citterio, A., et al., 2008. Spatial and temporal variability in sedimentation rates associated with cutoff channel infill deposits: Ain River, France. Water Resources Research, 44(5). 10.1029/2006wr005260.Google Scholar
Pierce, A. R., King, S. L., 2008. Spatial dynamics of overbank sedimentation in floodplain systems. Geomorphology, 100(3–4), 256268.Google Scholar
Pieri, D. C., 1984. Junction angles in drainage networks. Journal of Geophysical Research, 89(NB8), 68786884.Google Scholar
Pierson, T. C., 2005. Hyperconcentrated flow – transitional process between water flow and debris flow. In: Jakob, M., Hungr, O. (eds.), Debris-Flow Hazards and Related Phenomena. Springer, Berlin, pp. 159202.Google Scholar
Pietsch, T. J., Nanson, G. C., 2011. Bankfull hydraulic geometry; the role of in-channel vegetation and downstream declining discharges in the anabranching and distributary channels of the Gwydir distributive fluvial system, southeastern Australia. Geomorphology, 129(1–2), 152165.Google Scholar
Pitlick, J., 1992. Flow resistance under conditions of intense gravel transport. Water Resources Research, 28(3), 891903.Google Scholar
Pittaluga, M. B., Nobile, G., Seminara, G., 2009. A nonlinear model for river meandering. Water Resources Research, 45. 10.1029/2008wr007298.Google Scholar
Pizzuto, J. E., 1984a. An evaluation of methods for calculating the concentration of suspended bed material in rivers. Water Resources Research, 20(10), 13811389.Google Scholar
Pizzuto, J. E., 1984b. Bank erodibility of shallow sandbed streams. Earth Surface Processes and Landforms, 9(2), 113124.Google Scholar
Pizzuto, J. E., 1987. Sediment diffusion during overbank flows. Sedimentology, 34(2), 301317.Google Scholar
Pizzuto, J. E., 1995. Downstream fining in a network of gravel-bedded rivers. Water Resources Research, 31(3), 753759.Google Scholar
Pizzuto, J. E., 2002. Effects of dam removal on river form and process. Bioscience, 52(8), 683691.Google Scholar
Pizzuto, J. E., 2016. Modelling fluvial morphodynamics. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology, 2nd Edition. Wiley, Chichester, UK, pp. 442455.Google Scholar
Pizzuto, J. E., Meckelnburg, T. S., 1989. Evaluation of a linear bank erosion equation. Water Resources Research, 25(5), 10051013.Google Scholar
Pizzuto, J. E., Hession, W. C., McBride, M., 2000. Comparing gravel-bed rivers in paired urban and rural catchments of southeastern Pennsylvania. Geology, 28(1), 7982.Google Scholar
Placzkowska, E., Gornik, M., Mocior, E., et al., 2015. Spatial distribution of channel heads in the Polish Flysch Carpathians. Catena, 127, 240249.Google Scholar
Playfair, J., 1802. Illustrations of the Huttonian Theory of the Earth. Edinburgh; London: Printed for Cadell and Davies; William Creech. Nineteenth Century Collections Online, http://tinyurl.gale.com/tinyurl/CPVmk4 (accessed December 1, 2019).Google Scholar
Plumb, B. D., Annable, W. K., Thompson, P. J., Hassan, M. A., 2017. The impact of urbanization on temporal changes in sediment transport in a gravel bed channel in southern Ontario, Canada. Water Resources Research, 53(10), 84438458. 10.1002/2016WR020288.Google Scholar
Poesen, J., 2018. Soil erosion in the Anthropocene: research needs. Earth Surface Processes and Landforms, 43(1), 6484.Google Scholar
Poff, N. L., Olden, J. D., Merritt, D. M., Pepin, D. M., 2007. Homogenization of regional river dynamics by dams and global biodiversity implications. Proceedings of the National Academy of Sciences of the United States of America, 104(14), 57325737.Google Scholar
Pollen, N., 2007. Temporal and spatial variability in root reinforcement of streambanks: accounting for soil shear strength and moisture. Catena, 69(3), 197205.Google Scholar
Pollen, N., Simon, A., 2005. Estimating the mechanical effects of riparian vegetation on stream bank stability using a fiber bundle model. Water Resources Research, 41(7). 10.1029/2004wr003801.Google Scholar
Pollock, M. M., Beechie, T. J., Wheaton, J. M., et al., 2014. Using beaver dams to restore incised stream ecosystems. Bioscience, 64(4), 279290.Google Scholar
Polvi, L. E., Wohl, E., Merritt, D. M., 2014. Modeling the functional influence of vegetation type on streambank cohesion. Earth Surface Processes and Landforms, 39, 12451258.Google Scholar
Pornprommin, A., Izumi, N., 2010. Inception of stream incision by seepage erosion. Journal of Geophysical Research – Earth Surface, 115. 10.1029/2009jf001369.Google Scholar
Pornprommin, A., Takei, Y., Wubneh, A. M., Izumi, N., 2010. Channel inception in cohesionless sediment by seepage erosion. Journal of Hydro-Environment Research, 3(4), 232238.Google Scholar
Portenga, E. W, Bierman, P.A., 2011. Understanding Earth’s eroding surface with 10Be. GSA Today, 21, 410.Google Scholar
Portenga, E. W., Westaway, K. E., Bishop, P., 2016. Timing of post-European settlement alluvium deposition in SE Australia: a legacy of European land-use in the Goulburn Plains. Holocene, 26(9), 14721485.Google Scholar
Potter, K. W., 1991. Hydrological impacts of changing land management practices in a moderate-sized agricultural catchment. Water Resources Research, 27(5), 845855.Google Scholar
Powell, D. M., 1998. Patterns and processes of sediment sorting in gravel-bed rivers. Progress in Physical Geography, 22(1), 132.Google Scholar
Powell, D. M., Ashworth, P. J., 1995. Spatial pattern of flow competence and bed load transport in a divided gravel bed river. Water Resources Research, 31(3), 741752.Google Scholar
Powell, D. M., Ockelford, A., Rice, S. P., et al., 2016. Structural properties of mobile armors formed at different flow strengths in gravel-bed rivers. Journal of Geophysical Research – Earth Surface, 121(8), 14941515. 10.1002/2015jf003794.Google Scholar
Powell, G. E., Ward, A. D., Mecklenburg, D. E., Jayakaran, A. D., 2007. Two-stage channel systems: Part 1, a practical approach for sizing agricultural ditches. Journal of Soil and Water Conservation, 62(4), 277286.Google Scholar
Powell, J. W., 1875. Exploration of the Colorado River of the West and its tributaries. Smithsonian Institution, U.S. Government Printing Office, Washington, DC.Google Scholar
Prancevic, J. P., Lamb, M. P., 2015a. Unraveling bed slope from relative roughness in initial sediment motion. Journal of Geophysical Research – Earth Surface, 120(3), 474489. 10.1002/2014jf003323.Google Scholar
Prancevic, J. P., Lamb, M. P., 2015b. Particle friction angles in steep mountain channels. Journal of Geophysical Research – Earth Surface, 120(2), 242259. 10.1002/2014jf003286.Google Scholar
Praskievicz, S., 2015. A coupled hierarchical modeling approach to simulating the geomorphic response of river systems to anthropogenic climate change. Earth Surface Processes and Landforms, 40(12), 16161630.Google Scholar
Prestegaard, K. L., 1983. Variables influencing water-surface slopes in gravel-bed streams at bankfull stage. Geological Society of America Bulletin, 94(5), 673678.Google Scholar
Pringle, C. M., Naiman, R. J., Bretschko, G., et al., 1988. Patch dynamics in lotic systems – the stream as a mosaic. Journal of the North American Benthological Society, 7(4), 503524.Google Scholar
Pritchard, D., Roberts, G. G., White, N. J., Richardson, C. N., 2009. Uplift histories from river profiles. Geophysical Research Letters, 36. 10.1029/2009gl040928.Google Scholar
Pyne, S. J., 1980. Grove Karl Gilbert: A Great Engine of Research. University of Texas Press, Austin, TX.Google Scholar
Pyrce, R. S., Ashmore, P. E., 2003a. The relation between particle path length distributions and channel morphology in gravel-bed streams: a synthesis. Geomorphology, 56(1–2), 167187.Google Scholar
Pyrce, R. S., Ashmore, P. E., 2003b. Particle path length distributions in meandering gravel-bed streams: results from physical models. Earth Surface Processes and Landforms, 28(9), 951966.Google Scholar
Pyrce, R. S., Ashmore, P. E., 2005. Bedload path length and point bar development in gravel-bed river models. Sedimentology, 52(4), 839857.Google Scholar
Qian, H., Cao, Z., Liu, H., Pender, G., 2017. Numerical modelling of alternate bar formation, development and sediment sorting in straight channels. Earth Surface Processes and Landforms, 42(4), 555574.Google Scholar
Qin, J., Zhong, D. Y., Wang, G. Q., Ng, S. L., 2012. On characterization of the imbrication of armored gravel surfaces. Geomorphology, 159, 116124.Google Scholar
Qing-Yuan, Y., Xian-Ye, W., Wei-Zhen, L., Xie-Kang, W., 2009. Experimental study on characteristics of separation zone in confluence zones in rivers. Journal of Hydrologic Engineering, 14(2), 166171.Google Scholar
Quraishy, M. S., 1944. The origin of curves in rivers. Current Science, 13, 3639.Google Scholar
Radoane, M., Radoane, N., Dumitriu, D., Miclaus, C., 2008. Downstream variation in bed sediment size along the East Carpathian rivers: evidence of the role of sediment sources. Earth Surface Processes and Landforms, 33(5), 674694.Google Scholar
Ramamurthy, A. S., Carballada, L. B., Tran, D. M., 1988. Combining open channel flow at right angled junctions. Journal of Hydraulic Engineering, 114, 14491460.Google Scholar
Ramon, C. L., Hoyer, A. B., Armengol, J., Dolz, J., Rueda, F. J., 2013. Mixing and circulation at the confluence of two rivers entering a meandering reservoir. Water Resources Research, 49(3), 14291445.Google Scholar
Ramon, C. L., Armengol, J., Dolz, J., Prats, J., Rueda, F. J., 2014. Mixing dynamics at the confluence of two large rivers undergoing weak density variations. Journal of Geophysical Research – Oceans, 119(4), 23862402.Google Scholar
Ramon, C. L., Prats, J., Rueda, F. J., 2016. The influence of flow inertia, buoyancy, wind, and flow unsteadiness on mixing at the asymmetrical confluence of two large rivers. Journal of Hydrology, 539, 1126.Google Scholar
Rantz, S. E., and others, 1982. Measurement and computation of streamflow: Vol. 2. Computation of discharge, U.S. Geological Survey Water Supply Paper 2175. U.S. Government Printing Office, Washington, DC.Google Scholar
Rathbun, R. E., Rostad, C. E., 2004. Lateral mixing in the Mississippi River below the confluence with the Ohio River. Water Resources Research, 40(5). 10.1029/2003wr002381.Google Scholar
Rathburn, S. L., Rubin, Z. K., Wohl, E. E., 2013. Evaluating channel response to an extreme sedimentation event in the context of historical range of variability: Upper Colorado River, USA. Earth Surface Processes and Landforms, 38(4), 391406.Google Scholar
Recking, A., 2009. Theoretical development on the effects of changing flow hydraulics on incipient bed load motion. Water Resources Research, 45. 10.1029/2008wr006826.Google Scholar
Recking, A., 2016. A generalized threshold model for computing bed load grain size distribution. Water Resources Research, 52(12), 92749289.Google Scholar
Recking, A., Liebault, F., Peteuil, C., Jolimet, T., 2012. Testing bedload transport equations with consideration of time scales. Earth Surface Processes and Landforms, 37(7), 774789.Google Scholar
Refice, A., Giachetta, E., Capolongo, D., 2012. SIGNUM: a Matlab, TIN-based landscape evolution model. Computers & Geosciences, 45, 293303.Google Scholar
Reid, D. A., Hassan, M. A., Floyd, W., 2016. Reach-scale contributions of road-surface sediment to the Honna River, Haida Gwaii, BC. Hydrological Processes, 30(19), 34503465.Google Scholar
Reid, D. E., Hickin, E. J., Babakaiff, S. C., 2010. Low-flow hydraulic geometry of small, steep mountain streams in southwest British Columbia. Geomorphology, 122(1–2), 3955.Google Scholar
Reid, I., Laronne, J. B., 1995. Bedload sediment transport in an ephemeral stream and a comparison with seasonal and perennial counterparts. Water Resources Research, 31(3), 773781.Google Scholar
Reid, L. M., Dunne, T., 1984. Sediment production from forest road surfaces. Water Resources Research, 20(11), 17531761.Google Scholar
Reid, L. M., Dunne, T., 2016. Sediment budgets as an organizing framework in fluvial geomorphology. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology. Wiley and Sons, Chichester, UK, pp. 357380.Google Scholar
Reinfelds, I., Nanson, G., 1993. Formation of braided river floodplains, Waimakiriri River, New Zealand. Sedimentology, 40(6), 11131127.Google Scholar
Renard, K. G., Foster, G. R., Weesies, G. A., Porter, J. P., 1991. RUSLE – revised universal soil loss equation. Journal of Soil and Water Conservation, 46(1), 3033.Google Scholar
Renschler, C. S., 2003. Designing geo-spatial interfaces to scale process models: the GeoWEPP approach. Hydrological Processes, 17(5), 10051017.Google Scholar
Renwick, W. H., 1992. Equilibrium, disequilibrium, and nonequilibrium landforms in the landscape. Geomorphology, 5(3–5), 265276.Google Scholar
Renwick, W. H., Andereck, Z. D., 2006. Reservoir sedimentation trends in Ohio, USA: sediment delivery and response to land-use change. In: Rowan, J. S., Duck, R. W., Werritty, A. (eds.), Sediment Dynamics and the Hydromorphology of Fluvial Systems. IAHS Publication No. 306, IAHS Press, Wallingford, UK, pp. 341347.Google Scholar
Rhoads, B. L., 1987a. Stream power terminology. Professional Geographer, 39(2), 189195.Google Scholar
Rhoads, B. L., 1987b. Changes in stream channel characteristics at tributary junctions. Physical Geography, 8, 346361.Google Scholar
Rhoads, B. L., 1988. Mutual adjustments between process and form in a desert mountain fluvial system. Annals of the Association of American Geographers, 78(2), 271287.Google Scholar
Rhoads, B. L., 1989. Longitudinal variations in the size and sorting of bed material along six arid-region mountain streams. In: Yair, A., Berkowicz, S. M. (eds.), Arid and Semiarid Environments – Geomorphological and Pedological Aspects, Catena Supplement 14. Catena Verlag, Cremlingen-Destadt, W. Germany, pp. 87105.Google Scholar
Rhoads, B. L., 1990a. Hydrologic characteristics of a small desert mountain stream: implications for short-term magnitude and frequency of bedload transport. Journal of Arid Environments, 18, 151163.Google Scholar
Rhoads, B. L., 1990b. The impact of stream channelization on the geomorphic stability of an arid-region river. National Geographic Research, 6, 157177.Google Scholar
Rhoads, B. L., 1991a. A continuously varying parameter model of downstream hydraulic geometry. Water Resources Research, 27(8), 18651872.Google Scholar
Rhoads, B. L., 1991b. Multicollinearity and parameter estimation in simultaneous-equation models of fluvial systems. Geographical Analysis, 23(4), 346361.Google Scholar
Rhoads, B. L., 1991c. Impact of agricultural development on regional drainage in the lower Santa Cruz Valley, Arizona, USA. Environmental Geology and Water Sciences, 18(2), 119135.Google Scholar
Rhoads, B. L., 1992. Statistical models of fluvial systems. Geomorphology, 5(3–5), 433455.Google Scholar
Rhoads, B. L., 1995. Stream power: a unifying theme for urban fluvial geomorphology. In: Herricks, E. E. (ed.), Stormwater Runoff and Receiving Systems: Impact, Monitoring, and Assessment. Lewis Publishers, Boca Raton, FL, pp. 6575.Google Scholar
Rhoads, B. L., 1996. Mean structure of transport-effective flows at an asymmetrical confluence when the main stream is dominant. In: Ashworth, P. J., Bennett, S. J., Best, J. L., McLelland, S. J. (eds.), Coherent Flow Structures in Open Channels. John Wiley & Sons Ltd, Chichester, UK, pp. 491517.Google Scholar
Rhoads, B. L., 1999. Beyond pragmatism: the value of philosophical discourse for physical geography. Annals of the Association of American Geographers, 89(4), 760771.Google Scholar
Rhoads, B. L., 2006a. The dynamic basis of geomorphology reenvisioned. Annals of the Association of American Geographers, 96(1), 1430.Google Scholar
Rhoads, B. L., 2006b. Scaling of confluence dynamics in river systems: some general considerations. In: Parker, G., Garcia, M. H. (eds.), River, Coastal and Estuarine Morphodynamics, RCEM 2005. Taylor and Francis, London, pp. 379387.Google Scholar
Rhoads, B. L., 2013. Process in geomorphology. In: Shroder, J. W. (ed.), Treatise on Geomorphology, Vol. 1, The Foundations of Geomorphology, Orme, A. R. and Sack, D. (vol. eds.). Elsevier, New York, pp. 190204.Google Scholar
Rhoads, B. L., Cahill, R. A., 1999. Geomorphological assessment of sediment contamination in an urban stream system. Applied Geochemistry, 14(4), 459483.Google Scholar
Rhoads, B. L., Herricks, E. E., 1996. Naturalization of headwater streams in Illinois: challenges and possibilities. In: Shields, F. D., Jr., Brookes, A. (eds.), River Channel Restoration. Wiley, Chichester, UK, pp. 331367.Google Scholar
Rhoads, B. L., Johnson, K. K., 2018. Three-dimensional flow structure, morphodynamics, suspended sediment, and thermal mixing at an asymmetrical river confluence of a straight tributary and curving main channel. Geomorphology, 323, 5169. 10.1016/j.geomorph.2018.09.009.Google Scholar
Rhoads, B. L., Kenworthy, S. T., 1995. Flow structure at an asymmetrical stream confluence. Geomorphology, 11, 273293.Google Scholar
Rhoads, B. L., Kenworthy, S. T., 1998. Time-averaged flow structure in the central region of a stream confluence. Earth Surface Processes and Landforms, 23, 171191.Google Scholar
Rhoads, B. L., Kenworthy, S. T., 1999. On secondary circulation, helical motion and Rozovskii-based analysis of time-averaged two-dimensional velocity fields at confluences. Earth Surface Processes and Landforms, 24(4), 369375.Google Scholar
Rhoads, B. L., Massey, K. D., 2012. Flow structure and channel change in a sinuous grass-lined stream within an agricultural drainage ditch: implications for ditch stability and aquatic habitat. River Research and Applications, 28(1), 3952.Google Scholar
Rhoads, B. L., Miller, M. V., 1991. Impact of flow variability on the morphology of a low-energy meandering river. Earth Surface Processes and Landforms, 16(4), 357367.Google Scholar
Rhoads, B. L., Sukhodolov, A. N., 2001. Field investigation of three-dimensional flow structure at stream confluences: 1. thermal mixing and time-averaged velocities. Water Resources Research, 37(9), 23932410.Google Scholar
Rhoads, B. L., Sukhodolov, A. N., 2004. Spatial and temporal structure of shear layer turbulence at a stream confluence. Water Resources Research, 40(6). 10.1029/2003WR002811.Google Scholar
Rhoads, B. L., Sukhodolov, A. N., 2008. Lateral momentum flux and the spatial evolution of flow within a confluence mixing interface. Water Resources Research, 44(8). 10.1029/2007wr006634.Google Scholar
Rhoads, B. L., Thorn, C. E., 1993. Geomorphology as science – the role of theory. Geomorphology, 6(4), 287307.Google Scholar
Rhoads, B. L., Thorn, C. E., 1996. Toward a philosophy of geomorphology. In: Rhoads, B. L., Thorn, C. E. (eds.), The Scientific Nature of Geomorphology. Wiley, Chichester, UK, pp. 115143.Google Scholar
Rhoads, B. L., Thorn, C. E., 2011. The role and character of theory in geomorphology. In: Gregory, K. J., Goudie, A. S. (eds.), The Sage Handbook of Geomorphology. Sage, Los Angeles, CA, pp. 5977.Google Scholar
Rhoads, B. L., Welford, M. R., 1991. Initiation of river meandering. Progress in Physical Geography, 15(2), 127156.Google Scholar
Rhoads, B. L., Wilson, D., Urban, M., Herricks, E. E., 1999. Interaction between scientists and nonscientists in community-based watershed management: emergence of the concept of stream naturalization. Environmental Management, 24(3), 297308.Google Scholar
Rhoads, B. L., Schwartz, J. S., Porter, S., 2003. Stream geomorphology, bank vegetation, and three-dimensional habitat hydraulics for fish in midwestern agricultural streams. Water Resources Research, 39(8). 10.1029/2003WR002294Google Scholar
Rhoads, B .L., Garcia, M. H., Rodriguez, J., et al., 2008. Methods for evaluating the geomorphological performance of naturalized rivers: examples from the Chicago Metropolitan Area. In: Darby, S. E., Sear, D. A. (eds.), River Restoration: Managing for Uncertainty in Restoring Physical Habitat. Wiley, Chichester, UK, pp. 209228.Google Scholar
Rhoads, B. L., Riley, J. D., Mayer, D. R., 2009. Response of bed morphology and bed material texture to hydrological conditions at an asymmetrical stream confluence. Geomorphology, 109(3–4), 161173.Google Scholar
Rhoads, B. L., Engel, F. L., Abad, J. D., 2011. Pool-riffle design based on geomorphic principles for naturalizing straight channels. In: Simon, A., Bennett, S. J., Castro, J. M. (eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools. American Geophysical Union, Washington, DC, pp. 367384.Google Scholar
Rhoads, B. L., Lewis, Q. W., Andresen, W., 2016. Historical changes in channel network extent and channel planform in an intensively managed landscape: natural versus human-induced effects. Geomorphology, 252, 1731. 10.1016/j.geomorph.2015.04.021.Google Scholar
Rhodes, D. D., 1977. The b-f-m diagram – graphical representation and interpretation of at-a-station hydraulic geometry. American Journal of Science, 277(1), 7396.Google Scholar
Rhodes, D. D., 1978. World-wide variations in hydraulic geometry exponents of stream channels – analysis and some observations – comments. Journal of Hydrology, 39(1–2), 193197.Google Scholar
Rhodes, D. D., 1987. The b-f-m diagram for downstream hydraulic geometry. Geografiska Annaler Series A Physical Geography, 69(1), 147161.Google Scholar
Rice, S., 1998. Which tributaries disrupt downstream fining along gravel-bed rivers? Geomorphology, 22(1), 3956.Google Scholar
Rice, S., 1999. The nature and controls on downstream fining within sedimentary links. Journal of Sedimentary Research, 69(1), 3239.Google Scholar
Rice, S. P., 2017. Tributary connectivity, confluence aggradation and network biodiversity. Geomorphology, 277, 616.Google Scholar
Rice, S., Church, M., 1998. Grain size along two gravel-bed rivers: statistical variation, spatial pattern and sedimentary links. Earth Surface Processes and Landforms, 23(4), 345363.Google Scholar
Rice, S. P., Church, M., 2001. Longitudinal profiles in simple alluvial systems. Water Resources Research, 37(2), 417426.Google Scholar
Rice, S. P., Church, M., 2010. Grain-size sorting within river bars in relation to downstream fining along a wandering channel. Sedimentology, 57(1), 232251.Google Scholar
Rice, S. P., Kiffney, P., Greene, C., Pess, G. R., 2008. The ecological importance of tributaries and confluences. In: Rice, S. P., Roy, A. G., Rhoads, B. L. (eds.), River Confluences, Tributaries and the Fluvial Network. Wiley, Chichester, UK, pp. 209242.Google Scholar
Rice, S.P., Roy, A.G., Rhoads, B.L. (eds.), 2008. River Confluences, Tributaries, and the Fluvial Network. Wiley, Chichester, UK.Google Scholar
Rice, S. P., Church, M., Wooldridge, C. L., Hickin, E. J., 2009. Morphology and evolution of bars in a wandering gravel-bed river; lower Fraser river, British Columbia, Canada. Sedimentology, 56(3), 709736.Google Scholar
Richards, D., 2018. Three-dimensional flow, morphologic change, and sediment deposition and distribution of actively evolving neck cutoffs located on the White River, Arkansas. PhD Dissertation, Louisiana State University.Google Scholar
Richards, K. S., 1973. Hydraulic geometry and channel roughness – a nonlinear system. American Journal of Science, 273(10), 877896.Google Scholar
Richards, K. S., 1976a. Morphology of pool-riffle sequences. Earth Surface Processes and Landforms, 1(1), 7188.Google Scholar
Richards, K. S., 1976b. Channel width and the riffle pool sequence. Geological Society of America Bulletin, 87(6), 883890.Google Scholar
Richards, K. S., 1978. Channel geometry in the riffle-pool sequence. Geografiska Annaler Series A Physical Geography, 60(1–2), 2327.Google Scholar
Richards, K. S., 1980. A note on changes in channel geometry at tributary junctions. Water Resources Research, 16(1), 241244.Google Scholar
Richards, K. S., 1993. Sediment delivery and the drainage network. In: Beven, K., Kirkby, M. J. (eds.), Channel Network Hydrology. Wiley and Sons, Chichester, UK, pp. 221254.Google Scholar
Richards, K. S., 1996. Samples and cases: generalisation and explanation in geomorphology. In: Rhoads, B. L., Thorn, C. E. (eds.), The Scientific Nature of Geomorphology. Wiley and Sons, Chichester, UK, pp. 171190.Google Scholar
Richards, K. S., Clifford, N., 1991. Fluvial geomorphology – structured beds in gravelly rivers. Progress in Physical Geography, 15(4), 407422.Google Scholar
Richardson, W. R., Thorne, C. R., 1998. Secondary currents around braid bar in Brahmaputra River, Bangladesh. Journal of Hydraulic Engineering, 124(3), 325328.Google Scholar
Richardson, W. R., Thorne, C. R., 2001. Multiple thread flow and channel bifurcation in a braided river: Brahmaputra-Jamuna River, Bangladesh. Geomorphology, 38(3–4), 185196.Google Scholar
Rickenmann, D., Recking, A., 2011. Evaluation of flow resistance in gravel-bed rivers through a large field data set. Water Resources Research, 47. 10.1029/2010wr009793.Google Scholar
Ridenour, G. S., Giardino, J. R., 1991. The statistical study of hydraulic geometry – a new direction for compositional data analysis. Mathematical Geology, 23(3), 349366.Google Scholar
Ridenour, G. S., Giardino, J. R., 1995. Logratio linear modeling of hydraulic geometry using indexes of flow resistance as covariates. Geomorphology, 14(1), 6572.Google Scholar
Rigby, J. R., Wren, D. G., Kuhnle, R. A., 2016. Passive acoustic monitoring of bed load for fluvial applications. Journal of Hydraulic Engineering, 142(9). 10.1061/(asce)hy.1943–7900.0001122.Google Scholar
Righini, M., Surian, N., Wohl, E., et al., 2017. Geomorphic response to an extreme flood in two Mediterranean rivers (northeastern Sardinia, Italy): analysis of controlling factors. Geomorphology, 290, 184199.Google Scholar
Rigon, R., Rodriguez-Iturbe, I., Maritan, A., et al., 1996. On Hack’s law. Water Resources Research, 32(11), 33673374.Google Scholar
Riley, J. D., Rhoads, B. L., 2012. Flow structure and channel morphology at a natural confluent meander bend. Geomorphology, 163, 8498.Google Scholar
Riley, J. D., Rhoads, B. L., Parsons, D. R., Johnson, K. K., 2015. Influence of junction angle on three-dimensional flow structure and bed morphology at confluent meander bends during different hydrological conditions. Earth Surface Processes and Landforms, 40(2), 252271.Google Scholar
Riley, S. J., 1972. A comparison of morphometric measures of bankfull. Journal of Hydrology, 17, 2331.Google Scholar
Rinaldi, M., Wyzga, B., Surian, N., 2005. Sediment mining in alluvial channels: physical effects and management perspectives. River Research and Applications, 21(7), 805828.Google Scholar
Rinaldi, M., Mengoni, B., Luppi, L., Darby, S. E., Mosselman, E., 2008. Numerical simulation of hydrodynamics and bank erosion in a river bend. Water Resources Research, 44(9). 10.1029/2008wr007008.Google Scholar
Rinaldi, M., Surian, N., Comiti, F., Bussettini, M., 2013. A method for the assessment and analysis of the hydromorphological condition of Italian streams: the Morphological Quality Index (MQI). Geomorphology, 180, 96108. 10.1016/j.geomorph.2012.09.009.Google Scholar
Rinaldo, A., Dietrich, W. E., Rigon, R., Vogel, G. K., Rodriguez-Iturbe, I., 1995. Geomorphological signatures of varying climate. Nature, 374(6523), 632635.Google Scholar
Rinaldo, A., Rigon, R., Banavar, J. R., Maritan, A., Rodriguez-Iturbe, I., 2014. Evolution and selection of river networks: statics, dynamics, and complexity. Proceedings of the National Academy of Sciences of the United States of America, 111(7), 24172424.Google Scholar
Riquier, J., Piegay, H., Lamouroux, N., Vaudor, L., 2017. Are restored side channels sustainable aquatic habitat features? Predicting the potential persistence of side channels as aquatic habitats based on their fine sedimentation dynamics. Geomorphology, 295, 507528.Google Scholar
Rittenour, T. M., Blum, M. D., Goble, R. J., 2007. Fluvial evolution of the lower Mississippi River valley during the last 100 k.y. glacial cycle: response to glaciation and sea-level change. Geological Society of America Bulletin, 119 (5–6),586608.Google Scholar
Ritter, D. F., 1975. Stratigraphic implications of coarse-grained gravel deposited as overbank sediment, southern Illinois. Journal of Geology, 83(5), 645650.Google Scholar
Ritter, D. F., 1986. Process Geomorphology, 2nd Edition. W.C. Brown, Dubuque, IA.Google Scholar
Riviere, N., Wei, C., Kouyi, G. L., Momplot, A., Mignot, E., 2015. Mixing downstream of a 90° open channel junction. Proceedings of the 36th IAHR World Congress: Deltas of the Future and What Happens Upstream. International Association of Hydro-Environment Engineering and Research.Google Scholar
Robert, A., 2003. River Processes: An Introduction to Fluvial Dynamics. Arnold, London.Google Scholar
Robert, A., Roy, A. G., 1990. On the fractal interpretation of the mainstream length-drainage area relationship. Water Resources Research, 26(5), 839842.Google Scholar
Robertson, J. M., Augspurger, C. K., 1999. Geomorphic processes and spatial patterns of primary forest succession on the Bogue Chitto River, USA. Journal of Ecology, 87(6), 10521063.Google Scholar
Robertson-Rintoul, M. S. E., Richards, K. S., 1993. Braided-channel pattern and palaeohydrology using an index of total sinuosity. In: Best, J. L., Bristow, C. S. (eds.), Braided Rivers. Geological Society of London Special Publication No. 75, Geological Society, London, pp. 113118.Google Scholar
Rodrigues, S., Mosselman, E., Claude, N., Wintenberger, C. L., Juge, P., 2015. Alternate bars in a sandy gravel bed river: generation, migration and interactions with superimposed dunes. Earth Surface Processes and Landforms, 40(5), 610628.Google Scholar
Rodriguez, J. F., Garcia, M. H., 2008. Laboratory measurements of 3-D flow patterns and turbulence in straight open channel with rough bed. Journal of Hydraulic Research, 46(4), 454465.Google Scholar
Rodriguez-Iturbe, I., 1993. The geomorphological unit hydrograph. In: Beven, K., Kirkby, M. J. (eds.), Channel Network Hydrology. Wiley, Chichester, UK, pp. 4368.Google Scholar
Rodriguez-Iturbe, I., Rinaldo, A., 1997. Fractal River Basins. Cambridge University Press, Cambridge, UK.Google Scholar
Rodriguez-Iturbe, I., Valdes, J. B., 1979. Geomorphologic structure of hydrologic response. Water Resources Research, 15(6), 14091420.Google Scholar
Roehl, J. E., 1962. Sediment source areas, delivery ratios and influencing morphological factors. IAHS Publication No. 59, IAHS Press, Wallingford, UK, pp. 202213.Google Scholar
Rogers, M. M., Moser, R. D., 1992. The 3-dimensional evolution of a plane mixing layer – the Kelvin-Helmholtz rollup. Journal of Fluid Mechanics, 243, 183226.Google Scholar
Roll-Hansen, N., 2017. A historical perspective on the distinction between basic and applied science. Journal for General Philosophy of Science, 48(4), 535551.Google Scholar
Roni, P., Beechie, T. (eds.), 2013. Stream and Watershed Restoration: A Guide to Restoring Riverine Processes and Habitats. Wiley-Blackwell, Chichester, UK.Google Scholar
Roni, P., Pess, G. R., Hanson, K., Pearsons, M., 2013a. Selecting appropriate stream and watershed restoration techniques. In: Roni, P., Beechie, T. (eds.), Stream and Watershed Restoration: A Guide to Restoring Riverine Processes and Habitats. Wiley-Blackwell, Chichester, UK, pp. 144188.Google Scholar
Roni, P., Liermann, M., Muhar, S., Schmutz, S., 2013b. Monitoring and evaluation of restoration actions. In: Roni, P., Beechie, T. (eds.), Stream and Watershed Restoration: A Guide to Restoring Riverine Processes and Habitats. Wiley-Blackwell, Chichester, UK, pp. 254279.Google Scholar
Roodsari, B. K., Chandler, D. G., 2017. Distribution of surface imperviousness in small urban catchments predicts runoff peak flows and stream flashiness. Hydrological Processes, 31(17), 29903002.Google Scholar
Roper, B. B., Buffington, J. M., Archer, E., Moyer, C., Ward, M., 2008. The role of observer variation in determining Rosgen stream types in northeastern Oregon mountain streams. Journal of the American Water Resources Association, 44(2), 417427.Google Scholar
Rosburg, T. T., Nelson, P. A., Bledsoe, B. P., 2017. Effects of urbanization on flow duration and stream flashiness: a case study of Puget Sound streams, western Washington, USA. Journal of the American Water Resources Association, 53(2), 493507.Google Scholar
Rosgen, D. L., 1994. A classification of natural rivers. Catena, 22(3), 169199.Google Scholar
Rosgen, D. L., 1996. Applied River Morphology. Wildland Hydrology, Pagosa Springs, CO.Google Scholar
Rosgen, D. L., 2007. Rosgen geomorphic channel design (Chapter 11). Natural Resources Conservation Service National Engineering Handbook Part 654 Stream Restoration Design, U.S. Department of Agriculture, Washington, DC.Google Scholar
Rosgen, D. L., 2009. Watershed Assessment of River Stability and Sediment Supply (WARSSS), 2nd Edition. Wildland Hydrology, Pagosa Springs, CO.Google Scholar
Rosgen, D. L., 2011. Natural channel design: fundamental concepts, assumptions, and methods. In: Simon, A., Bennett, S. J., Castro, J. M. (eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools. American Geophysical Union, Washington, DC, pp. 6993.Google Scholar
Rosgen, D. L., 2014. River Stability Field Guide, 2nd edition. Wildland Hydrology, Fort Collins, CO.Google Scholar
Rosso, R., 1984. Nash model relation to Horton order ratios. Water Resources Research, 20(7), 914920.Google Scholar
Rouse, H., 1939. Discussion of “Laboratory investigation of flume traction and transportation. Transactions of the American Society of Civil Engineers, 104, 13031308.Google 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(9), 11611179.Google 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-Earth Surface, 114. 10.1029/2008jf001073.Google Scholar
Roy, A. G., 1983. Optimal angular geometry models of river branching. Geographical Analysis, 15(2), 8796.Google Scholar
Roy, A. G., Bergeron, N., 1990. Flow and particle paths in a natural river confluence with coarse bed material. Geomorphology, 3, 99112.Google Scholar
Roy, A. G., Roy, R., 1988. Short communication: changes in channel size at river confluences with coarse bed material. Earth Surface Processes and Landforms, 13, 7784.Google Scholar
Roy, A. G., Woldenberg, M. J., 1986. A model for changes in channel form at a river confluence. Journal of Geology, 94(3), 402411. 10.1086/629038.Google Scholar
Roy, A. G., Roy, R., Bergeron, N., 1988. Hydraulic geometry and changes in flow velocity at a river confluence with coarse bed material. Earth Surface Processes and Landforms, 13, 583598.Google Scholar
Roy, A. G., Buffin-Belanger, T., Deland, S., 1996. Scales of coherent turbulent flow structures in a gravel-bed river. In: Ashworth, P. J., Bennett, S. J., Best, J. L., McLelland, S. (eds.), Coherent Flow Structures in Open Channels. John Wiley, New York, pp. 147164.Google Scholar
Roy, A. G., Buffin-Belanger, T., Lamarre, H., Kirkbride, A. D., 2004. Size, shape and dynamics of large-scale turbulent flow structures in a gravel-bed river. Journal of Fluid Mechanics, 500, 127.Google Scholar
Roy, N. G., Sinha, R., 2014. Effective discharge for suspended sediment transport of the Ganga River and its geomorphic implication. Geomorphology, 227, 1830.Google Scholar
Royall, D., 2013. Land use impacts on the hydrogeomorphology of small watersheds. In: Shroder, J. C. (ed.), Treatise on Geomorphology, Vol. 13, Geomorphology of Human Disturbances, Climate Change, and Natural Hazards, James, L. A., Harden, C. P., Clague, J. J. (vol. eds.). Academic Press, San Diego, CA, pp. 2847.Google Scholar
Royden, L., Perron, J. T., 2013. Solutions of the stream power equation and application to the evolution of river longitudinal profiles. Journal of Geophysical Research-Earth Surface, 118(2), 497518. 10.1002/jgrf.20031.Google Scholar
Rozovskii, I.L., 1957. Flow of Water in Bends of Open Channels. Academy of Sciences of the Ukrainian S.S.R., Kiev. (Translated from Russian by Y. Prushansky, Israel Program for Scientific Translations, 1961.)Google Scholar
Rubey, W. W., 1933a. Settling velocities of gravel, sand, and silt particles. American Journal of Science, 25(148), 325338.Google Scholar
Rubey, W. W., 1933b. Equilibrium-conditions in debris-laden streams. Transactions, American Geophysical Union, 14, 497505.Google Scholar
Rubey, W. W., 1952. Geology and mineral resources of the Hardin and Brussels quadrangles (in Illinois). U.S. Geological Survey Professional Paper 218. U.S. Government Printing Office, Washington, DC.Google Scholar
Ruhe, R. V., 1952. Topographic discontinuities of the Des Moines Lobe. American Journal of Science, 250(1), 4656.Google Scholar
Russell, K. L., Vietz, G. J., Fletcher, T. D., 2017. Global sediment yields from urban and urbanizing watersheds. Earth-Science Reviews, 168, 7380.Google Scholar
Russell, K. L., Vietz, G. J., Fletcher, T. D., 2018. Urban catchment runoff increases bedload sediment yield and particle size in stream channels. Anthropocene, 23, 5366.Google Scholar
Rust, B. R., 1981. Sedimentation in an arid-zone anastomosing fluvial system – Cooper’s Creek, central Australia. Journal of Sedimentary Petrology, 51(3), 745755.Google Scholar
Ruther, N., Olsen, N. R. B., 2007. Modelling free-forming meander evolution in a laboratory channel using three-dimensional computational fluid dynamics. Geomorphology, 89(3–4), 308319.Google Scholar
Rutherford, J. C., 1994. River Mixing. Wiley, Chichester, UK.Google Scholar
Sack, D., 2013. Geomorphology and nineteenth century explorations of the American West In: Shroder, J. W. (ed.), Treatise on Geomorphology, Vol. 1, The Foundations of Geomorphology, Orme, A.R., Sack, D. (vol. eds.). Elsevier, New York, pp. 5363.Google Scholar
Saletti, M., Molnar, P., Zimmermann, A., Hassan, M. A., Church, M., 2015. Temporal variability and memory in sediment transport in an experimental step-pool channel. Water Resources Research, 51(11), 93259337. 10.1002/2015wr016929.Google Scholar
Saletti, M., Molnar, P., Hassan, M. A., Burlando, P., 2016. A reduced-complexity model for sediment transport and step-pool morphology. Earth Surface Dynamics, 4(3), 549566.Google Scholar
Sambrook Smith, G. H., Ferguson, R. I., 1995. The gravel sand transition along river channels. Journal of Sedimentary Research Section a – Sedimentary Petrology and Processes, 65(2), 423430.Google Scholar
Sambrook Smith, G. H., Ferguson, R. I., 1996. The gravel-sand transition: flume study of channel response to reduced slope. Geomorphology, 16(2), 147159.Google Scholar
Sambrook Smith, G. H., Ashworth, P. J., Best, J. L., Woodward, J., Simpson, C. J., 2005. The morphology and facies of sandy braided rivers: some considerations of scale invariance. In: Blum, M. D., Marriott, S. B., Leclair, S. F. (eds.), Fluvial Sedimentology VII, International Association of Sedimentologists Special Publication No. 35. Blackwell, Malden, MA, pp. 145158.Google Scholar
Sambrook Smith, G. H., Ashworth, P. J., Best, J. L., Woodward, J., Simpson, C. J., 2006. The sedimentology and alluvial architecture of the sandy braided South Saskatchewan River, Canada. Sedimentology, 53(2), 413434.Google Scholar
Sambrook Smith, G. H., Best, J. L., Leroy, J. Z., Orfeo, O., 2016. The alluvial architecture of a suspended sediment dominated meandering river: the Rio Bermejo, Argentina. Sedimentology, 63(5), 11871208.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.Google Scholar
Sangireddy, H., Carothers, R. A., Stark, C. P., Passalacqua, P., 2016a. Controls of climate, topography, vegetation, and lithology on drainage density extracted from high resolution topography data. Journal of Hydrology, 537, 271282.Google Scholar
Sangireddy, H., Stark, C. P., Kladzyk, A., Passalacqua, P., 2016b. GeoNet: an open source software for the automatic and objective extraction of channel heads, channel network, and channel morphology from high resolution topography data. Environmental Modelling & Software, 83, 5873.Google Scholar
Sapozhnikov, V., Foufoula-Georgiou, E., 1996. Self-affinity in braided rivers. Water Resources Research, 32(5), 14291439.Google Scholar
Sapozhnikov, V. B., Foufoula-Georgiou, E., 1999. Horizontal and vertical self-organization of braided rivers toward a critical state. Water Resources Research, 35(3), 843851.Google Scholar
Saucier, R. T., 1994a. Evidence of late glacial runoff in the lower Mississippi Valley. Quaternary Science Reviews, 13(9–10), 973981.Google Scholar
Saucier, R. T., 1994b. Geomorphology and Quaternary geologic history of the lower Mississippi Valley. U.S. Army Corps of Engineers, Vicksburg, MS.Google Scholar
Sawyer, A. M., Pasternack, G. B., Moir, H. J., Fulton, A. A., 2010. Riffle-pool maintenance and flow convergence routing observed on a large gravel-bed river. Geomorphology, 114(3), 143160.Google Scholar
Scheidegger, A. E., 1967. A stochastic model for drainage patterns into an intramontane trench. Bulletin for the Association of Scientific Hydrology, 12, 1520.Google Scholar
Scheingross, J. S., Lamb, M. P., 2017a. A mechanistic model of waterfall plunge pool erosion into bedrock. Journal of Geophysical Research-Earth Surface, 122(11), 20792104.Google Scholar
Scheingross, J. S., Lo, D. Y., Lamb, M. P., 2017b. Self-formed waterfall plunge pools in homogeneous rock. Geophysical Research Letters, 44(1), 200208.Google Scholar
Schick, A. P., Hassan, M. A., Lekach, J., 1987. A vertical exchange model for coarse bedload movement – numerical considerations. Catena Supplement, 10, 7383.Google Scholar
Schiefer, E., Hassan, M. A., Menounos, B., Pelpola, C. P., Slaymaker, O., 2010. Interdecadal patterns of total sediment yield from a montane catchment, southern Coast Mountains, British Columbia, Canada. Geomorphology, 118(1–2), 207212.Google Scholar
Schielen, R., Doelman, A., De Swart, H. E., 1993. On the nonlinear dynamics of free bars in straight channels. Journal of Fluid Mechanics, 252, 325356.Google Scholar
Schindfessel, L., Creelle, S., De Mulder, T., 2015. Flow patterns in an open channel confluence with increasingly dominant tributary inflow. Water, 7(9), 47244751.Google Scholar
Schindfessel, L., Creelle, S., De Mulder, T., 2017. How different cross-sectional shapes influence the separation zone of an open-channel confluence. Journal of Hydraulic Engineering, 143(9). 10.1061/(asce)hy.1943–7900.0001336.Google Scholar
Schlichting, H., Gersten, K., 2016. Boundary-Layer Theory. Springer, Berlin.Google Scholar
Schmeeckle, M. W., Nelson, J. M., Shreve, R. L., 2007. Forces on stationary particles in near-bed turbulent flows. Journal of Geophysical Research-Earth Surface, 112(F2). 10.1029/2006jf000536.Google Scholar
Schmidt, J. C., Wilcock, P. R., 2008. Metrics for assessing the downstream effects of dams. Water Resources Research, 44(4). 10.1029/2006wr005092.Google Scholar
Schmidt, K. H., Ergenzinger, P., 1992. Bedload entrainment, travel lengths, step lengths, rest periods – studied with passive (iron, magnetic) and active (radio) tracer techniques. Earth Surface Processes and Landforms, 17(2), 147165.Google Scholar
Schneider, A., Gerke, H. H., Maurer, T., Nenov, R., 2013. Initial hydro-geomorphic development and rill network evolution in an artificial catchment. Earth Surface Processes and Landforms, 38(13), 14961512.Google Scholar
Schorghofer, N., Jensen, B., Kudrolli, A., Rothman, D. H., 2004. Spontaneous channelization in permeable ground: theory, experiment, and observation. Journal of Fluid Mechanics, 503, 357374.Google Scholar
Schroder, R., 1991. Test of Hack’s slope to bed material relationship in the southern Eifel Uplands, Germany. Earth Surface Processes and Landforms, 16(8), 731736.Google Scholar
Schumann, R. R., 1989. Morphology of Red Creek, Wyoming, an arid-region anastomosing channel system. Earth Surface Processes and Landforms, 14(4), 277288.Google Scholar
Schumm, S. A., 1956. Evolution of drainage systems and slopes in badlands at Perth Amboy, New Jersey. Geological Society of America Bulletin, 67(5), 597646.Google Scholar
Schumm, S. A., 1960. The shape of alluvial channels in relation to sediment type. U.S. Geological Survey Professional Paper 352-B, U.S. Government Printing Office, Washington, DC.Google Scholar
Schumm, S. A., 1968. River adjustment to altered hydrologic regime – Murrumbidgee River and paleochannels. U.S. Geological Survey Professional Paper 598. U.S. Government Printing Office, Washington, DC.Google Scholar
Schumm, S. A., 1969. River metamorphosis. Journal of the Hydraulics Division – ASCE, HY1, 255273.Google Scholar
Schumm, S. A., 1971. Fluvial geomorphology: channel adjustment and river metamorphosis. In: Shen, H. W. (ed.), River Mechanics, Vol. 1. H. W. Shen, Ft. Collins, CO, pp. 5.15.22.Google Scholar
Schumm, S. A., 1977. The Fluvial System. Wiley, New York.Google Scholar
Schumm, S. A., 1979. Geomorphic thresholds – concepts and its application. Transactions of the Institute of British Geographers, 4(4), 485515.Google Scholar
Schumm, S. A., 1981. Evolution and response of the fluvial system, sedimentological implications. Society of Economic Paleontologists and Mineralogists Special Publication No. 31, 1929.Google Scholar
Schumm, S.A., 1985. Patterns of alluvial rivers. Annual Review of Earth and Planetary Sciences, 13, 527.Google Scholar
Schumm, S. A., 1993. River response to baselevel change – implications for sequence stratigraphy. Journal of Geology, 101(2), 279294.Google Scholar
Schumm, S. A., Khan, H. R., 1972. Experimental study of channel patterns. Geological Society of America Bulletin, 83(6), 17551770.Google Scholar
Schumm, S. A., Lichty, R. W., 1963. Channel widening and floodplain construction along the Cimarron River in southwestern Kansas. U.S. Geological Survey Professional Paper 352-D. U.S. Government Printing Office, Washington, DC.Google Scholar
Schumm, S. A., Lichty, R. W., 1965. Time, space, and causality in geomorphology. American Journal of Science, 263(2), 110119.Google Scholar
Schumm, S. A., Rea, D. K., 1995. Sediment yield from disturbed earth systems. Geology, 23(5), 391394.Google Scholar
Schumm, S. A., Stevens, M. A., 1973. Abrasion in place: a mechanism for rounding and size reduction of coarse sediment in rivers. Geology, 1, 3740.Google Scholar
Schumm, S. A., Harvey, M. D., Watson, C. C., 1984. Incised Channels: Morphology, Dynamics and Controls. Water Resources Publications, Littleton, CO.Google Scholar
Schumm, S. A., Mosley, M. P., Weaver, W. E., 1987. Experimental Fluvial Geomorphology. Wiley, Chichester, UK.Google Scholar
Schumm, S. A., Boyd, K. F., Wolff, C. G., Spitz, W. J., 1995. A groundwater sapping landscape in the Florida panhandle. Geomorphology, 12(4), 281297.Google Scholar
Schumm, S. A., Erskine, W. D., Tilleard, J. W., 1996. Morphology, hydrology, and evolution of the anastomosing Owens and King Rivers, Victoria, Australia. Geological Society of America Bulletin, 108(10), 12121224.Google Scholar
Schuurman, F., Kleinhans, M. G., 2015. Bar dynamics and bifurcation evolution in a modelled braided sand-bed river. Earth Surface Processes and Landforms, 40(10), 13181333.Google Scholar
Schuurman, F., Marra, W. A., Kleinhans, M. G., 2013. Physics-based modeling of large braided sand-bed rivers: bar pattern formation, dynamics, and sensitivity. Journal of Geophysical Research-Earth Surface, 118(4), 25092527. 10.1002/2013jf002896.Google Scholar
Schuurman, F., Shimizu, Y., Iwasaki, T., Kleinhans, M. G., 2016. Dynamic meandering in response to upstream perturbations and floodplain formation. Geomorphology, 253, 94109.Google Scholar
Schwanghart, W., Scherler, D., 2017. Bumps in river profiles: uncertainty assessment and smoothing using quantile regression techniques. Earth Surface Dynamics, 5(4), 821839.Google Scholar
Schwartz, J. S., Herricks, E. E., 2007. Evaluation of pool-riffle naturalization structures on habitat complexity and the fish community in an urban Illinois stream. River Research and Applications, 23(4), 451466.Google Scholar
Schwendel, A. C., Nicholas, A. P., Aalto, R. E., Smith, G. H. S., Buckley, S., 2015. Interaction between meander dynamics and floodplain heterogeneity in a large tropical sandbed river: the Rio Beni, Bolivian Amazon. Earth Surface Processes and Landforms, 40(15), 20262040.Google Scholar
Schwendel, A., Aalto, R., Nicholas, A. P., Parsons, D., 2018. Fill characteristics of abandoned river channels and resulting stratigraphy of a mobile sand-bed river floodplain. In: Ghinassi, M., Columbera, L., Mountney, N. P., Reesink, A. J. H. (eds.), Fluvial Meanders and their Sedimentary Products in the Rock Record, International Association of Sedimentologists Special Publication 48. Wiley and Sons, Hoboken, NJ, pp. 251272.Google Scholar
Schwenk, J., Foufoula-Georgiou, E., 2016. Meander cutoffs nonlocally accelerate upstream and downstream migration and channel widening. Geophysical Research Letters, 43(24), 1243712445.Google Scholar
Schwenk, J., Lanzoni, S., Foufoula-Georgiou, E., 2015. The life of a meander bend: connecting shape and dynamics via analysis of a numerical model. Journal of Geophysical Research – Earth Surface, 120(4), 690710. 10.1002/2014jf003252.Google Scholar
Schwenk, J., Khandelwal, A., Fratkin, M., Kumar, V., Foufoula-Georgiou, E., 2017. High spatiotemporal resolution of river planform dynamics from Landsat: the RivMAP toolbox and results from the Ucayali River. Earth and Space Science, 4(2), 4675. 10.1002/2016ea000196.Google Scholar
Schwindt, S., Pasternack, G. B., Bratovich, P. M., Rabone, G., Simodynes, D., 2019. Hydro-morphological parameters generate lifespan maps for stream restoration management. Journal of Environmental Management, 232, 475489.Google Scholar
Scorpio, V., Zen, S., Bertoldi, W., et al., 2018. Channelization of a large alpine river: what is left of its original morphodynamics? Earth Surface Processes and Landforms, 43(5), 10441062.Google Scholar
Scown, M. W., Thoms, M. C., De Jager, N. R., 2015. Floodplain complexity and surface metrics: influences of scale and geomorphology. Geomorphology, 245, 102116.Google Scholar
Scown, M. W., Thoms, M. C., De Jager, N. R., 2016. An index of floodplain surface complexity. Hydrology and Earth System Sciences, 20(1), 431441.Google Scholar
Seal, R., Paola, C., 1995. Observations of downstream fining on the North Fork Toutle River near Mount St. Helens, Washington. Water Resources Research, 31(5), 14091419.Google Scholar
Seal, R., Paola, C., Parker, G., Southard, J. B., Wilcock, P. R., 1997. Experiments on downstream fining of gravel. 1. Narrow-channel runs. Journal of Hydraulic Engineering, 123(10), 874884.Google Scholar
Sear, D. A., 1994. River restoration and geomorphology. Aquatic Conservation – Marine and Freshwater Ecosystems, 4(2), 169177.Google Scholar
Sear, D. A., 1996. Sediment transport processes in pool-riffle sequences. Earth Surface Processes and Landforms, 21(3), 241262.Google Scholar
Sear, D., Newson, M., Hill, C., Old, J., Branson, J., 2009. A method for applying fluvial geomorphology in support of catchment-scale river restoration planning. Aquatic Conservation – Marine and Freshwater Ecosystems, 19(5), 506519.Google Scholar
Sear, D. A., Millington, C. E., Kitts, D. R., Jeffries, R., 2010. Logjam controls on channel: floodplain interactions in wooded catchments and their role in the formation of multi-channel patterns. Geomorphology, 116(3–4), 305319.Google Scholar
Searcy, J. K., 1959. Flow-duration curves. U.S. Geological Survey Water-supply Paper 1542-A. U.S. Government Printing Office, Washington, DC.Google Scholar
Selby, M., 1985. Earth’s Changing Surface: An Introduction to Geomorphology. Clarendon Press, Oxford, UK.Google Scholar
Selley, R. C., 2000. Applied Sedimentology. Academic Press, San Diego, CA.Google Scholar
Seminara, G., Tubino, M., 1989. Alternate bars and meandering: free, forced and mixed interactions. In: Ikeda, S., Parker, G. (eds.), River Meandering. Water Resources Monograph 12. American Geophysical Union, Washington, DC, pp. 267320.Google Scholar
Seminara, G., Tubino, M., 1992. Weakly nonlinear theory of regular meanders. Journal of Fluid Mechanics, 244, 257288.Google Scholar
Seminara, G., Zolezzi, G., Tubino, M., Zardi, D., 2001. Downstream and upstream influence in river meandering. Part 2. Planimetric development. Journal of Fluid Mechanics, 438, 213230.Google Scholar
Sennatt, K. M., Salant, N. L., Renshaw, C. E., Magilligan, F. J., 2006. Assessment of methods for measuring embeddedness: application to sedimentation in flow regulated streams. Journal of the American Water Resources Association, 42(6), 16711682.Google Scholar
Seto, K. C., Fragkias, M., Guneralp, B., Reilly, M. K., 2011. A meta-analysis of global urban land expansion. PLoS ONE, 6(8). 10.1371/journal.pone.0023777.Google Scholar
Seybold, H., Rothman, D. H., Kirchner, J. W., 2017. Climate’s watermark in the geometry of stream networks. Geophysical Research Letters, 44(5), 22722280. 10.1002/2016gl072089.Google Scholar
Shabayek, S., Steffler, P., Hicks, F., 2002. Dynamic model for subcritical combining flows in channel junctions. Journal of Hydraulic Engineering, 128, 821828.Google Scholar
Shakibainia, A., Tabatabai, M. R. M., Zarrati, A. R., 2010. Three-dimensional numerical study of flow structure in channel confluences. Canadian Journal of Civil Engineering, 37(5), 772781.Google Scholar
Shchepetkina, A., Gingras, M. K., Pemberton, S. G., 2015. The removal-cap suction corer: an inexpensive and durable device to extract unconsolidated, wet sediments. Journal of Sedimentary Research, 85(12), 14311437. 10.2110/jsr.2015.91.Google Scholar
Shelef, E., Hilley, G. E., 2013. Impact of flow routing on catchment area calculations, slope estimates, and numerical simulations of landscape development. Journal of Geophysical Research-Earth Surface, 118(4), 21052123. 10.1002/jgrf.20127.Google Scholar
Shen, C., Wang, S., Liu, X., 2016. Geomorphological significance of at-many-stations hydraulic geometry. Geophysical Research Letters, 43(8), 37623770. 10.1002/2016gl068364.Google Scholar
Shen, H., Zheng, F., Wen, L., Lu, J., Jiang, Y., 2015a. An experimental study of rill erosion and morphology. Geomorphology, 231, 193201.Google Scholar
Shen, Z., Tornqvist, T. E., Mauz, B., et al., 2015b. Episodic overbank deposition as a dominant mechanism of floodplain and delta-plain aggradation. Geology, 43(10), 875878.Google Scholar
Shepherd, R. G., 1985. Regression analysis of river profiles. Journal of Geology, 93(3), 377384.Google Scholar
Sherriff, S. C., Rowan, J. S., Fenton, O., et al., 2016. Storm event suspended sediment-discharge hysteresis and controls in agricultural watersheds: implications for watershed scale sediment management. Environmental Science & Technology, 50(4), 17691778.Google Scholar
Shields, F. D., Jr., Abt, S. R., 1989. Sediment deposition in cutoff meander bends and implications for effective management. Regulated Rivers Research and Management, 4(4), 381396.Google Scholar
Shields, F. D., Jr., Copeland, R. R., Klingeman, P. C., Doyle, M. W., Simon, A., 2003. Design for stream restoration. Journal of Hydraulic Engineering, 129(8), 575584.Google Scholar
Shields, F. D., Jr., Copeland, R. R., Klingeman, P. C., Doyle, M. W., Simon, A., 2008. Stream restoration. In: Garcia, M. H. (ed.), Sedimentation Engineering: Processes, Measurements, Modeling, and Practice. American Society of Civil Engineers, New York, pp. 461503.Google Scholar
Shih, W., Diplas, P., Celik, A. O., Dancey, C., 2017. Accounting for the role of turbulent flow on particle dislodgement via a coupled quadrant analysis of velocity and pressure sequences. Advances in Water Resources, 101, 3748.Google Scholar
Shiklomanov, I.A., 1993. World fresh water resources. In: Gleick, P. (ed.), Water in Crisis. Oxford University Press, New York, pp. 1323.Google Scholar
Shiono, K., Knight, D. W., 1991. Turbulent open channel flows with variable depth across the channel. Journal of Fluid Mechanics, 222, 617646.Google Scholar
Sholtes, J. S., Bledsoe, B. P., 2016. Half-yield discharge: process-based predictor of bankfull discharge. Journal of Hydraulic Engineering, 142(8). 10.1061/(asce)hy.1943–7900.0001137.Google Scholar
Sholtes, J., Werbylo, K., Bledsoe, B., 2014. Physical context for theoretical approaches to sediment transport magnitude-frequency analysis in alluvial channels. Water Resources Research, 50(10), 79007914. 10.1002/2014wr015639.Google Scholar
Shreve, R. L., 1966. Statistical law of stream numbers. Journal of Geology, 74(1), 1737.Google Scholar
Shreve, R. L., 1967. Infinite topologically random channel networks. Journal of Geology, 75(2), 178186.Google Scholar
Shulits, S., 1941. Rational equation of river-bed profile. Transactions, American Geophysical Union, 22, 622630.Google Scholar
Shvidchenko, A. B., Pender, G., 2000. Flume study of the effect of relative depth on the incipient motion of coarse uniform sediments. Water Resources Research, 36(2), 619628.Google Scholar
Shvidchenko, A. B., Pender, G., 2001b. Macroturbulent structure of open-channel flow over gravel beds. Water Resources Research, 37(3), 709719.Google Scholar
Shvidchenko, A. B., Pender, G., Hoey, T. B., 2001a. Critical shear stress for incipient motion of sand/gravel streambeds. Water Resources Research, 37(8), 22732283.Google Scholar
Sichingabula, H. M., 1999. Magnitude-frequency characteristics of effective discharge for suspended sediment transport, Fraser River, British Columbia, Canada. Hydrological Processes, 13(9), 13611380.Google Scholar
Simoes, F. J. M., 2014. Shear velocity criterion for incipient motion of sediment. Water Science and Engineering, 7, 183193.Google Scholar
Simon, A., 1989a. A model of channel response in disturbed alluvial channels. Earth Surface Processes and Landforms, 14(1), 1126.Google Scholar
Simon, A., 1989b. The discharge of sediment in channelized alluvial streams. Water Resources Bulletin, 25(6), 11771188.Google Scholar
Simon, A., 1992. Energy, time, and channel evolution in catastrophically disturbed fluvial systems. Geomorphology, 5(3–5), 345372.Google Scholar
Simon, A., Collison, A. J. C., 2002. Quantifying the mechanical and hydrologic effects of riparian vegetation on streambank stability. Earth Surface Processes and Landforms, 27(5), 527546.Google Scholar
Simon, A., Rinaldi, M., 2000. Channel instability in the loess area of the midwestern United States. Journal of the American Water Resources Association, 36(1), 133150.Google Scholar
Simon, A., Rinaldi, M., 2006. Disturbance, stream incision, and channel evolution: the roles of excess transport capacity and boundary materials in controlling channel response. Geomorphology, 79(3–4), 361383.Google Scholar
Simon, A., Rinaldi, M., 2013. Incised channels: disturbance, evolution and the roles of excess transport capacity and boundary materials in controlling channel response. In: Shroder, J. C. (ed.), Treatise on Geomorphology, Vol. 9, Fluvial Geomorphology, Wohl, E. (vol. ed.), Academic Press, San Diego, CA, pp. 574594.Google Scholar
Simon, A., Dickerson, W., Heins, A., 2004. Suspended-sediment transport rates at the 1.5-year recurrence interval for ecoregions of the United States: transport conditions at the bankfull and effective discharge. Geomorphology, 58 (1–4),243262.Google Scholar
Simon, A., Doyle, M., Kondolf, M., et al., 2007. Critical evaluation of how the Rosgen classification and associated “natural channel design” methods fail to integrate and quantify fluvial processes and channel response. Journal of the American Water Resources Association, 43(5), 11171131.Google Scholar
Simons, D. B., Richardson, E. V., 1966. Resistance to flow in alluvial channels. U.S. Geological Survey Professional Paper 422-J. U.S. Government Printing Office, Washington, DC.Google Scholar
Simpson, G., Schlunegger, F., 2003. Topographic evolution and morphology of surfaces evolving in response to coupled fluvial and hillslope sediment transport. Journal of Geophysical Research – Solid Earth, 108(B6), 10.1029/2002jb002162.Google Scholar
Singer, M. B., 2008a. Downstream patterns of bed material grain size in a large, lowland alluvial river subject to low sediment supply. Water Resources Research, 44(12). 10.1029/2008wr007183.Google Scholar
Singer, M.B., 2008b. A new sampler for extracting bed material sediment from sand and gravel beds in navigable rivers. Earth Surface Processes and Landforms, 33(14), 22772284.Google Scholar
Singh, U., Crosato, A., Giri, S., Hicks, M., 2017. Sediment heterogeneity and mobility in the morphodynamic modelling of gravel-bed braided rivers. Advances in Water Resources, 104, 127144.Google Scholar
Singh, V. P., 2003. On the theories of hydraulic geometry. International Journal of Sediment Research, 18, 196218.Google Scholar
Singh, V. P., Zhang, L., 2008a. At-a-station hydraulic geometry relations, 1: theoretical development. Hydrological Processes, 22(2), 189215.Google Scholar
Singh, V.P., Zhang, L., 2008b. At-a-station hydraulic geometry relations, 2: calibration and testing. Hydrological Processes, 22(2), 216228.Google Scholar
Singh, V. P., Yang, C. T., Deng, Q.-Z., 2003. Downstream hydraulic geometry relations: 2. calibration and testing. Water Resources Research, 12. 10.1029/2003WR002498.Google Scholar
Sinha, S. K., Parker, G., 1996. Causes of concavity in longitudinal profiles of rivers. Water Resources Research, 32(5), 14171428.Google Scholar
Siviglia, A., Stecca, G., Vanzo, D., et al., 2013. Numerical modelling of two-dimensional morphodynamics with applications to river bars and bifurcations. Advances in Water Resources, 52, 243260.Google Scholar
Skalak, K., Pizzuto, J., Hart, D. D., 2009. Influence of small dams on downstream channel characteristics in Pennsylvania and Maryland: implications for the long-term geomorphic effects of dam removal. Journal of the American Water Resources Association, 45(1), 97109.Google Scholar
Sklar, L., Dietrich, W. E., 1998. River longitudinal profiles and bedrock incision models: stream power and the influence of sediment supply. In: Tinkler, K., Wohl, E. E. (eds.), Rivers over Rock: Fluvial Process in Bedrock Channels, 107. American Geophysical Union, Washington, DC, pp. 237260.Google Scholar
Sklar, L. S., Dietrich, W. E., 2001. Sediment and rock strength controls on river incision into bedrock. Geology, 29(12), 10871090.Google Scholar
Sklar, L. S., Dietrich, W. E., 2004. A mechanistic model for river incision into bedrock by saltating bed load. Water Resources Research, 40(6). 10.1029/2003wr002496.Google Scholar
Sklar, L. S., Dietrich, W. E., 2008. Implications of the saltation-abrasion bedrock incision model for steady-state river longitudinal profile relief and concavity. Earth Surface Processes and Landforms, 33(7), 11291151.Google Scholar
Sklar, L. S., Dietrich, W. E., 2012. Correction to “A mechanistic model for river incision into bedrock by saltating bed load”. Water Resources Research, 48. 10.1029/2012wr012267.Google Scholar
Skolasinska, K., 2014. Inquiry of levee formation by grain size analysis – a case study from the Warta River (central Poland). Catena, 122, 103110.Google Scholar
Slater, L. J., Singer, M. B., 2013. Imprint of climate and climate change in alluvial riverbeds: continental United States, 1950–2011. Geology, 41(5), 595598.Google Scholar
Slattery, M. C., Bryan, R. B., 1992. Hydraulic conditions for rill incision under simulated rainfall – a laboratory experiment. Earth Surface Processes and Landforms, 17(2), 127146.Google Scholar
Slaymaker, O., 2003. The sediment budget as conceptual framework and management tool. Hydrobiologia, 494(1–3), 7182.Google Scholar
Slaymaker, O., 2006. Towards the identification of scaling relations in drainage basin sediment budgets. Geomorphology, 80(1–2), 819.Google Scholar
Slingerland, R., Smith, N. D., 1998. Necessary conditions for a meandering-river avulsion. Geology, 26(5), 435438.Google Scholar
Slowik, M., 2018. The formation of an anabranching planform in a sandy floodplain by increased flows and sediment load. Earth Surface Processes and Landforms, 43(3), 623638.Google Scholar
Small, M. J., Doyle, M. W., 2012. Historical perspectives on river restoration design in the USA. Progress in Physical Geography, 36(2), 138153.Google Scholar
Smart, J. S., 1968. Statistical properties of stream lengths. Water Resources Research, 4(5), 10011014.Google Scholar
Smith, B., Clifford, N. J., Mant, J., 2014. The changing nature of river restoration. Wiley Interdisciplinary Reviews – Water, 1(3), 249261.Google Scholar
Smith, C. E., 1998. Modeling high sinuosity meanders in a small flume. Geomorphology, 25(1–2), 1930.Google Scholar
Smith, D. G., 1976. Effect of vegetation on lateral migration of anastomosed channels of a glacier meltwater river. Geological Society of America Bulletin, 87(6), 857860.Google Scholar
Smith, D. G., 1983. Anastomosed fluvial deposits: examples from western Canada. In: Collinson, J. D., Lewin, J. (eds.), Modern and Ancient Fluvial Systems, Special Publication of the International Association of Sedimentologists No. 6, Blackwell, Oxford, UK, pp. 155168.Google Scholar
Smith, D. G., 1984. Vibracoring fluvial and deltaic sediments – tips on improving penetration and recovery. Journal of Sedimentary Petrology, 54(2), 660663.Google Scholar
Smith, D. G., 1986. Anastomosing river deposits, sedimentation rates and basin subsidence, Magdalena River, northwestern Columbia, South America. Sedimentary Geology, 46(3–4), 177196.Google Scholar
Smith, D., Elmore, A. C., 2014. A modification of freeze-core technology for collecting granular fluvial sediment samples. Environmental Earth Sciences, 71(9), 41494156.Google Scholar
Smith, D. G., Smith, N. D., 1980. Sedimentation in anastomosed river systems: examples from alluvial valleys near Banff, Alberta. Journal of Sedimentary Petrology, 50, 157164.Google Scholar
Smith, D. G., Hubbard, S. M., Leckie, D. A., Fustic, M., 2009a. Counter point bar deposits: lithofacies and reservoir significance in the meandering modern Peace River and ancient McMurray Formation, Alberta, Canada. Sedimentology, 56(6), 16551669. 10.1111/j.1365–3091.2009.01050.x.Google Scholar
Smith, L. C., Isacks, B. L., Bloom, A. L., Murray, A. B., 1996. Estimation of discharge from three braided rivers using synthetic aperture radar satellite imagery: potential application to ungaged basins. Water Resources Research, 32(7), 20212034.Google Scholar
Smith, N. D., 1974. Sedimentology and bar formation in upper Kicking Horse River, a braided outwash stream. Journal of Geology, 82(2), 205223.Google Scholar
Smith, N. D., Perez-Arlucea, M., 1994. Fine-grained splay deposition in the avulsion belt of the lower Saskatchewan River, Canada. Journal of Sedimentary Research Section B – Stratigraphy and Global Studies, 64(2), 159168.Google Scholar
Smith, N. D., Perez-Arlucea, M., 2008. Natural levee deposition during the 2005 flood of the Saskatchewan River. Geomorphology, 101(4), 583594.Google Scholar
Smith, N. D., Smith, D. G., 1984. William River – an outstanding example of channel widening and braiding caused by bedload addition. Geology, 12(2), 7882.Google Scholar
Smith, N. D., Cross, T. A., Dufficy, J. P., Clough, S. R., 1989. Anatomy of an avulsion. Sedimentology, 36(1), 123.Google Scholar
Smith, N. D., McCarthy, T. S., Ellery, W. N., Merry, C. L., Ruther, H., 1997. Avulsion and anastomosis in the panhandle region of the Okavango Fan, Botswana. Geomorphology, 20(1–2), 4965.Google Scholar
Smith, N. D., Slingerland, R. L., Perez-Arlucea, M., Morozova, G. S., 1998. The 1870s avulsion of the Saskatchewan River. Canadian Journal of Earth Sciences, 35(4), 453466. 10.1139/e97-113.Google Scholar
Smith, N. D., Perez-Arlucea, M., Edmonds, D. A., Slingerland, R. L., 2009b. Elevation adjustments of paired natural levees during flooding of the Saskatchewan River. Earth Surface Processes and Landforms, 34(8), 10601068. 10.1002/esp.1792.Google Scholar
Smith, S. M., Prestegaard, K. L., 2005. Hydraulic performance of a morphology-based stream channel design. Water Resources Research, 41(11). 10.1029/2004wr003926.Google Scholar
Smith, T. R., 2010. A theory for the emergence of channelized drainage. Journal of Geophysical Research – Earth Surface, 115. 10.1029/2008jf001114.Google Scholar
Smith, T. R., Bretherton, F. P., 1972. Stability and the conservation of mass in drainage basin evolution. Water Resources Research, 8, 15061529.Google Scholar
Sneddon, C. S., Magilligan, F. J., Fox, C. A., 2017. Science of the dammed: expertise and knowledge claims in contested dam removals. Water Alternatives-an Interdisciplinary Journal on Water Politics and Development, 10(3), 677696.Google Scholar
Snow, R. S., Slingerland, R. L., 1987. Mathematical modeling of graded river profiles. Journal of Geology, 95(1), 1533.Google Scholar
Snyder, N. P., Whipple, K. X., Tucker, G. E., Merritts, D. J., 2003. Importance of a stochastic distribution of floods and erosion thresholds in the bedrock river incision problem. Journal of Geophysical Research – Solid Earth, 108(B2). 10.1029/2001jb001655.Google Scholar
Soar, P. J., Thorne, C. R., 2011. Design discharge for river restoration. In: Simon, A., Bennett, S. J., Castro, J. M. (eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools. American Geophysical Union, Washington, DC, pp. 123149.Google Scholar
Song, X., Bai, Y., 2015. A new empirical river pattern discriminant method based on flow resistance characteristics. Catena, 135, 163172.Google Scholar
Soong, D. T., Prater, C. D., Halfar, T. M., Wobig, L. A., 2012. Manning’s roughness coefficients for Illinois streams. U.S. Geological Survey Data Series 668. U.S. Geological Survey, Reston, VA.Google Scholar
Spink, A., Hillman, M., Fryirs, K., Brierley, G., Lloyd, K., 2010. Has river rehabilitation begun? Social perspectives from the Upper Hunter catchment, New South Wales, Australia. Geoforum, 41(3), 399409.Google Scholar
Stanistreet, I. G., McCarthy, T. S., 1993. The Okavango Fan and the classification of subaerial fan systems. Sedimentary Geology, 85(1–4), 115133.Google Scholar
Stanistreet, I. G., Cairncross, B., McCarthy, T. S., 1993. Low sinuosity and meandering bedload rivers of the Okavango Fan: channel confinement by vegetated levees. Sedimentary Geology, 85(1–4), 135156.Google Scholar
Stanley, E. H., Doyle, M. W., 2003. Trading off: the ecological removal effects of dam removal. Frontiers in Ecology and the Environment, 1(1), 1522.Google Scholar
Stark, C. P., 1991. An invasion percolation model of drainage network evolution. Nature, 352(6334), 423425.Google Scholar
Stark, C. P., Barbour, J. R., Hayakawa, Y. S., et al., 2010. The climatic signature of incised river meanders. Science, 327(5972), 14971501. 10.1126/science.1184406.Google Scholar
Stecca, G., Measures, R., Hicks, D. M., 2017. A framework for the analysis of noncohesive bank erosion algorithms in morphodynamic modeling. Water Resources Research, 53(8), 66636686. 10.1002/2017wr020756.Google Scholar
Steffen, W., Persson, A., Deutsch, L., et al., 2011. The Anthropocene: from global change to planetary stewardship. Ambio, 40(7), 739761.Google Scholar
Steffen, W., Broadgate, W., Deutsch, L., Gaffney, O., Ludwig, C., 2015. The trajectory of the Anthropocene: the Great Acceleration. Anthropocene Review, 2(1), 8198.Google Scholar
Stein, O. R., Julien, P. Y., 1993. Criterion delineating the mode of the headcut migration. Journal of Hydraulic Engineering, 119(1), 3750.Google Scholar
Stein, O. R., LaTray, D. A., 2002. Experiments and modeling of head cut migration in stratified soils. Water Resources Research, 38(12). 10.1029/2001wr001166.Google Scholar
Sternberg, H., 1875. Untersuchen uber Langen- und Querprofile geschiebefuhrende Flusse. Zeitschrift Fur Bauwesen, 25, 483506.Google Scholar
Stevaux, J. C., Souza, I. A., 2004. Floodplain construction in an anastomosed river. Quaternary International, 114, 5565.Google Scholar
Stevens, A. J., Clarke, D., Nicholls, R. J., 2016. Trends in reported flooding in the UK: 1884–2013. Hydrological Sciences Journal, 61(1), 5063.Google Scholar
Stevens, M. A., Simons, D. B., Richardson, E. V., 1975. Nonequilibrium river form. Journal of the Hydraulics Division-ASCE, 101(NHY5), 557566.Google Scholar
Stewardson, M., 2005. Hydraulic geometry of stream reaches. Journal of Hydrology, 306(1–4), 97111.Google Scholar
Stewardson, M., Rutherford, I., 2008. Conceptual and mathematical modelling in river restoration: do we have unreasonable confidence? In: Darby, S. E., Sear, D. A. (eds.), River Restoration: Managing for Uncertainty in Restoring Physical Habitat. Wiley, Chichester, UK, pp. 6178.Google Scholar
Stolum, H. H., 1996. River meandering as a self-organization process. Science, 271(5256), 17101713.Google Scholar
Stout, J. C., Belmont, P., 2014. TerEx Toolbox for semi-automated selection of fluvial terrace and floodplain features from lidar. Earth Surface Processes and Landforms, 39(5), 569580.Google Scholar
Stout, J. C., Belmont, P., Schottler, S. P., Willenbring, J. K., 2014. Identifying sediment sources and sinks in the Root River, southeastern Minnesota. Annals of the Association of American Geographers, 104(1), 2039.Google Scholar
Stouthamer, E., Berendsen, H. J. A., 2001. Avulsion frequency, avulsion duration, and interavulsion period of Holocene channel belts in the Rhine-Meuse Delta, the Netherlands. Journal of Sedimentary Research, 71(4), 589598.Google Scholar
Stouthamer, E., Berendsen, H. J. A., 2007. Avulsion: the relative roles of autogenic and allogenic processes. Sedimentary Geology, 198(3–4), 309325.Google Scholar
Strahler, A. N., 1952a. Dynamic basis of geomorphology. Geological Society of America Bulletin, 63, 923938.Google Scholar
Strahler, A. N., 1952b. Hypsometric (area-altitude) analysis of erosional topography. Geological Society of America Bulletin, 63(11), 11171142.Google Scholar
Strahler, A. N., 1965. Introduction to Physical Geography. Wiley, New York.Google Scholar
Stranko, S. A., Hilderbrand, R. H., Palmer, M. A., 2012. Comparing the fish and benthic macroinvertebrate diversity of restored urban streams to reference streams. Restoration Ecology, 20(6), 747755.Google Scholar
Strick, R. J. P., Ashworth, P. J., Awcock, G., Lewin, J., 2018. Morphology and spacing of river meander scrolls. Geomorphology, 310, 5768.Google Scholar
Strom, K. B., Papanicolaou, A. N., 2008. Morphological characterization of cluster microforms. Sedimentology, 55(1), 137153.Google Scholar
Stubblefield, A. P., Reuter, J. E., Goldman, C. R., 2009. Sediment budget for subalpine watersheds, Lake Tahoe, California, USA. Catena, 76(3), 163172.Google Scholar
Sui, B., Huang, S.-h., 2017. Numerical analysis of flow separation zone in a confluent meander bend channel. Journal of Hydrodynamics, 29(4), 716723.Google Scholar
Sukhodolov, A. N., 2015. Field based research in fluvial hydraulics: potential, paradigms and challenges. Journal of Hydraulic Research, 53(1), 119.Google Scholar
Sukhodolov, A. N., Rhoads, B. L., 2001. Field investigation of three-dimensional flow structure at stream confluences 2. Turbulence. Water Resources Research, 37(9), 24112424.Google Scholar
Sukhodolov, A. N., Sukhodolova, T. A., 2019. Dynamics of flow at concordant gravel bed river confluences: effects of junction angle and momentum flux ratio. Journal of Geophysical Research – Earth Surface, 124(2), 588615. 10.1029/2018jf004648.Google Scholar
Sukhodolov, A., Thiele, M., Bungartz, H., 1998. Turbulence structure in a river reach with sand bed. Water Resources Research, 34(5), 13171334.Google Scholar
Sukhodolov, A. N., Schnauder, I., Uijttewaal, W. S. J., 2010. Dynamics of shallow lateral shear layers: experimental study in a river with a sandy bed. Water Resources Research, 46. 10.1029/2010wr009245.Google Scholar
Sukhodolov, A. N., Krick, J., Sukhodolova, T. A., et al., 2017. Turbulent flow structure at a discordant river confluence: asymmetric jet dynamics with implications for channel morphology. Journal of Geophysical Research – Earth Surface, 122(6), 12781293. 10.1002/2016jf004126.Google Scholar
Summerfield, M. A., Hulton, N. J., 1994. Natural controls of fluvial denudation rates in major world drainage basins. Journal of Geophysical Research – Solid Earth, 99(B7), 1387113883.Google Scholar
Sun, T., Meakin, P., Jossang, T., Schwarz, K., 1996. A simulation model for meandering rivers. Water Resources Research, 32(9), 29372954.Google Scholar
Sun, T., Meakin, P., Jossang, T., 2001. A computer model for meandering rivers with multiple bed load sediment sizes 2. Computer simulations. Water Resources Research, 37(8), 22432258.Google Scholar
Surian, N., 2006. Effects of human impact on braided river morphology: examples from northern Italy. In: Sambrook Smith, G. H., Best, J. L., Bristow, C. S., Petts, G. E. (eds.), Braided Rivers: Process, Deposits, Ecology and Management, International Association of Sedimentologists Special Publication 36. Blackwell, Oxford, UK, pp. 327338.Google Scholar
Sutherland, A. J., 1987. Static armor layers by progressive erosion. In: Thorne, C. R., Bathurst, J.C., and Hey, R.D. (eds.), Sediment Transport in Gravel-bed Rivers. Wiley, Chichester, UK, pp. 243260.Google Scholar
Swanson, B. J., Meyer, G., 2014. Tributary confluences and discontinuities in channel form and sediment texture: Rio Chama, NM. Earth Surface Processes and Landforms, 39(14), 19271943.Google Scholar
Swanson, M. L., Kondolf, G. M., Boison, P. J., 1989. An example of rapid gully initiation and extension by subsurface erosion: coastal San Mateo County, California. Geomorphology, 2(4), 393403.Google Scholar
Sweet, W. V., Geratz, J. W., 2003. Bankfull hydraulic geometry relationships and recurrence intervals for North Carolina’s Coastal Plain. Journal of the American Water Resources Association, 39(4), 861871.Google Scholar
Sylvester, Z., Durkin, P., Covault, J. A., 2019. High curvatures drive river meandering. Geology, 47(3), 263266.Google Scholar
Syvitski, J. P. M., 2003. Sediment fluxes and rates of sedimentation. In: Middleton, G. (ed.), Encyclopedia of Sediments and Sedimentary Rocks. Kluwer Academic Publishers, Dordrecht, Netherlands, pp. 600606.Google Scholar
Syvitski, J. P. M., Kettner, A., 2011. Sediment flux and the Anthropocene. Philosophical Transactions of the Royal Society A – Mathematical Physical and Engineering Sciences, 369(1938), 957975.Google Scholar
Syvitski, J. P. M., Milliman, J. D., 2007. Geology, geography, and humans battle for dominance over the delivery of fluvial sediment to the coastal ocean. Journal of Geology, 115(1), 119.Google Scholar
Syvitski, J. P., Morehead, M. D., 1999. Estimating river-sediment discharge to the ocean: application to the Eel margin, northern California. Marine Geology, 154(1–4), 1328.Google Scholar
Syvitski, J. P. M., Saito, Y., 2007. Morphodynamics of deltas under the influence of humans. Global and Planetary Change, 57(3–4), 261282.Google Scholar
Syvitski, J. P. M., Peckham, S. D., Hilberman, R., Mulder, T., 2003. Predicting the terrestrial flux of sediment to the global ocean: a planetary perspective. Sedimentary Geology, 162(1–2), 524.Google Scholar
Syvitski, J. P. M., Vorosmarty, C. J., Kettner, A. J., Green, P., 2005. Impact of humans on the flux of terrestrial sediment to the global coastal ocean. Science, 308(5720), 376380.Google Scholar
Syvitski, J. P. M., Kettner, A. J., Overeem, I., et al., 2009. Sinking deltas due to human activities. Nature Geoscience, 2(10), 681686.Google Scholar
Syvitski, J. P., Kettner, A., Overeem, I., Brakenridge, G. R., Cohen, S., 2019. Latitudinal controls on siliciclastic sediment production and transport, Latitudinal Controls on Stratigraphic Models and Sedimentary Concepts, SEPM Special Publication 108, 1428.Google Scholar
Szupiany, R. N., Amsler, M.L., Parsons, D. R., Best, J. L., 2009. Morphology, flow structure, and suspended bed sediment transport at two large braid-bar confluences. Water Resources Research, 45. 10.1029/2008wr007428.Google Scholar
Szupiany, R. N., Amsler, M. L., Hernandez, J., et al., 2012. Flow fields, bed shear stresses, and suspended bed sediment dynamics in bifurcations of a large river. Water Resources Research, 48. 10.1029/2011wr011677.Google Scholar
Tabata, K. K., Hickin, E. J., 2003. Interchannel hydraulic geometry and hydraulic efficiency of the anastomosing Columbia River, southeastern British Columbia, Canada. Earth Surface Processes and Landforms, 28(8), 837852.Google Scholar
Tadaki, M., Brierley, G., Cullum, C., 2014. River classification: theory, practice, politics. Wiley Interdisciplinary Reviews – Water, 1(4), 349367.Google Scholar
Tal, M., Paola, C., 2007. Dynamic single-thread channels maintained by the interaction of flow and vegetation. Geology, 35(4), 347350.Google Scholar
Tal, M., Paola, C., 2010. Effects of vegetation on channel morphodynamics: results and insights from laboratory experiments. Earth Surface Processes and Landforms, 35(9), 10141028.Google Scholar
Tal, M., Gran, K., Murray, A. B., Paola, C., Hicks, D. M., 2004. Riparian vegetation as a primary control on channel characteristics in multi-thread rivers. In: Bennett, S. J., Simon, A. (eds.), Riparian Vegetation and Fluvial Geomorphology, 8, pp. 4358.Google Scholar
Talbot, T., Lapointe, M., 2002a. Numerical modeling of gravel bed river response to meander straightening: the coupling between the evolution of bed pavement and long profile. Water Resources Research, 38(6). 10.1029/2001wr000330.Google Scholar
Talbot, T., Lapointe, M., 2002b. Modes of response of a gravel bed river to meander straightening: the case of the Sainte-Marguerite River, Saguenay Region, Quebec, Canada. Water Resources Research, 38(6). 10.1029/2001wr000324.Google Scholar
Talling, P.J., Sowter, M.J., 1999. Drainage density on progressively tilted surfaces with different gradients, Wheeler Ridge, California. Earth Surface Processes and Landforms, 24(9), 809824.Google Scholar
Tang, H., Zhang, H., Yuan, S., 2018. Hydrodynamics and contaminant transport on a degraded bed at a 90-degree channel confluence. Environmental Fluid Mechanics, 18(2), 443463.Google Scholar
Taniguchi, K. T., Biggs, T. W., 2015. Regional impacts of urbanization on stream channel geometry: a case study in semiarid southern California. Geomorphology, 248, 228236.Google Scholar
Tanner, W. F., 1960. Helicoidal flow, a possible cause of meandering. Journal of Geophysical Research, 65, 993995.Google Scholar
Tanner, W. F., 1971. The river profile. Journal of Geology, 79(4), 482492.Google Scholar
Tarboton, D. G., Bras, R. L., Rodriguez-Iturbe, I., 1988. The fractal nature of river networks. Water Resources Research, 24(8), 13171322.Google Scholar
Tarboton, D. G., Bras, R. L., Rodriguez-Iturbe, I., 1990. On the fractal dimension of stream networks – comment. Water Resources Research, 26(9), 22432244.Google Scholar
Taylor, C. F. H., 1999. The role of overbank flow in governing the form of an anabranching river: the Fitzroy River, northwestern Australia. In: Smith, N. D., Rogers, J. (eds.), Fluvial Sedimentology VI, International Association of Sedimentologists Special Publication No. 28, Blackwell, Oxford, UK, pp. 7791.Google Scholar
Taylor, E. H., 1944. Flow characteristics at rectangular open-channel junctions. Transactions, American Society of Civil Engineers, 109, 893912.Google Scholar
Tennekes, H., Lumley, J. L., 1972. A First Course in Turbulence. MIT Press, Cambridge, MA.Google Scholar
Termini, D., 2009. Experimental observations of flow and bed processes in large-amplitude meandering flume. Journal of Hydraulic Engineering, 135(7), 575587.Google Scholar
Tessler, Z. D., Voeroesmarty, C. J., Grossberg, M., et al., 2015. Profiling risk and sustainability in coastal deltas of the world. Science, 349(6248), 638643.Google Scholar
Thayer, J. B., 2017. Downstream regime relations for single-thread channels. River Research and Applications, 33, 182186.Google Scholar
Thomas, R., Nicholas, A. P., 2002. Simulation of braided river flow using a new cellular routing scheme. Geomorphology, 43(3–4), 179195.Google Scholar
Thomas, R., Nicholas, A. P., Quine, T. A., 2007. Cellular modelling as a tool for interpreting historic braided river evolution. Geomorphology, 90(3–4), 302317.Google Scholar
Thomas, R. B., Megahan, W. F., 1998. Peak flow responses to clear-cutting and roads in small and large basins, western Cascades, Oregon: a second opinion. Water Resources Research, 34(12), 33933403.Google Scholar
Thomas, R. E., Parsons, D. R., Sandbach, S. D., et al., 2011. An experimental study of discharge partitioning and flow structure at symmetrical bifurcations. Earth Surface Processes and Landforms, 36(15), 20692082.Google Scholar
Thompson, A., 1986. Secondary flows and the pool-riffle unit – a case study of the processes of meander development. Earth Surface Processes and Landforms, 11(6), 631641.Google Scholar
Thompson, C., Croke, J., 2008. Channel flow competence and sediment transport in upland streams in southeast Australia. Earth Surface Processes and Landforms, 33(3), 329352.Google Scholar
Thompson, C., Croke, J., 2013. Geomorphic effects, flood power, and channel competence of a catastrophic flood in confined and unconfined reaches of the upper Lockyer valley, southeast Queensland, Australia. Geomorphology, 197, 156169.Google Scholar
Thompson, D. M., 2001. Random controls on semi-rhythmic spacing of pools and riffles in constriction-dominated rivers. Earth Surface Processes and Landforms, 26(11), 11951212.Google Scholar
Thompson, D. M., 2004. The influence of pool length on local turbulence production and energy slope: a flume experiment. Earth Surface Processes and Landforms, 29(11), 13411358.Google Scholar
Thompson, D. M., 2006. The role of vortex shedding in the scour of pools. Advances in Water Resources, 29(2), 121129.Google Scholar
Thompson, D. M., 2007. The characteristics of turbulence in a shear zone downstream of a channel constriction in a coarse-grained forced pool. Geomorphology, 83(3–4), 199214.Google Scholar
Thompson, D. M., Fixler, S. A., 2017. Formation and maintenance of a forced pool-riffle couplet following loading of large wood. Geomorphology, 296, 7490.Google Scholar
Thompson, D. M., Hoffman, K. S., 2001. Equilibrium pool dimensions and sediment-sorting patterns in coarse-grained, New England channels. Geomorphology, 38(3–4), 301316.Google Scholar
Thompson, D. M., McCarrick, C. R., 2010. A flume experiment on the effect of constriction shape on the formation of forced pools. Hydrology and Earth System Sciences, 14(7), 13211330.Google Scholar
Thompson, D. M., Wohl, E. E., 2009. The linkage between velocity patterns and sediment entrainment in a forced-pool and riffle unit. Earth Surface Processes and Landforms, 34(2), 177192.Google Scholar
Thompson, D. M., Wohl, E. E., Jarrett, R. D., 1996. Revised velocity-reversal and sediment-sorting model for a high-gradient, pool-riffle stream. Physical Geography, 17(2), 142156.Google Scholar
Thompson, D. M., Wohl, E. E., Jarrett, R. D., 1999. Velocity reversals and sediment sorting in pools and riffles controlled by channel constrictions. Geomorphology, 27(3–4), 229241.Google Scholar
Thoms, M. C., 1987. Channel sedimentation within the urbanized River Tame, UK. Regulated Rivers Research and Management, 1(3), 229246.Google Scholar
Thomson, J., 1876. On the origin of windings of rivers in alluvial plains. Proceedings of the Royal Society, 25, 58.Google Scholar
Thonemann, P., 2011. The Maeander Valley: a historical geography from antiquity to Byzantium. Cambridge University Press, Cambridge, UK.Google Scholar
Thorn, C. E., Welford, M. R., 1994. The equilibrium concept in geomorphology. Annals of the Association of American Geographers, 84(4), 666696.Google Scholar
Thorne, C. R., 1982. Processes and mechanisms of river bank erosion. In: Hey, R. D., Bathurst, J. C., Thorne, C. R. (eds.), Gravel-Bed Rivers. Wiley, Chichester, UK, pp. 227259.Google Scholar
Thorne, C. R., 1990. Effects of vegetation on river bank erosion and stability. In: Thornes, J. B. (ed.), Vegetation and Erosion: Processes and Environments. Wiley, Chichester, UK, pp. 125144.Google Scholar
Thorne, C. R., Abt, S. R., 1993. Analysis of riverbank stability due to toe scour and lateral erosion. Earth Surface Processes and Landforms, 18(9), 835843.Google Scholar
Thorne, C. R., Tovey, N. K., 1981. Stability of composite river banks. Earth Surface Processes and Landforms, 6(5), 469484.Google Scholar
Thorne, C. R., Zevenbergen, L. W., Pitlick, J. C., et al., 1985. Direct measurement of secondary currents in a meandering sand-bed river. Nature, 315(6022), 746747.Google Scholar
Thorne, S. D., Furbish, D. J., 1995. Influences of coarse bank roughness on flow within a sharply curved river bend. Geomorphology, 12(3), 241257.Google Scholar
Thorp, J. H., Thoms, M. C., Delong, M. D., 2006. The riverine ecosystem synthesis: biocomplexity in river networks across space and time. River Research and Applications, 22(2), 123147.Google Scholar
Tinkler, K. J., 1985. A Short History of Geomorphology. Barnes and Noble, Totowa, NJ.Google Scholar
Tockner, K., Paetzold, A., Karaus, U., Claret, C., Zettel, J., 2006. Ecology of braided rivers. In: Sambrook Smith, G. H., Best, J. L., Bristow, C. S., Petts, G. E. (eds.), Braided Rivers: Process, Deposits, Ecology and Management, International Association of Sedimentologists Special Publication 36. Blackwell, Oxford, UK, pp. 339359.Google Scholar
Toniolo, H., Parker, G., Voller, V., 2007. Role of ponded turbidity currents in reservoir trap efficiency. Journal of Hydraulic Engineering, 133(6), 579595.Google Scholar
Toonen, W. H. J., Kleinhans, M. G., Cohen, K. M., 2012. Sedimentary architecture of abandoned channel fills. Earth Surface Processes and Landforms, 37(4), 459472.Google Scholar
Tooth, S., 2005. Splay formation along the lower reaches of ephemeral rivers on the Northern Plains of arid central Australia. Journal of Sedimentary Research, 75(4), 636649.Google Scholar
Tooth, S., McCarthy, T. S., 2004a. Anabranching in mixed bedrock-alluvial rivers: the example of the Orange River above Augrabies Falls, Northern Cape Province, South Africa. Geomorphology, 57(3–4), 235262.Google Scholar
Tooth, S., McCarthy, T. S., 2004b. Controls on the transition from meandering to straight channels in the wetlands of the Okavango Delta, Botswana. Earth Surface Processes and Landforms, 29(13), 16271649.Google Scholar
Tooth, S., Nanson, G. C., 1999. Anabranching rivers on the Northern Plains of arid central Australia. Geomorphology, 29(3–4), 211233.Google Scholar
Tooth, S., Nanson, G. C., 2000a. Equilibrium and nonequilibrium conditions in dryland rivers. Physical Geography, 21(3), 183211.Google Scholar
Tooth, S., Nanson, G. C., 2000b. The role of vegetation in the formation of anabranching channels in an ephemeral river, Northern plains, arid central Australia. Hydrological Processes, 14(16–17), 30993117.Google Scholar
Tooth, S., Nanson, G. C., 2004. Forms and processes of two highly contrasting rivers in arid central Australia, and the implications for channel-pattern discrimination and prediction. Geological Society of America Bulletin, 116(7–8), 802816.Google Scholar
Tooth, S., Jansen, J. D., Nanson, G. C., Coulthard, T. J., Pietsch, T., 2008. Riparian vegetation and the late Holocene development of an anabranching river: Magela Creek, northern Australia. Geological Society of America Bulletin, 120(7–8), 10211035.Google Scholar
Topping, D. J., Rubin, D. M., Vierra, L. E., 2000. Colorado River sediment transport – 1. Natural sediment supply limitation and the influence of Glen Canyon Dam. Water Resources Research, 36(2), 515542.Google Scholar
Torizzo, M., Pitlick, J., 2004. Magnitude-frequency of bed load transport in mountain streams in Colorado. Journal of Hydrology, 290(1–2), 137151.Google Scholar
Törnqvist, T. E., Bridge, J. S., 2002. Spatial variation of overbank aggradation rate and its influence on avulsion frequency. Sedimentology, 49(5), 891905.Google Scholar
Törnqvist, T. E., van Dijk, G. J., 1993. Optimizing sampling strategy for radiocarbon dating of Holocene fluvial systems in a vertically aggrading setting. Boreas, 22(2), 129145.Google Scholar
Torres, A., Brandt, J., Lear, K., Liu, J. G., 2017. A looming tragedy of the sand commons. Science, 357(6355), 970971.Google Scholar
Trampush, S. M., Huzurbazar, S., McElroy, B., 2014. Empirical assessment of theory for bankfull characteristics of rivers. Water Resources Research, 50, 92119220.Google Scholar
Tranmer, A. W., Tonina, D., Benjankar, R., Tiedemann, M., Goodwin, P., 2015. Floodplain persistence and dynamic-equilibrium conditions in a canyon environment. Geomorphology, 250, 147158.Google Scholar
Trimble, S. W., 1974. Man-Induced Soil Erosion on the Southern Piedmont, 1700–1970. Soil Conservation Society of America, Ankeny, IA.Google Scholar
Trimble, S. W., 1975. Denudation studies – can we assume stream steady state? Science, 188(4194), 12071208.Google Scholar
Trimble, S. W., 1977. Fallacy of stream equilibrium in contemporary denudation studies. American Journal of Science, 277(7), 876887.Google Scholar
Trimble, S. W., 1981. Changes in sediment storage in the Coon Creek Basin, Driftless Area, Wisconsin, 1853–1975. Science, 214(4517), 181183.Google Scholar
Trimble, S. W., 1983. A sediment budget for Coon Creek Basin in the Driftless Area, Wisconsin, 1853–1977. American Journal of Science, 283(5), 454474.Google Scholar
Trimble, S. W., 1993. The distributed sediment budget model and watershed management in the Paleozoic plateau of the upper midwestern United States. Physical Geography, 14(3), 285303.Google Scholar
Trimble, S. W., 1994. Erosional effects of cattle on streambanks in Tennessee, USA. Earth Surface Processes and Landforms, 19(5), 451464.Google Scholar
Trimble, S. W., 1995. Catchment sediment budgets and change. In: Gurnell, A., Petts, G. (eds.), Changing River Channels. Wiley and Sons, Chichester, UK, pp. 201215.Google Scholar
Trimble, S. W., 1997. Contribution of stream channel erosion to sediment yield from an urbanizing watershed. Science, 278(5342), 14421444.Google Scholar
Trimble, S. W., 1999. Decreased rates of alluvial sediment storage in the Coon Creek Basin, Wisconsin, 1975–93. Science, 285(5431), 12441246.Google Scholar
Trimble, S. W., 2008. The use of historical data and artifacts in geomorphology. Progress in Physical Geography, 32(1), 329.Google Scholar
Trimble, S. W., 2009. Fluvial processes, morphology and sediment budgets in the Coon Creek Basin, WI, USA, 1975–1993. Geomorphology, 108(1–2), 823.Google Scholar
Trimble, S. W., 2013. Historical Agriculture and Soil Erosion in the Upper Mississippi River Hill Country. CRC Press, Boca Raton, FL.Google Scholar
Trimble, S. W., Cooke, R. U., 1991. Historical sources for geomorphological research in the United States. Professional Geographer, 43(2), 212228.Google Scholar
Trimble, S. W., Crosson, P., 2000. U.S. soil erosion rates – myth and reality. Science, 289(5477), 248250.Google Scholar
Trimble, S. W., Mendel, A. C., 1995. The cow as a geomorphic agent – a critical review. Geomorphology, 13(1–4), 233253.Google Scholar
Tubino, M., 1991. Growth of alternate bars in unsteady flow. Water Resources Research, 27(1), 3752.Google Scholar
Tubino, M., Seminara, G., 1990. Free forced interactions in developing meanders and suppression of free bars. Journal of Fluid Mechanics, 214, 131159.Google Scholar
Tucker, G. E., 2004. Drainage basin sensitivity to tectonic and climatic forcing: implications of a stochastic model for the role of entrainment and erosion thresholds. Earth Surface Processes and Landforms, 29(2), 185205.Google Scholar
Tucker, G. E., Hancock, G. R., 2010. Modelling landscape evolution. Earth Surface Processes and Landforms, 35(1), 2850.Google Scholar
Tucker, G. E., Slingerland, R. L., 1994. Erosional dynamics, flexural isostasy, and long-lived escarpments – a numerical modeling study. Journal of Geophysical Research-Solid Earth, 99(B6), 1222912243. 10.1029/94jb00320.Google Scholar
Tucker, G. E., Lancaster, S., Gasparini, N., Bras, R. L., 2001. The Channel-Hillslope Integrated Landscape Development Model (CHILD) In: Harmon, R. S., Doe, W. W. III (eds.), Landscape Erosion and Evolution Modeling. Kluwer Academic, New York, pp. 349388.Google Scholar
Turnipseed, D. P., Sauer, V. B., 2010. Discharge measurements at gaging stations. U.S. Geological Survey Techniques and Methods, Book 3, Chapter A8. http://pubs.usgs.gov/tm/tm3-a8/.Google Scholar
Turowski, J. M., Lague, D., Hovius, N., 2007. Cover effect in bedrock abrasion: a new derivation and its implications for the modeling of bedrock channel morphology. Journal of Geophysical Research-Earth Surface, 112(F4). 10.1029/2006jf000697.Google Scholar
Turowski, J. M., Hovius, N., Wilson, A., Horng, M.-J., 2008. Hydraulic geometry, river sediment and the definition of bedrock channels. Geomorphology, 99(1–4), 2638.Google Scholar
Turowski, J. M., Yager, E. M., Badoux, A., Rickenmann, D., Molnar, P., 2009. The impact of exceptional events on erosion, bedload transport and channel stability in a step-pool channel. Earth Surface Processes and Landforms, 34(12), 16611673.Google Scholar
Tyner, J. S., Yoder, D. C., Chomicki, B. J., Tyagi, A., 2011. A review of construction site best management practices for erosion control. Transactions of the American Society of Agricultural and Biological Engineers, 54(2), 441450.Google Scholar
Uijttewaal, W. S. J., Booij, R., 2000. Effects of shallowness on the development of free-surface mixing layers. Physics of Fluids, 12(2), 392402.Google Scholar
Uijttewaal, W. S. J., Tukker, J., 1998. Development of quasi two-dimensional structures in a shallow free-surface mixing layer. Experiments in Fluids, 24(3), 192200.Google Scholar
Umar, M., Rhoads, B. L., Greenberg, J. A., 2018. Use of multispectral satellite remote sensing to assess mixing of suspended sediment downstream of large river confluences. Journal of Hydrology, 556, 325338.Google Scholar
United Nations, 2018. World Urbanization Prospects: The 2018 Revision [key facts], United Nations, New York.Google Scholar
Urban, M. A., Rhoads, B. L., 2003. Catastrophic human-induced change in stream-channel planform and geometry in an agricultural watershed, Illinois, USA. Annals of the Association of American Geographers, 93(4), 783796.Google Scholar
U.S. Army Corps of Engineers, 2010. Illinois Stream Mitigation Guidance. U.S. Army Corps of Engineers. www.mvr.usace.army.mil/Portals/48/docs/regulatory/mitigation/IllinoisMethod.pdf.Google Scholar
U.S. Army Corps of Engineers, 2016. HEC-RAS River Analysis System, Hydraulic Reference Manual, Version 5.0. U.S. Army Corps of Engineers Hydrologic Engineering Center, Davis, CA.Google Scholar
U.S. Army Corps of Engineers, 2018. National Inventory of Dams. http://nid.usace.army.mil/ (accessed November 27, 2019).Google Scholar
U.S. Department of Agriculture, 1983. Sediment sources, yields, and delivery ratios, Chapter 6 of Section 3: Sedimentation, National Engineering Handbook. U.S. Department of Agriculture, Washington, D.C.Google Scholar
U.S. Department of Agriculture, 1995. U.S.D.A – Water Erosion Prediction Project – Technical Documentation. NSERL Report No. 10, National Soil Erosion Research Laboratory, West Lafayette, IN.Google Scholar
U.S. Department of Agriculture, 2013. Revised Universal Soil Loss Equation Version 2 (RUSLE2) Science Documentation A.R.S. U.S. Department of Agriculture, Washington, DC.Google Scholar
U.S. Department of Agriculture, 2018. Summary Report: 2015 National Resources Inventory. Natural Resources Conservation Service, Washington, DC, Center for Survey Statistics and Methodology, Iowa State University, Ames, IA.Google Scholar
U.S. Environmental Protection Agency, 2008. Handbook for Developing Watershed Plans to Restore and Protect Waters. U.S. Environmental Protection Agency, Washington, DC.Google Scholar
Vallé, B. L., Pasternack, G. B., 2006. Submerged and unsubmerged natural hydraulic jumps in a bedrock step-pool mountain channel. Geomorphology, 82(1–2), 146159.Google Scholar
Valyrakis, M., Diplas, P., Dancey, C. L., Greer, K., Celik, A. O., 2010. Role of instantaneous force magnitude and duration on particle entrainment. Journal of Geophysical Research – Earth Surface, 115. 10.1029/2008jf001247.Google Scholar
Valyrakis, M., Diplas, P., Dancey, C. L., 2011. Entrainment of coarse grains in turbulent flows: an extreme value theory approach. Water Resources Research, 47. 10.1029/2010wr010236.Google Scholar
Valyrakis, M., Diplas, P., Dancey, C. L., 2013. Entrainment of coarse particles in turbulent flows: an energy approach. Journal of Geophysical Research – Earth Surface, 118(1), 4253. 10.1029/2012jf002354.Google Scholar
van de Lageweg, W. I., van Dijk, W. M., Baar, A. W., Rutten, J., Kleinhans, M. G., 2014. Bank pull or bar push: what drives scroll-bar formation in meandering rivers? Geology, 42(4), 319322.Google Scholar
van de Wiel, M. J., Darby, S. E., 2007. A new model to analyse the impact of woody riparian vegetation on the geotechnical stability of riverbanks. Earth Surface Processes and Landforms, 32(14), 21852198.Google Scholar
van de Wiel, M. J., Rousseau, Y. Y., Darby, S. E., 2016. Models in fluvial geomorphology. In: Kondolf, G. M., Piegay, H. (eds.), Tools in Fluvial Geomorphology, 2nd Edition. Wiley, Chichester, UK, pp. 383411.Google Scholar
van den Berg, J. H., 1987. Bedform migration and bed-load transport in some rivers and tidal environments. Sedimentology, 34(4), 681698.Google Scholar
van den Berg, J. H., 1995. Prediction of alluvial channel pattern of perennial rivers. Geomorphology, 12(4), 259279.Google Scholar
van den Berg, J. H., Bledsoe, B. P., 2003. Comment on Lewin and Brewer (2001): “Predicting channel patterns”. Geomorphology, 53(3–4), 333337.Google Scholar
van der Mark, C. F., Mosselman, E., 2013. Effects of helical flow in one-dimensional modelling of sediment distribution at river bifurcations. Earth Surface Processes and Landforms, 38(5), 502511.Google Scholar
van Dijk, W. M., van de Lageweg, W. I., Kleinhans, M. G., 2012. Experimental meandering river with chute cutoffs. Journal of Geophysical Research-Earth Surface, 117. 10.1029/2011jf002314.Google Scholar
van Dijk, W. M., Schuurman, F., van de Lageweg, W. I., Kleinhans, M. G., 2014. Bifurcation instability and chute cutoff development in meandering gravel-bed rivers. Geomorphology, 213, 277291.Google Scholar
van Maren, D. S., Winterwerp, J. C., Wu, B. S., Zhou, J. J., 2009. Modelling hyperconcentrated flow in the Yellow River. Earth Surface Processes and Landforms, 34(4), 596612.Google Scholar
van Oorschot, M., Kleinhans, M., Geerling, G., Middelkoop, H., 2016. Distinct patterns of interaction between vegetation and morphodynamics. Earth Surface Processes and Landforms, 41(6), 791808.Google Scholar
van Rijn, L. C., 1984. Sediment transport, part II: suspended load transport. Journal of Hydraulic Engineering, 110(11), 16131641.Google Scholar
van Toorenenburg, K. A., Donselaar, M. E., Weltje, G. J., 2018. The life cycle of crevasse splays as a key mechanism in the aggradation of alluvial ridges and river avulsion. Earth Surface Processes and Landforms, 43(11), 24092420.Google Scholar
van Vliet, J., Eitelberg, D. A., Verburg, P. H., 2017. A global analysis of land take in cropland areas and production displacement from urbanization. Global Environmental Change – Human and Policy Dimensions, 43, 107115.Google Scholar
Vanapalli, S. K., Fredlund, D. G., Pufahl, D. E., Clifton, A. W., 1996. Model for the prediction of shear strength with respect to soil suction. Canadian Geotechnical Journal, 33(3), 379392.Google Scholar
Vandaele, K., Poesen, J., Govers, G., van Wesemael, B., 1996. Geomorphic threshold conditions for ephemeral gully incision. Geomorphology, 16(2), 161173.Google Scholar
Vanmaercke, M., Kettner, A. J., van den Ekhaut, M., et al., 2014. Moderate seismic activity affects contemporary sediment yields. Progress in Physical Geography, 38, 145172.Google Scholar
Vanmaercke, M., Ardizzone, F., Rossi, M., Guzzetti, F., 2017. Exploring the effects of seismicity on landslides and catchment sediment yield: an Italian case study. Geomorphology, 278, 171183.Google Scholar
Vanoni, V. A., Benedict, P. C., Bondurant, D. C., et al., 1966. Sediment transportation mechanics: initiation of motion. Journal of the Hydraulics Division – ASCE, 92, 291314.Google Scholar
Vasconselos, P. M., Farley, K. A., Stone, J., Piacentini, T., Fifield, L. K., 2019. Stranded landscapes in the humid tropics: Earth’s oldest land surfaces. Earth and Planetary Science Letters, 519, 152164.Google Scholar
Vaughan, I. P., Diamond, M., Gurnell, A. M., et al., 2009. Integrating ecology with hydromorphology: a priority for river science and management. Aquatic Conservation – Marine and Freshwater Ecosystems, 19(1), 113125.Google Scholar
Vaughn, D. M., 1990. Flood dynamics of a concrete‐lined, urban stream in Kansas City, Missouri. Earth Surface Processes and Landforms, 15(6), 525537.Google Scholar
Venditti, J. G., 2013. Bedforms in sand-bedded rivers. In: Schroder, J. L. (ed.), Treatise on Geomorphology, Vol. 9, Fluvial Geomorphology, Wohl, E. (vol. ed.), Academic Press, San Diego, CA, pp. 137162.Google Scholar
Venditti, J. G., Church, M., 2014. Morphology and controls on the position of a gravel-sand transition: Fraser River, British Columbia. Journal of Geophysical Research – Earth Surface, 119(9), 19591976.Google Scholar
Venditti, J. G., Nelson, P. A., Minear, J. T., Wooster, J., Dietrich, W. E., 2012. Alternate bar response to sediment supply termination. Journal of Geophysical Research – Earth Surface, 117. 10.1029/2011jf002254.Google Scholar
Venditti, J. G., Hardy, R. J., Church, M., Best, J. L., 2013. What is a coherent flow structure in geophysical flow? In: Venditti, J. G. (ed.), Coherent Flow Structures at Earth’s Surface. Wiley and Sons, Chichester, UK, pp. 116.Google Scholar
Venditti, J. G., Domarad, N., Church, M., Rennie, C. D., 2015. The gravel-sand transition: sediment dynamics in a diffuse extension. Journal of Geophysical Research – Earth Surface, 120(6), 943963. 10.1002/2014jf003328.Google Scholar
Vercruysse, K., Grabowski, R. C., Rickson, R. J., 2017. Suspended sediment transport dynamics in rivers: multi-scale drivers of temporal variation. Earth-Science Reviews, 166, 3852.Google Scholar
Vericat, D., Wheaton, J. M., Brasington, J., 2017. Revisiting the morphological approach: opportunities and challenges with repeat high resolution topography. In: Tsutsumi, D., Laronne, J. B. (eds.), Gravel-Bed Rivers: Processes and Disasters. Wiley, Chichester, UK, pp. 121158.Google Scholar
Vermeulen, B., Hoitink, A. J. F., Labeur, R. J., 2015. Flow structure caused by a local cross-sectional area increase and curvature in a sharp river bend. Journal of Geophysical Research – Earth Surface, 120(9), 17711783. 10.1002/2014jf003334.Google Scholar
Verpoorter, C., Kutser, T., Seekell, D. A., Tranvik, L. J., 2014. A global inventory of lakes based on high-resolution satellite imagery. Geophysical Research Letters, 41(18), 63966402. 10.1002/2014gl060641.Google Scholar
Verstraeten, G., Broothaerts, N., Van Loo, M., et al., 2017. Variability in fluvial geomorphic response to anthropogenic disturbance. Geomorphology, 294, 2039.Google Scholar
Vetter, T., 2011a. Riffle-pool morphometry and stage-dependent morphodynamics of a large floodplain river (Vereinigte Mulde, Sachsen-Anhalt, Germany). Earth Surface Processes and Landforms, 36(12), 16471657.Google Scholar
Vetter, T., 2011b. Analysing riffle-pool dynamics of a large floodplain river with a system-oriented approach. Zeitschrift fur Geomorphologie, 55, 355372.Google Scholar
Viero, D. P., Lopez Dupon, S., Lanzoni, S., 2018. Chute cutoffs in meandering rivers: formative mechanisms and hydrodynamic forcing. In: Ghinassi, M., Columbera, L., Mountney, N. P., Reesink, A. J. H., Bateman, M. (eds.), Fluvial Meanders and their Sedimentary Products in the Rock Record. International Association of Sedimentologists Special Publication 48. Wiley and Sons, Hoboken, NJ, pp. 201229.Google Scholar
Vietz, G. J., Rutherfurd, I. D., Stewardson, M. J., Finlayson, B. L., 2012. Hydrodynamics and sedimentology of concave benches in a lowland river. Geomorphology, 147, 86101.Google Scholar
Vietz, G. J., Walsh, C. J., Fletcher, T. D., 2016. Urban hydrogeomorphology and the urban stream syndrome: treating the symptoms and causes of geomorphic change. Progress in Physical Geography, 40(3), 480492.Google Scholar
Vigilar, G. G., Diplas, P., 1997. Stable channels with mobile bed: formulation and numerical solution. Journal of Hydraulic Engineering, 123(3), 189199.Google Scholar
Vigilar, G. G., Diplas, P., 1998. Stable channels with mobile bed: model verification and graphical solution. Journal of Hydraulic Engineering, 124(11), 10971108.Google Scholar
Vogel, R. M., Stedinger, J. R., Hooper, R. P., 2003. Discharge indices for water quality loads. Water Resources Research, 39(10). 10.1029/2002wr001872.Google Scholar
Voichick, N., Topping, D. J., 2014. Extending the turbidity record—making additional use of continuous data from turbidity, acoustic-doppler, and laser diffraction instruments and suspended-sediment samples in the Colorado River in Grand Canyon. U.S. Geological Survey Scientific Investigations Report 2014–5097. Available online only at http://dx.doi.org/10.3133/sir20145097.Google Scholar
Vollmer, S., Kleinhans, M. G., 2007. Predicting incipient motion, including the effect of turbulent pressure fluctuations in the bed. Water Resources Research, 43(5). 10.1029/2006wr004919.Google Scholar
von Blanckenburg, F., 2005. The control mechanisms of erosion and weathering at basin scale from cosmogenic nuclides in river sediment. Earth and Planetary Science Letters, 237(3–4), 462479.Google Scholar
Vorosmarty, C. J., Meybeck, M., Fekete, B., et al., 2003. Anthropogenic sediment retention: major global impact from registered river impoundments. Global and Planetary Change, 39(1–2), 169190.Google Scholar
Wade, R. J., Rhoads, B. L., Rodriguez, J., et al., 2002. Integrating science and technology to support stream naturalization near Chicago, Illinois. Journal of the American Water Resources Association, 38(4), 931944.Google Scholar
Walling, D. E., 1977. Assessing accuracy of suspended sediment rating curves for a small basin. Water Resources Research, 13(3), 530538.Google Scholar
Walling, D. E., 1983. The sediment delivery problem. Journal of Hydrology, 65(1–3), 209237.Google Scholar
Walling, D. E., 1999. Linking land use, erosion and sediment yields in river basins. Hydrobiologia, 410, 223240.Google Scholar
Walling, D. E., 2005. Tracing suspended sediment sources in catchments and river systems. Science of the Total Environment, 344(1–3), 159184.Google Scholar
Walling, D. E., 2006. Human impact on land-ocean sediment transfer by the world’s rivers. Geomorphology, 79(3–4), 192216.Google Scholar
Walling, D. E., Collins, A. L., 2008. The catchment sediment budget as a management tool. Environmental Science & Policy, 11(2), 136143.Google Scholar
Walling, D. E., Fang, D., 2003. Recent trends in the suspended sediment loads of the world’s rivers. Global and Planetary Change, 39(1–2), 111126.Google Scholar
Walling, D. E., Gregory, K. J., 1970. The measurement of the effects of building construction on drainage basin dynamics. Journal of Hydrology, 11, 129144.Google Scholar
Walling, D. E., He, Q., 1997. Investigating spatial patterns of overbank sedimentation on river floodplains. Water Air and Soil Pollution, 99(1–4), 920.Google Scholar
Walling, D. E., He, Q., 1998. The spatial variability of overbank sedimentation on river floodplains. Geomorphology, 24(2–3), 209223.Google Scholar
Walling, D. E., He, Q., 1999. Using fallout lead-210 measurements to estimate soil erosion on cultivated land. Soil Science Society of America Journal, 63(5), 14041412.Google Scholar
Walling, D. E., Kleo, A. H. A., 1979. Sediment yields in areas of low precipitation: an overview. In: The Hydrology of Areas of Low Precipitation. IAHS Publication No. 128, IAHS Press, Wallingford, UK, pp. 479493.Google Scholar
Walling, D. E., Webb, B. W., 1983. Patterns of sediment yield. In: Gregory, K. J. (ed.), Background to Palaeohydrology. Wiley and Sons, New York, pp. 69100.Google Scholar
Walling, D. E., Webb, B. W., 1996. Erosion and sediment yield: a global overview. In: Walling, D. E., Webb, B. W. (eds.), Erosion and Sediment Yield: Global and Regional Perspectives. IAHS Publication No. 236, IAHS Press, Wallingford, UK, pp. 319.Google Scholar
Walling, D. E., Russell, M. A., Hodgkinson, R. A., Zhang, Y., 2002. Establishing sediment budgets for two small lowland agricultural catchments in the UK. Catena, 47(4), 323353.Google Scholar
Walling, D. E., Collins, A. L., Jones, P. A., Leeks, G. J. L., Old, G., 2006. Establishing fine-grained sediment budgets for the Pang and Lambourn LOCAR catchments, UK. Journal of Hydrology, 330(1–2), 126141.Google Scholar
Wallinga, J., 2002. Optically stimulated luminescence dating of fluvial deposits: a review. Boreas, 31(4), 303322.Google Scholar
Walsh, J., Hicks, D. M., 2002. Braided channels: self-similar or self-affine? Water Resources Research, 38(6). 10.1029/2001wr000749.Google Scholar
Walter, R. C., Merritts, D. J., 2008. Natural streams and the legacy of water-powered mills. Science, 319(5861), 299304.Google Scholar
Wang, Y., Rhoads, B. L., Wang, D., 2016. Assessment of the flow regime alterations in the middle reach of the Yangtze River associated with dam construction: potential ecological implications. Hydrological Processes, 30(21), 39493966.Google Scholar
Wang, Y., Rhoads, B. L., Wang, D., Wu, J., Zhang, X., 2018. Impacts of large dams on the complexity of suspended sediment dynamics in the Yangtze River. Journal of Hydrology, 558, 184195.Google Scholar
Warner, R. F., 1997. Floodplain stripping: another form of adjustment to secular hydrologic regime change in Southeast Australia. Catena, 30(4), 263282.Google Scholar
Warrick, J. A., Milliman, J. D., Walling, D. E., et al., 2014. Earth is (mostly) flat: apportionment of the flux of continental sediment over millennial time scales: comment. Geology, 42(1), e316. 10.1130/g34846c.1.Google Scholar
Wathen, S. J., Ferguson, R. I., Hoey, T. B., Werritty, A., 1995. Unequal mobility of gravel and sand in weakly bimodal river sediments. Water Resources Research, 31(8), 20872096.Google Scholar
Watson, R. L., 1969. Modified Rubey’s law accurately predicts sediment settling velocities. Water Resources Research, 5(5), 11471150.Google Scholar
Waylen, P., Woo, M. K., 1982. Prediction of annual floods generated by mixed processes. Water Resources Research, 18(4), 12831286.Google Scholar
Waythomas, C., Williams, G. P., 1988. Sediment yield and spurious correlation – toward a better portrayal of the annual suspended-sediment load of rivers. Geomorphology, 1, 309316.Google Scholar
Webber, N. B., Greated, C. A., 1966. An investigation of flow behavior at the junction of rectangular channels. Proceedings of the Institute of Civil Engineers, 34, 321334.Google Scholar
Weber, L. J., Schumate, E. D., Mawer, N., 2001. Experiments on flow at a 90° open-channel junction. Journal of Hydraulic Engineering, 127(5), 340350.Google Scholar
Weichert, R. B., Bezzola, G. R., Minor, H.-E., 2008. Bed morphology and generation of step-pool channels. Earth Surface Processes and Landforms, 33(11), 16781692.Google Scholar
Welber, M., Bertoldi, W., Tubino, M., 2012. The response of braided planform configuration to flow variations, bed reworking and vegetation: the case of the Tagliamento River, Italy. Earth Surface Processes and Landforms, 37(5), 572582.Google Scholar
Welford, M. R., 1993. Field evaluation of empirical equations in straight alluvial channels. Physical Geography, 14(6), 581598.Google Scholar
Welford, M. R., 1994. A field test of Tubino’s (1991) model of alternate bar formation. Earth Surface Processes and Landforms, 19(4), 287297.Google Scholar
Wellmeyer, J. L., Slattery, M. C., Phillips, J. D., 2005. Quantifying downstream impacts of impoundment on flow regime and channel planform, lower Trinity River, Texas. Geomorphology, 69(1–4), 113.Google Scholar
Wende, R., Nanson, G. C., 1998. Anabranching rivers: ridge-form alluvial channels in tropical northern Australia. Geomorphology, 22(3–4), 205224.Google Scholar
Wentworth, C. K., 1919. A laboratory and field study of cobble abrasion. Journal of Geology, 27(7), 507521.Google Scholar
Wentworth, C. K., 1922. A scale of grade and class terms for clastic sediments. Journal of Geology, 30, 377392.Google Scholar
Werritty, A., 1992. Downstream fining in a gravel-bed river in southern Poland – lithologic controls and the role of abrasion. In: Billi, P., Hey, R.D., Thorne, C.R., Tacconi, P. (eds.), Dynamics of Gravel-Bed Rivers. Wiley, Chichester, UK, pp. 333350.Google Scholar
Wheaton, J., Darby, S. E., Sear, D. A., 2008. The scope of uncertainties in river restoration. In: Darby, S. E., Sear, D. A. (eds.), River Restoration: Managing the Uncertainty of Restoring Physical Habitat. Wiley, Chichester, UK, pp. 2139.Google Scholar
Wheaton, J. M., Brasington, J., Darby, S. E., Sear, D. A., 2010. Accounting for uncertainty in DEMs from repeat topographic surveys: improved sediment budgets. Earth Surface Processes and Landforms, 35(2), 136156.Google Scholar
Wheaton, J. M., Brasington, J., Darby, S. E., et al., 2013. Morphodynamic signatures of braiding mechanisms as expressed through change in sediment storage in a gravel-bed river. Journal of Geophysical Research – Earth Surface, 118(2), 759779.Google Scholar
Wheaton, J. M., Fryirs, K. A., Brierley, G., et al., 2015. Geomorphic mapping and taxonomy of fluvial landforms. Geomorphology, 248, 273295.Google Scholar
Whipple, K. X., 2001. Fluvial landscape response time: how plausible is steady-state denudation? American Journal of Science, 301(4–5), 313325.Google Scholar
Whipple, K. X., 2004. Bedrock rivers and the geomorphology of active orogens. Annual Review of Earth and Planetary Sciences, 32, 151185.Google Scholar
Whipple, K. X., Tucker, G. E., 1999. Dynamics of the stream-power river incision model: implications for height limits of mountain ranges, landscape response timescales, and research needs. Journal of Geophysical Research-Solid Earth, 104(B8), 1766117674.Google Scholar
Whipple, K. X., DiBiase, R. A., Crosby, B. T., 2013. Bedrock rivers. In: Schroder, J. (ed.), Treatise on Geomorphology, Vol. 9, Fluvial Geomorphology, Wohl, E. E. (vol. ed.). Academic Press, San Diego, CA, pp. 550573.Google Scholar
Whipple, W., Jr., DiLouie, J., 1981. Coping with increased stream erosion in urbanizing areas. Water Resources Research, 17(5), 15611564.Google Scholar
Whitaker, A. C., Potts, D. F., 2007. Analysis of flow competence in an alluvial gravel bed stream, Dupuyer Creek, Montana. Water Resources Research, 43(7). 10.1029/2006wr005289.Google Scholar
White, W. R., Bettess, R., Paris, E., 1982. Analytical approach to river regime. Journal of the Hydraulics Division – ASCE, 108(10), 11791193.Google Scholar
Whitehead, A. N., 1925. Science and the Modern World. The Free Press, New York.Google Scholar
Whiting, P. J., Dietrich, W. E., 1990. Boundary shear stress and roughness over mobile alluvial beds. Journal of Hydraulic Engineering, 116(12), 14951511.Google Scholar
Whiting, P. J., Dietrich, W. E., 1991. Convective accelerations and boundary shear stress over a channel bar. Water Resources Research, 27(5), 783796.Google Scholar
Whiting, P. J., Dietrich, W. E., 1993a. Experimental studies of bed topography and flow patterns in large-amplitude meanders. 1. Observations. Water Resources Research, 29(11), 36053614.Google Scholar
Whiting, P. J., Dietrich, W. E., 1993b. Experimental studies of bed topography and flow patterns in large-amplitude meanders 2. Mechanisms. Water Resources Research, 29(11), 36153622.Google Scholar
Whittaker, A. C., Boulton, S. J., 2012. Tectonic and climatic controls on knickpoint retreat rates and landscape response times. Journal of Geophysical Research – Earth Surface, 117. 10.1029/2011jf002157.Google Scholar
Whittaker, J. G., Jaeggi, M. N. R., 1982. Origin of step-pool systems in mountain streams. Journal of the Hydraulics Division – ASCE, 108(6), 758773.Google Scholar
Wiberg, P. L., Smith, J. D., 1987. Calculations of the critical shear stress for motion of uniform and heterogeneous sediments. Water Resources Research, 23(8), 14711480.Google Scholar
Wiberg, P. L., Smith, J. D., 1991. Velocity distribution and bed roughness in high gradient streams. Water Resources Research, 27(5), 825838.Google Scholar
Wicks, J. M., Bathurst, J. C., 1996. SHESED: a physically based, distributed erosion and sediment yield component for the SHE hydrological modelling system. Journal of Hydrology, 175(1–4), 213238.Google Scholar
Wilcock, P. R., 1996. Estimating local bed shear stress from velocity observations. Water Resources Research, 32(11), 33613366.Google Scholar
Wilcock, P. R., 1997. The components of fractional transport rate. Water Resources Research, 33(1), 247258.Google Scholar
Wilcock, P. R., 1998. Two-fraction model of initial sediment motion in gravel-bed rivers. Science, 280(5362), 410412.Google Scholar
Wilcock, P. R., 2012. Stream restoration in gravel-bed rivers. In: Church, M., Biron, P., Roy, A. (eds.), Gravel-Bed Rivers: Processes, Tools, Environments. Wiley, Chichester, UK, pp. 137146.Google Scholar
Wilcock, P. R., Crowe, J. C., 2003. Surface-based transport model for mixed-size sediment. Journal of Hydraulic Engineering, 129(2), 120128.Google Scholar
Wilcock, P. R., DeTemple, B. T., 2005. Persistence of armor layers in gravel-bed streams. Geophysical Research Letters, 32(8). 10.1029/2004gl021772.Google Scholar
Wilcock, P. R., Kenworthy, S. T., 2002. A two-fraction model for the transport of sand/gravel mixtures. Water Resources Research, 38(10). 10.1029/2001wr000684.Google Scholar
Wilcock, P. R., McArdell, B. W., 1993. Surface-based fractional transport rates – mobilization thresholds and partial transport of a sand-gravel mixture. Water Resources Research, 29(4), 12971312.Google Scholar
Wilcock, P. R., McArdell, B. W., 1997. Partial transport of a sand/gravel sediment. Water Resources Research, 33(1), 235245.Google Scholar
Wilcox, A. C., Wohl, E. E., 2006. Flow resistance dynamics in step-pool stream channels: 1. large woody debris and controls on total resistance. Water Resources Research, 42(5). 10.1029/2005wr004277.Google Scholar
Wilcox, A. C., Wohl, E. E., 2007. Field measurements of three-dimensional hydraulics in a step-pool channel. Geomorphology, 83(3–4), 215231.Google Scholar
Wilcox, A. C., Nelson, J. M., Wohl, E. E., 2006. Flow resistance dynamics in step-pool channels: 2. partitioning between grain, spill, and woody debris resistance. Water Resources Research, 42(5). 10.1029/2005wr004278.Google Scholar
Wilcox, A. C., Wohl, E. E., Comiti, F., Mao, L., 2011. Hydraulics, morphology, and energy dissipation in an alpine step-pool channel. Water Resources Research, 47. 10.1029/2010wr010192.Google Scholar
Wilkerson, G. V., Parker, G., 2011. Physical basis for quasi-universal relationships describing bankfull hydraulic geometry of sand-bed rivers. Journal of Hydraulic Engineering, 137(7), 739753.Google Scholar
Wilkinson, B. H., McElroy, B. J., 2007. The impact of humans on continental erosion and sedimentation. Geological Society of America Bulletin, 119(1–2), 140156.Google Scholar
Wilkinson, S. N., Keller, R. J., Rutherfurd, I. D., 2004. Phase-shifts in shear stress as an explanation for the maintenance of pool-riffle sequences. Earth Surface Processes and Landforms, 29(6), 737753.Google Scholar
Wilkinson, S. N., Prosser, I. P., Hughes, A. O., 2006. Predicting the distribution of bed material accumulation using river network sediment budgets. Water Resources Research, 42(10). 10.1029/2006wr004958.Google Scholar
Wilkinson, S. N., Henderson, A., Chen, Y., Sherman, B., 2008. SedNet User Guide. Client Report, CSIRO Land and Water Canberra.Google Scholar
Wilkinson, S. N., Prosser, I. P., Rustomji, P., Read, A. M., 2009a. Modelling and testing spatially distributed sediment budgets to relate erosion processes to sediment yields. Environmental Modelling & Software, 24(4), 489501.Google Scholar
Wilkinson, S. N., Wallbrink, P. J., Hancock, G. J., et al., 2009b. Fallout radionuclide tracers identify a switch in sediment sources and transport-limited sediment yield following wildfire in a eucalypt forest. Geomorphology, 110(3–4), 140151.Google Scholar
Wilkinson, S. N., Dougall, C., Kinsey-Henderson, A. E., et al., 2014. Development of a time-stepping sediment budget model for assessing land use impacts in large river basins. Science of the Total Environment, 468, 12101224.Google Scholar
Wilkinson, S. N., Kinsey-Henderson, A. E., Hawdon, A. A., et al., 2018. Grazing impacts on gully dynamics indicate approaches for gully erosion control in northeast Australia. Earth Surface Processes and Landforms, 43(8), 17111725.Google Scholar
Willemin, J. H., 2000. Hack’s law: sinuosity, convexity, elongation. Water Resources Research, 36(11), 33653374.Google Scholar
Willenbring, J. K., Codilean, A. T., McElroy, B., 2013. Earth is (mostly) flat: apportionment of the flux of continental sediment over millennial time scales. Geology, 41(3), 343346.Google Scholar
Willett, S. D., Brandon, M. T., 2002. On steady states in mountain belts. Geology, 30(2), 175178.Google Scholar
Willett, S. D., McCoy, S. W., Perron, J. T., Goren, L., Chen, C.-Y., 2014. Dynamic reorganization of river basins. Science, 343(6175). 10.1126/science.1248765.Google Scholar
Willgoose, G., 2005. Mathematical modeling of whole landscape evolution. Annual Review of Earth and Planetary Sciences, 33, 443459.Google Scholar
Willgoose, G., Bras, R. L., Rodriguez-Iturbe, I., 1991a. A coupled channel network growth and hillslope evolution model 1. theory. Water Resources Research, 27(7), 16711684.Google Scholar
Willgoose, G., Bras, R. L., Rodriguez-Iturbe, I., 1991b. A coupled channel network and hillslope evolution model 2. nondimensionalization and applications. Water Resources Research, 27(7), 16851696.Google Scholar
Willgoose, G., Bras, R. L., Rodriguez-Iturbe, I., 1991c. Results from a new model of river basin evolution. Earth Surface Processes and Landforms, 16(3), 237254.Google Scholar
Williams, G. P., 1978a. Bank-full discharge of rivers. Water Resources Research, 14(6), 11411154.Google Scholar
Williams, G. P., 1978b. Hydraulic geometry of river cross sections – theory of minimum variance. U.S. Geological Survey Professional Paper 1029. U.S. Government Printing Office, Washington, DC.Google Scholar
Williams, G. P., 1986. River meanders and channel size. Journal of Hydrology, 88, 147164.Google Scholar
Williams, G. P., 1988. Paleofluvial estimates from dimensions of former channels and meanders. In: Baker, V. R., Kochel, R. C., Patton, P. C. (eds.), Flood Geomorphology. Wiley, New York, pp. 321334.Google Scholar
Williams, G. P., Wolman, M. G., 1984. Downstream effects of dams on alluvial rivers. U.S. Geological Survey Professional Paper 1286. U.S. Government Printing Office, Washington, DC.Google Scholar
Williams, J., 1996. Turbulent flow in rivers. In: Carling, P. A., Dawson, M. R. (eds.), Advances in Fluvial Dynamics and Stratigraphy. Wiley, Chichester, UK, pp. 132.Google Scholar
Williams, J. R., Berndt, H. D., 1977. Sediment yield prediction based on watershed hydrology. Transactions of the American Society of Agricultural Engineers, 20(6), 11001104.Google Scholar
Williams, R. D., Brasington, J., Vericat, D., Hicks, D. M., 2014. Hyperscale terrain modelling of braided rivers: fusing mobile terrestrial laser scanning and optical bathymetric mapping. Earth Surface Processes and Landforms, 39(2), 167183.Google Scholar
Williams, R. D., Rennie, C. D., Brasington, J., Hicks, D. M., Vericat, D., 2015. Linking the spatial distribution of bed load transport to morphological change during high-flow events in a shallow braided river. Journal of Geophysical Research – Earth Surface, 120(3), 604622. 10.1002/2014jf003346.Google Scholar
Williams, R. D., Brasington, J., Hicks, D. M., 2016a. Numerical modelling of braided river morphodynamics: review and future challenges. Geography Compass, 10(3), 102127.Google Scholar
Williams, R. D., Measures, R., Hicks, D. M., Brasington, J., 2016b. Assessment of a numerical model to reproduce event-scale erosion and deposition distributions in a braided river. Water Resources Research, 52(8), 66216642.Google Scholar
Wilson, C. G., Papanicolaou, A. N. T., Denn, K. D., 2012. Partitioning fine sediment loads in a headwater system with intensive agriculture. Journal of Soils and Sediments, 12(6), 966981. 10.1007/s11368-012-0504-2.Google Scholar
Wilson, G., 2011. Understanding soil-pipe flow and its role in ephemeral gully erosion. Hydrological Processes, 25(15), 23542364.Google Scholar
Wilson, K. C., 1966. Bedload transport at high shear stresses. Journal of the Hydraulics Division – ASCE, 92, 4959.Google Scholar
Wilson, L., 1971. Drainage density, length ratios, and lithology in a glaciated area of southern Connecticut. Geological Society of America Bulletin, 82(10), 29552956.Google Scholar
Wilson, L., 1973. Variations in mean annual sediment yield as a function of mean annual precipitation. American Journal of Science, 273(4), 335349.Google Scholar
Winant, C. D., Browand, F. K., 1974. Vortex pairing: the mechanism of turbulent mixing layer growth at moderate Reynolds numbers. Journal of Fluid Mechanics, 63, 237255.Google Scholar
Winemiller, K. O., McIntyre, P. B., Castello, L., et al., 2016. Balancing hydropower and biodiversity in the Amazon, Congo, and Mekong. Science, 351(6269), 128129.Google Scholar
Wishart, D., Warburton, J., Bracken, L., 2008. Gravel extraction and planform change in a wandering gravel-bed river: the River Wear, Northern England. Geomorphology, 94(1–2), 131152.Google Scholar
Wobus, C., Whipple, K. X., Kirby, E., et al., 2006a. Tectonics from topography: procedures, promise, and pitfalls. In: Willett, S. D., Hovius, N., Brandon, M. T., Fisher, D. M. (eds.), Tectonics, Climate, and Landscape Evolution. Geological Society of America Special Papers, Special Paper 398. Geological Society of America, Denver, CO, pp. 5574.Google Scholar
Wobus, C. W., Crosby, B. T., Whipple, K. X., 2006b. Hanging valleys in fluvial systems: controls on occurrence and implications for landscape evolution. Journal of Geophysical Research – Earth Surface, 111(F2). 10.1029/2005jf000406.Google Scholar
Woelfle-Erskine, C., Wilcox, A. C., Moore, J. N., 2012. Combining historical and process perspectives to infer ranges of geomorphic variability and inform river restoration in a wandering gravel-bed river. Earth Surface Processes and Landforms, 37(12), 13021312.Google Scholar
Wohl, E., 2004. Limits of downstream hydraulic geometry. Geology, 32(10), 897900.Google Scholar
Wohl, E., 2008. Review of effects of large floods in resistant-boundary channels. In: Habersack, H., Piegay, H., Rinaldi, M. (eds.), Gravel-Bed Rivers VI: From Process Understanding to River Restoration. Elsevier, Amsterdam, the Netherlands, pp. 181212.Google Scholar
Wohl, E., 2011a. What should these rivers look like? Historical range of variability and human impacts in the Colorado Front Range, USA. Earth Surface Processes and Landforms, 36(10), 13781390.Google Scholar
Wohl, E., 2011b. Seeing the forest and the trees: wood in stream restoration in the Colorado Front Range, United States. In: Simon, A., Bennett, S. J., Castro, J. M. (eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools. American Geophysical Union, Washington, DC, pp. 399418.Google Scholar
Wohl, E., 2013a. Field and laboratory experiments in fluvial geomorphology. In: Shroder, J. W. (ed.), Treatise on Geomorphology, Vol. 9, Fluvial geomorphology, Wohl, E. (vol. ed.). Academic Press, San Diego, CA, pp. 679693.Google Scholar
Wohl, E., 2013b. Migration of channel heads following wildfire in the Colorado Front Range, USA. Earth Surface Processes and Landforms, 38(9), 10491053.Google Scholar
Wohl, E., 2014. A legacy of absence: wood removal in US rivers. Progress in Physical Geography, 38(5), 637663.Google Scholar
Wohl, E., 2015a. Legacy effects on sediments in river corridors. Earth-Science Reviews, 147, 3053.Google Scholar
Wohl, E., 2015b. Of wood and rivers: bridging the perception gap. Wiley Interdisciplinary Reviews – Water, 2(3), 167176.Google Scholar
Wohl, E., David, G. C. L., 2008. Consistency of scaling relations among bedrock and alluvial channels. Journal of Geophysical Research – Earth Surface, 113(F4). 10.1029/2008jf000989.Google Scholar
Wohl, E. E., Grodek, T., 1994. Channel bed-steps along Nahael Yael, Negev Desert, Israel. Geomorphology, 9(2), 117126.Google Scholar
Wohl, E., Legleiter, C. J., 2003. Controls on pool characteristics along a resistant-boundary channel. Journal of Geology, 111(1), 103114.Google Scholar
Wohl, E., Merritt, D. M., 2007. What is a natural river? Geography Compass, 1, 871900.Google Scholar
Wohl, E., Merritt, D. M., 2008. Reach-scale channel geometry of mountain streams. Geomorphology, 93(3–4), 168185.Google Scholar
Wohl, E., Scott, D. N., 2017. Wood and sediment storage and dynamics in river corridors. Earth Surface Processes and Landforms, 42(1), 523.Google Scholar
Wohl, E. E., Thompson, D. M., 2000. Velocity characteristics along a small step-pool channel. Earth Surface Processes and Landforms, 25(4), 353367.Google Scholar
Wohl, E. E., Vincent, K. R., Merritts, D. J., 1993. Pool and riffle characteristics in relation to channel gradient. Geomorphology, 6(2), 99110.Google Scholar
Wohl, E., Madsen, S., MacDonald, L., 1997. Characteristics of log and clast bed-steps in step-pool streams of northwestern Montana, USA. Geomorphology, 20(1–2), 110.Google Scholar
Wohl, E., Cenderelli, D., Mejia-Navarro, M., 2001. Channel change from extreme floods in bedrock canyons. In: Anthony, D. J., Harvey, M. D., Laronne, D., Mosley, M. P. (eds.), Applying Geomorphology to Environmental Management. Water Resources Publications, LLC, Highlands Ranch, CO, pp. 149174.Google Scholar
Wohl, E., Angermeier, P. L., Bledsoe, B., et al., 2005. River restoration. Water Resources Research, 41(10). 10.1029/2005wr003985.Google Scholar
Wohl, E., Lane, S. N., Wilcox, A. C., 2015. The science and practice of river restoration. Water Resources Research, 51(8), 59745997. 10.1002/2014wr016874.Google Scholar
Wohl, E., Lininger, K. B., Fox, M., Baillie, B. R., Erskine, W. D., 2017. Instream large wood loads across bioclimatic regions. Forest Ecology and Management, 404, 370380.Google Scholar
Wolman, M. G., 1955. The natural channel of Brandywine Creek Pennsylvania, U.S. Geological Survey Professional Paper 271. U.S. Government Printing Office, Washington, DC.Google Scholar
Wolman, M. G., 1967. A cycle of erosion and sedimentation in urban river channels. Geografiska Annaler Series A Physical Geography, 49, 385395.Google Scholar
Wolman, M. G., Gerson, R., 1978. Relative scales of time and effectiveness in watershed geomorphology. Earth Surface Processes and Landforms, 3(2), 189208.Google Scholar
Wolman, M. G., Leopold, L. B., 1957. River floodplains: some observations on their formation, U.S. Geological Survey Professional Paper 282-C. U.S. Government Printing Office, Washington, DC.Google Scholar
Wolman, M. G., Miller, J. P., 1960. Magnitude and frequency of forces in geomorphic processes. Journal of Geology, 68(1), 5474.Google Scholar
Wolman, M. G., Schick, A. P., 1967. Effects of construction on fluvial sediment, urban and suburban areas of Maryland. Water Resources Research, 3(2), 451464.Google Scholar
Wolter, A., Ward, B., Millard, T., 2010. Instability in eight sub-basins of the Chilliwack River Valley, British Columbia, Canada: a comparison of natural and logging-related landslides. Geomorphology, 120(3–4), 123132.Google Scholar
Womble, P., Doyle, M., 2012. The geography of trading ecosystem services: a case study of wetland and stream compensatory mitigation markets. Harvard Environmental Law Review, 36(1), 229296.Google Scholar
Wong, M., Parker, G., 2006. Reanalysis and correction of bed-load relation of Meyer-Peter and Muller using their own database. Journal of Hydraulic Engineering, 132(11), 11591168.Google Scholar
Wong, M., Parker, G., DeVries, P., Brown, T. M., Burges, S. J., 2007. Experiments on dispersion of tracer stones under lower-regime plane-bed equilibrium bed load transport. Water Resources Research, 43(3). 10.1029/2006wr005172.Google Scholar
Wood, A. L., Simon, A., Downs, P. W., Thorne, C. R., 2001. Bank-toe processes in incised channels: the role of apparent cohesion in the entrainment of failed bank materials. Hydrological Processes, 15(1), 3961.Google Scholar
Wood, P. J., Armitage, P. D., 1997. Biological effects of fine sediment in the lotic environment. Environmental Management, 21(2), 203217.Google Scholar
Woodget, A. S., Carbonneau, P. E., Visser, F., Maddock, I. P., 2015. Quantifying submerged fluvial topography using hyperspatial resolution UAS imagery and structure from motion photogrammetry. Earth Surface Processes and Landforms, 40(1), 4764.Google Scholar
Woodget, A. S., Austrums, R., Maddock, I. P., Habit, E., 2017. Drones and digital photogrammetry: from classifications to continuums for monitoring river habitat and hydromorphology. WIREs Water, 4(4), e1222.Google Scholar
Woodyer, K. D., 1968. Bankfull frequency in rivers. Journal of Hydrology, 6(2), 114142.Google Scholar
Wooldridge, C. L., Hickin, E. J., 2002. Step-pool and cascade morphology, Mosquito Creek, British Columbia: a test of four analytical techniques. Canadian Journal of Earth Sciences, 39(4), 493503.Google Scholar
Wooldridge, C. L., Hickin, E. J., 2005. Radar architecture and evolution of channel bars in wandering gravel-bed rivers: Fraser and Squamish rivers, British Columbia, Canada. Journal of Sedimentary Research, 75(5), 844860.Google Scholar
World Energy Council, 2019. Energy Resources: Hydropower. www.worldenergy.org/data/resources/resource/hydropower/ (accessed June 6, 2019).Google Scholar
Worrall, F., Burt, T. P., Howden, N. J. K., Hancock, G. R., 2014. Variation in suspended sediment yield across the UK – a failure of the concept and interpretation of the sediment delivery ratio. Journal of Hydrology, 519, 19851996.Google Scholar
Wren, D. G., Davidson, G. R., Walker, W. G., Galicki, S. J., 2008. The evolution of an oxbow lake in the Mississippi alluvial floodplain. Journal of Soil and Water Conservation, 63(3), 129135.Google Scholar
Wright, K. A., Sendek, K. H., Rice, R. M., Thomas, R. B., 1990. Logging effects on streamflow – storm runoff at Caspar Creek in northwestern California. Water Resources Research, 26(7), 16571667.Google Scholar
Wright, S., Parker, G., 2004. Density stratification effects in sand-bed rivers. Journal of Hydraulic Engineering – ASCE, 130(8), 783795.Google Scholar
Wright, S., Parker, G., 2005. Modeling downstream fining in sand-bed rivers. II: application. Journal of Hydraulic Research, 43(6), 621631.Google Scholar
Wu, B., Zheng, S., Thorne, C. R., 2012. A general framework for using the rate law to simulate morphological response to disturbance in the fluvial system. Progress in Physical Geography, 36(5), 575597.Google Scholar
Wyrick, J. R., Pasternack, G. B., 2008. Modeling energy dissipation and hydraulic jump regime responses to channel nonuniformity at river steps. Journal of Geophysical Research – Earth Surface, 113(F3). 10.1029/2007jf000873.Google Scholar
Wyrick, J. R., Pasternack, G. B., 2014. Geospatial organization of fluvial landforms in a gravel-cobble river: beyond the riffle-pool couplet. Geomorphology, 213, 4865.Google Scholar
Wyzga, B., 1996. Changes in the magnitude and transformation of flood waves subsequent to the channelization of the Raba River, Polish Carpathians. Earth Surface Processes and Landforms, 21(8), 749763.Google Scholar
Xia, J., Wu, B., Wang, G., Wang, Y., 2010. Estimation of bankfull discharge in the Lower Yellow River using different approaches. Geomorphology, 117(1–2), 6677.Google Scholar
Xu, J. X., 1996. Channel pattern change downstream from a reservoir: an example of wandering braided rivers. Geomorphology, 15(2), 147158.Google Scholar
Xu, J. X., 2002. Implication of relationships among suspended sediment size, water discharge and suspended sediment concentration: the Yellow River basin, China. Catena, 49(4), 289307.Google Scholar
Xu, J. X., 2004. Channel pattern discrimination based on the relationship between channel slope and width. Zeitschrift fur Geomorphologie, 48(3), 391401.Google Scholar
Xu, J., 2008. Discrimination of channel patterns for gravel-and sand-bed rivers. Zeitschrift fur Geomorphologie, 52(4), 503523.Google Scholar
Xu, J., Yan., Y., 2005. Scale effects on specific yield in the Yellow River basin and geomorphological explanations. Journal of Hydrology, 307, 219232.Google Scholar
Yaeger, M., Coopersmith, E., Ye, S., et al., 2012. Exploring the physical controls of regional patterns of flow duration curves – Part 4: a synthesis of empirical analysis, process modeling and catchment classification. Hydrology and Earth System Sciences, 16(11), 44834498.Google Scholar
Yager, E. M., Kirchner, J. W., Dietrich, W. E., 2007. Calculating bed load transport in steep boulder bed channels. Water Resources Research, 43(7). 10.1029/2006wr005432.Google Scholar
Yager, E. M., Dietrich, W. E., Kirchner, J. W., McArdell, B. W., 2012. Prediction of sediment transport in step-pool channels. Water Resources Research, 48. 10.1029/2011wr010829.Google Scholar
Yalin, M., 1971. On the formation of dunes and meanders, Proceedings of the 14th Congress, International Association of Hydraulic Research, 3, Paper C13, pp. 18.Google Scholar
Yalin, M. S., 2006. Large-scale turbulence and river morphology. In: Ferreira, R. M. L., Alves, C. T. L., Leal, G. A. B., Cardoso, A. H. (eds.), River Flow 2006, Vols 1 and 2, CRC Press, Boca Raton, FL, pp. 12431249.Google Scholar
Yalin, M. S., Da Silva, A. M. F., 2000. Computation of regime channel characteristics on thermodynamic basis. Journal of Hydraulic Research, 38(1), 5763.Google Scholar
Yalin, M. S., Karahan, E., 1979. Inception of sediment transport. Journal of the Hydraulics Division – ASCE, 105(11), 14331443.Google Scholar
Yan, Q., Iwasaki, T., Stumpf, A., et al., 2018. Hydrogeomorphological differentiation between floodplains and terraces. Earth Surface Processes and Landforms, 43(1), 218228.Google Scholar
Yang, C. T., 1971. On river meanders. Journal of Hydrology, 13, 231253.Google Scholar
Yang, C. T., 1973. Incipient motion and sediment transport. Journal of the Hydraulics Division – ASCE, 99, 16791704.Google Scholar
Yang, C. T., 1984. Unit stream power equation for gravel. Journal of Hydraulic Engineering, 110(12), 17831797.Google Scholar
Yang, C. T., 1994. Variational theories in hydrodynamics and hydraulics. Journal of Hydraulic Engineering, 120(6), 737756.Google Scholar
Yang, C. T., Song, C. C. S., 1979. Theory of minimum rate of energy dissipation. Journal of the Hydraulics Division – ASCE, 105(7), 769784.Google Scholar
Yang, C. T., Wan, S. G., 1991. Comparisons of selected bed-material load formulas. Journal of Hydraulic Engineering, 117(8), 973989.Google Scholar
Yang, C. T., Song, C. C. S., Woldenberg, M. J., 1981. Hydraulic geometry and minimum rate of energy dissipation. Water Resources Research, 17(4), 10141018.Google Scholar
Yang, K., Cao, S., Knight, D. W., 2007. Flow patterns in compound channels with vegetated floodplains. Journal of Hydraulic Engineering, 133(2), 148159.Google Scholar
Yang, Q. Y., Liu, T. H., Lu, W. Z., Wang, X. K., 2013. Numerical simulation of confluence flow in open channel with dynamic meshes techniques. Advances in Mechanical Engineering. 10.1155/2013/860431.Google Scholar
Yang, S. L., Xu, K. H., Milliman, J. D., Yang, H. F., Wu, C. S., 2015a. Decline of Yangtze River water and sediment discharge: impact from natural and anthropogenic changes. Scientific Reports, 5. 10.1038/srep12581.Google Scholar
Yang, S. Q., 2005. Sediment transport capacity in rivers. Journal of Hydraulic Research, 43(2), 131138.Google Scholar
Yang, S. Q., Tan, S. K., Lim, S. Y., 2004. Velocity distribution and dip-phenomenon in smooth uniform open channel flows. Journal of Hydraulic Engineering, 130(12), 11791186.Google Scholar
Yang, S. Q., Tan, S. K., Wang, X.-K., 2012. Mechanism of secondary currents in open channel flows. Journal of Geophysical Research – Earth Surface, 117. 10.1029/2012jf002510.Google Scholar
Yang, X. L., Zhang, Q. Y., Li, X. Z., et al., 2015b. Determination of soil texture by laser diffraction method. Soil Science Society of America Journal, 79(6), 15561566.Google Scholar
Yanites, B. J., 2018. The dynamics of channel slope, width, and sediment in actively eroding bedrock river systems. Journal of Geophysical Research – Earth Surface, 123(7), 15041527. 10.1029/2017jf004405.Google Scholar
Yanites, B. J., Tucker, G. E., 2010. Controls and limits on bedrock channel geometry. Journal of Geophysical Research – Earth Surface, 115(F4). 10.1029/2009jf001601.Google Scholar
Yatsu, E., 1955. On the longitudinal profile of the graded river. Transactions, American Geophysical Union, 36(4), 655663.Google Scholar
Yeakley, J. A., Ervin, D., Chang, H., et al., 2016. Ecosystem services of streams and rivers. In: Gilvear, D. J., Greenwood, M. T., Thoms, M. C., Woods, P. J. (eds.), River Science: Research and Management for the 21st Century. Wiley, Chichester, UK, pp. 335352.Google Scholar
Yeh, T. H., Parker, G., 2013. Software for evaluating sediment-induced stratification in open-channel flows. Computers & Geosciences, 53, 94104.Google Scholar
Yen, B. C., 2002. Open channel flow resistance. Journal of Hydraulic Engineering, 128(1), 2039.Google Scholar
Yochum, S. E., 2018. Guidance for Stream Restoration. Forest Service: National Stream & Aquatic Ecology Center Technical Note TN-102.4. U.S. Department of Agriculture, Ft. Collins, CO.Google Scholar
Yochum, S. E., Bledsoe, B. P., David, G. C. L., Wohl, E., 2012. Velocity prediction in high-gradient channels. Journal of Hydrology, 424, 8498.Google Scholar
Yochum, S. E., Comiti, F., Wohl, E., David, G. C. L., Mao, L., 2014. Photographic Guidance for Selecting Flow Resistance Coefficients in High-gradient Channels. Gen. Tech. Rep. RMRS-GTR-323. Forest Service, Rocky Mountain Research Station, U.S. Department of Agriculture, Ft. Collins, CO.Google Scholar
Yochum, S. E., Sholtes, J. S., Scott, J. A., Bledsoe, B. P., 2017. Stream power framework for predicting geomorphic change: the 2013 Colorado Front Range flood. Geomorphology, 292, 178192.Google Scholar
Yokoo, Y., Sivapalan, M., 2011. Towards reconstruction of the flow duration curve: development of a conceptual framework with a physical basis. Hydrology and Earth System Sciences, 15(9), 28052819.Google Scholar
Young, W. J., Davies, T. R. H., 1991. Bedload transport processes in a braided gravel-bed river model. Earth Surface Processes and Landforms, 16(6), 499511.Google Scholar
Yu, B., Wolman, M. G., 1987. Some dynamic aspects of river geometry. Water Resources Research, 23(3), 501509.Google Scholar
Yu, M., Rhoads, B. L., 2018. Floodplains as a source of fine sediment in grazed landscapes: tracing the source of suspended sediment in the headwaters of an intensively managed agricultural landscape. Geomorphology, 308, 278292.Google Scholar
Yuan, S., Tang, H., Xiao, Y., et al., 2016. Turbulent flow structure at a 90-degree open channel confluence: accounting for the distortion of the shear layer. Journal of Hydro-Environment Research, 12, 130147.Google Scholar
Yuan, S., Tang, H., Xiao, Y., Qui, X., Xia, Y., 2017. Water flow and sediment transport at open-channel confluences: an experimental study. Journal of Hydraulic Research, 56(3). 10.1080/00221686.2017.1354932.Google Scholar
Yuill, B. T., Khadka, A. K., Pereira, J., Allison, M. A., Meselhe, E. A., 2016a. Morphodynamics of the erosional phase of crevasse-splay evolution and implications for river sediment diversion function. Geomorphology, 259, 1229.Google Scholar
Yuill, B. T., Gaweesh, A., Allison, M. A., Meselhe, E. A., 2016b. Morphodynamic evolution of a lower Mississippi River channel bar after sand mining. Earth Surface Processes and Landforms, 41(4), 526542.Google Scholar
Zalasiewicz, J., Waters, C. N., Summerhayes, C. P., et al., 2017. The Working Group on the Anthropocene: summary of evidence and interim recommendations. Anthropocene, 19, 5560.Google Scholar
Zanoni, L., Gurnell, A., Drake, N., Surian, N., 2008. Island dynamics in a braided river from analysis of historical maps and air photographs. River Research and Applications, 24(8), 11411159.Google Scholar
Zarfl, C., Lumsdon, A. E., Berlekamp, J., Tydecks, L., Tockner, K., 2015. A global boom in hydropower dam construction. Aquatic Sciences, 77(1), 161170.Google Scholar
Zen, S., Zolezzi, G., Toffolon, M., Gurnell, A. M., 2016. Biomorphodynamic modelling of inner bank advance in migrating meander bends. Advances in Water Resources, 93, 166181.Google Scholar
Zhang, C., Xu, M., Hassan, M. A., Chartrand, S. M., Wang, Z., 2018. Experimental study on the stability and failure of individual step-pool. Geomorphology, 311, 5162.Google Scholar
Zhang, L. T., Li, Z. B., Wang, S. S., 2016b. Spatial scale effect on sediment dynamics in basin-wide floods within a typical agro-watershed: a case study in the hilly loess region of the Chinese Loess Plateau. Science of the Total Environment, 572, 476486.Google Scholar
Zhang, Y., Slingerland, R., Duffy, C., 2016a. Fully-coupled hydrologic processes for modeling landscape evolution. Environmental Modelling & Software, 82, 89107.Google Scholar
Zhou, Z., Coco, G., Townend, I., et al., 2017. Is “Morphodynamic Equilibrium” an oxymoron? Earth-Science Reviews, 165, 257267.Google Scholar
Ziliani, L., Surian, N., Coulthard, T. J., Tarantola, S., 2013. Reduced-complexity modeling of braided rivers: assessing model performance by sensitivity analysis, calibration, and validation. Journal of Geophysical Research – Earth Surface, 118(4), 22432262. 10.1002/jgrf.20154.Google Scholar
Zimmermann, A. E., 2013. Step-pool channel features. In: Shroder, J. (ed.), Treatise on Geomorphology, Vol. 9, Fluvial Geomorphology, Wohl, E. (vol. ed.). Academic Press, San Diego, CA, pp. 346363.Google Scholar
Zimmermann, A. E., Church, M., 2001. Channel morphology, gradient profiles and bed stresses during flood in a step-pool channel. Geomorphology, 40(3–4), 311327.Google Scholar
Zimmermann, A. E., Church, M., Hassan, M. A., 2008. Identification of steps and pools from stream longitudinal profile data. Geomorphology, 102(3–4), 395406.Google Scholar
Zimmermann, A. E., Church, M., Hassan, M. A., 2010. Step-pool stability: testing the jammed state hypothesis. Journal of Geophysical Research – Earth Surface, 115. 10.1029/2009jf001365.Google Scholar
Zinger, J. A., Rhoads, B. L., Best, J. L., 2011. Extreme sediment pulses generated by bend cutoffs along a large meandering river. Nature Geoscience, 4(10), 675678.Google Scholar
Zinger, J. A., Rhoads, B. L., Best, J. L., Johnson, K. K., 2013. Flow structure and channel morphodynamics of meander bend chute cutoffs: a case study of the Wabash River, USA. Journal of Geophysical Research – Earth Surface, 118(4), 24682487. 10.1002/jgrf.20155.Google Scholar
Zingg, T., 1935. Beitrage zur Schotteranalyse. Schweizerische Mineralogische und Petrologische Mitteilungen, 15, 39140.Google Scholar
Zolezzi, G., Seminara, G., 2001. Downstream and upstream influence in river meandering. Part 1. General theory and application to overdeepening. Journal of Fluid Mechanics, 438, 183211.Google Scholar
Zolezzi, G., Bertoldi, W., Tubino, M., 2006. Morphological analysis and prediction of river bifurcations. In: Smith, G. H. S., Best, J. L., Bristow, C. S., Petts, G. E. (eds.), Braided Rivers: Process, Deposits, Ecology and Management. Special Publications of the International Association of Sedimentologists, 36, Blackwell, Oxford, UK, pp. 233256.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org 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
×